1. Introduction
The advancement of battery electric vehicles (BEVs) has become central to the transformation of the automotive industry. BEVs offer significant advantages over traditional internal combustion engine (ICE) vehicles, including the elimination of tailpipe greenhouse gas emissions, improved energy conversion efficiency, and a reduced number of mechanical parts, contributing to lower maintenance requirements [
1]. Among various energy storage technologies, lithium–ion batteries have emerged as the dominant power source for new energy vehicles. Since their initial commercialization by Sony Corporation in 1990, lithium–ion batteries have played a crucial role in shaping the development of BEVs [
2].
Despite the rapid progress in battery technologies, the mass adoption of BEVs is currently experiencing stagnation. This deceleration, often described in terms of the “chasm” phenomenon, has been linked to persistent technical challenges, most notably limited driving range and long charging times [
3]. These constraints remain primary concerns for consumers and have hindered broader market penetration. Addressing such challenges necessitates not only improvements in electrochemical performance but also the development of high-power fast-charging capabilities, often exceeding 300 kW. However, increasing energy density and rapid charging rates are directly associated with higher heat generation, thereby intensifying the need for robust BTMS.
Various BTMS technologies have been proposed and implemented to maintain safe and efficient operating temperatures. These include air cooling, direct and indirect liquid cooling, phase change material (PCM)-based methods, and heat pipe-assisted approaches. Each configuration offers distinct advantages and is subject to specific limitations in terms of thermal conductivity, system integration, reliability, and cost.
Parallel to advancements in thermal control technologies, the evolution of battery cell formats has significantly influenced BEV design and performance. The industry has progressed from the 18650 cylindrical format, known for its manufacturing maturity and reliability, to the 2170 format, which provides improved capacity and thermal behavior. More recently, the 4680 cylindrical cell has gained attention due to its increased energy density, tabless internal structure, and reduced internal resistance [
4]. While this format offers improved volumetric efficiency and performance, it also introduces substantial thermal management challenges, owing to its larger thermal mass and greater internal heat generation.
Thermal regulation of 4680 format cells has become an increasingly important research topic. In experimental investigations, Li et al. [
5] observed that during 1.5 C discharge, the surface temperature of 4680 cells exceeded 70 °C, despite the use of tabless design and top and bottom indirect liquid cooling. This suggests that even with architectural innovations, conventional BTMS strategies are not sufficient to handle the thermal loads of large-format cells. Similarly, Baazouzi et al. [
4] reported limited thermal improvements in 4680 cells compared to 2170 cells under high load conditions, based on teardown analysis. To explore possible improvements in cooling strategies, Eze et al. [
6] performed a coupled electrochemical–thermal simulation to evaluate the effectiveness of different cooling approaches. Their findings indicated that a combined top and bottom cooling method is more suitable for large-format cylindrical cells. However, they also noted that the analysis was conducted under ideal thermal boundary conditions and lacked experimental validation. Furthermore, practical integration aspects, including module design constraints and coolant routing, remain underexplored.
The aforementioned studies collectively reveal a critical limitation in existing thermal management practices. While the 4680 format improves energy density and enables simplified module design, its substantial internal heat generation and nonuniform thermal distribution pose serious challenges for maintaining thermal stability. In particular, the literature rarely addresses important design factors such as wick structure optimization, thermal interface design, or the effective use of internal chamber geometry. Moreover, experimental validation of BTMS strategies tailored specifically for the 4680 format remains insufficient.
To address these gaps, the present study introduces a novel hybrid BTMS that combines heat pipe-based thermal transport with two-phase immersion cooling. The proposed configuration incorporates a capillary wick structure that actively facilitates liquid return within the heat pipe through capillary action. This mechanism enables efficient two-phase heat transfer using a small volume of working fluid and enhances the thermal uniformity of the system. Unlike conventional immersion cooling systems, which rely solely on natural convection or PCM integration, this approach allows for passive circulation without the need for mechanical pumps. The overall result is improved thermal performance and reduced system weight, which are both essential for high-power density applications such as BEV battery modules.
In this study, the proposed system is evaluated through both experimental investigations and multiphysics simulations. Key parameters analyzed include thermal behavior under high-rate charging conditions, the influence of wick structure configuration, and the effects of varying filling ratios. The integrated approach aims to establish a scalable and reliable solution that addresses the unique thermal demands of large-format cylindrical lithium–ion cells.
Understanding the operational thermal limits of lithium–ion batteries is essential for evaluating the significance of the proposed BTMS design. These batteries typically operate within a functional range of minus 20 to 60 °C and have an optimal temperature window between 15 and 35 °C [
7,
8]. Exceeding these boundaries can result in reduced charge efficiency, capacity degradation, and, in extreme cases, thermal runaway, a highly dangerous and irreversible failure process [
9]. Tesla’s 4680 battery, for example, exhibits an average temperature of 67.77 °C and a thermal gradient of 2.64 °C under a 1.5 C discharge rate with top and bottom cooling. In contrast, the 2170 cell under the same conditions reaches 86.84 °C and exhibits a much steeper thermal gradient of 20.92 °C [
5].
These results demonstrate the superior capacity, output performance, and thermal stability of the 4680 battery compared to the commercialized 2170 cell. However, even at a 1.5 C discharge rate, the 4680 cell still exceeds the upper allowable temperature limit of 60 °C. This suggests that during fast charging at rates of 2 C or higher, excessive heat generation may lead not only to reduced charging efficiency and capacity loss but also to an increased risk of thermal runaway [
5].
Air cooling, one of the earliest and most widely adopted BTMS methods, is appreciated for its simplicity, light weight, and cost-effectiveness [
10]. However, due to the low thermal conductivity and specific heat capacity of air, this approach often results in uneven temperature distribution and insufficient cooling performance [
11]. Several enhancements, such as pressure-relieved ventilation [
12], parallel airflow design [
13], reciprocating flow configurations [
14], and specialized heat sink structure [
15], have been investigated to improve heat dissipation. Although hybrid air cooling systems combining series and parallel flow paths show potential [
16], air cooling still suffers from critical limitations when applied to high-density battery systems. In addition, the ingress of airborne particulates can further degrade cooling performance and increase thermal resistance, as shown by Feng et al. [
15].
To overcome these limitations, liquid cooling has been increasingly adopted in modern BTMS designs. Direct liquid cooling offers excellent thermal performance but presents challenges in ensuring uniform coolant distribution and avoiding leakage [
12]. Indirect methods, which use cooling plates and low-viscosity fluids to transfer heat without direct contact with battery cells, are more practical for modular integration [
17,
18,
19]. Nevertheless, issues such as thermal contact resistance, increased component weight, and higher installation costs remain [
20,
21]. The use of nanofluids containing aluminum oxide or titanium dioxide particles has been proposed to improve the thermal conductivity of liquid coolants [
22,
23], and structural optimizations, such as serpentine cooling channels, have also been explored [
24]. For example, Ahmad et al. [
15] investigate the performance of a microchannel heat sink enhanced with sidewall ribs to improve heat removal under single-phase liquid cooling, which is a relevant indirect cooling strategy for high-power density systems such as electric vehicle battery packs.
PCM-based BTMS approaches have attracted attention for their ability to regulate temperature passively through latent heat absorption and release during phase transitions [
11,
25]. However, the inherently low thermal conductivity of PCMs, along with volume changes during melting and solidification, limits their overall effectiveness [
26,
27,
28,
29]. To mitigate these issues, composite structures incorporating metal foam or expanded graphite have been proposed, and hybrid PCM–heat pipe systems have been introduced to improve thermal response [
30,
31].
Heat pipe systems provide high thermal conductivity and passive operation without external energy input. They typically consist of evaporation, adiabatic, and condensation regions, allowing working fluid to circulate through evaporation and condensation cycles. These systems offer promising potential for regulating temperature in battery modules [
30,
32,
33]. Modular implementations have demonstrated reduced peak temperatures [
34] and improved thermal uniformity through enhanced surface contact designs [
35]. However, heat pipes face limitations related to limited cooling capacity, small contact area, and manufacturing cost, especially when copper materials are used [
36].
Two-phase immersion cooling using electrically insulating fluids offers another promising approach. By directly submerging cells in dielectric fluids, this method achieves efficient thermal regulation and reduces the risk of thermal runaway [
30]. When combined with heat pipes, immersion cooling further improves thermal performance while eliminating the need for active pumping. For high-energy density batteries such as the 4680 cell, where thermal buildup under fast charging can surpass 70 °C [
18], such an integrated solution can play a vital role in maintaining safe and efficient operation.
In consideration of the aforementioned issues, the present study proposes and experimentally validates a hybrid thermal management strategy that integrates heat pipes with two-phase immersion cooling, as depicted in
Figure 1. This figure presents the proposed cooling configuration specifically designed for the 4680-format cylindrical battery cell. In this system, the battery pack is vacuum-sealed, and a working fluid is introduced to form a closed-loop structure that operates analogously to a heat pipe with an internal heat source. During the charge and discharge processes, heat generated by the battery is effectively transported to the upper cooling plate through the continuous evaporation and condensation of the working fluid. This mechanism significantly reduces thermal resistance between the battery surface and the cooling interface.
Unlike conventional immersion cooling systems, which rely on mechanical circulation of the coolant to ensure contact across the entire cell surface, the proposed method utilizes a smaller quantity of working fluid while simplifying the internal flow path of the cooling plate. This also reduces the power requirements of auxiliary components such as pumps. Accordingly, the objective of this study is to develop a compact, passively operated, and thermally efficient battery thermal management system that is well-suited for next-generation cylindrical lithium–ion batteries in electric vehicles and other high-power applications.
3. 1D Simulation
3.1. Objective and Overview of the Simulation
To systematically compare and analyze the heat transfer performance differences observed experimentally depending on wick structure and working fluid filling ratio, a one-dimensional (1D) simulation model was developed based on the same structural configuration. This model reflects the physical structure of the experimental setup and functionally implements the major components, including the heat source, wick, and working fluid. The primary focus was to predict the temperature response within the heater block under various heat input conditions. Through this approach, quantitative comparison with experimental results becomes feasible, and the thermal behavior trends according to wick design can be assessed via simulation. This provides foundational insight for both structural design and performance evaluation.
3.2. Theoretical Background of 1D Modeling
The simulation was conducted using Siemens Simcenter AMESim (ver. 2310) based on a one-dimensional model. Thermodynamic variables, such as pressure and specific enthalpy, were used in the simulation process, and by inputting these into AMESim, the software automatically calculated the related physical properties and output results [
47]. To simplify the analysis and focus on key parameters, several assumptions were adopted in constructing the simulation model [
48].
First, the wick structure was assumed to be fully saturated with the working fluid, implying complete contact between the wick and the fluid throughout the entire domain.
Second, the internal flow within the wick was considered to follow Darcy’s law, allowing for simplified treatment of fluid flow through porous media. Third, the concept of effective thermal conductivity was applied to the wick structure to enable more accurate heat transfer analysis.
Lastly, the filling ratio was calculated based on the ratio of the area occupied by the working fluid to the total cross-sectional area of the internal space. The interior of the wick is assumed to be fully saturated with the working fluid, and its internal structure is considered to follow a packed spherical particle model based on SEM images [
44].
This full saturation assumption is supported by experimental observations: when the wick was placed in contact with the working fluid, complete saturation occurred over time due to capillary action, as the wick consists of a hydrophilic porous medium. As the wick is a composite structure comprising both solid and fluid phases, directly modeling its microscopic pore structure poses significant challenges. Therefore, it was treated as a homogeneous equivalent material, and the effective thermal conductivity was calculated using the Maxwell–Eucken model.
The thermal properties of the wick structures—such as effective thermal conductivity, effective density, and specific heat—were then determined based on the material properties of the solid and working fluid, as well as the porosity.
Appendix A presents all relevant equations, including those used to calculate the volumes of the working fluid and the vapor passage, both of which were derived from cross-sectional area calculations. In addition, the hydraulic diameter was calculated based on the tortuosity values of the wick structures, which were estimated from SEM image analysis. The calculated thermal properties and the hydraulic diameters for each wick structure are summarized in
Table 4 [
49,
50].
Implementation of Heat Transfer Modeling in AMESim
In this study, a heat transfer model based on heat transfer coefficients and thermal resistance was implemented within the AMESim environment. The built-in thermal–fluid analysis module in AMESim automatically calculates flow states and heat transfer coefficients based on user-defined physical parameters, such as temperature, pressure, and flow rate, and simulates the overall thermal behavior of the system through a thermal resistance network across different regions.
Using Equations (6) and (7), the latent heat
of the system can be calculated from the saturated vapor enthalpy
and saturated liquid enthalpy
:
This value is a key factor in determining the fluid flow within the heat pipe, as internal flow is computed using the correlation between mass flow rate
, supplied heat Q, and latent heat
:
In AMESim, conductive heat transfer between components is governed by predefined theoretical expressions, and the conduction is calculated based on Equation (8) [
47]:
The heat transfer Q within the working fluid is represented by Newton’s law of cooling, as described in Equation (9):
where
is the surface temperature,
is the ambient temperature,
is the convective heat transfer coefficient, and
is the heat transfer area [
51].
The convective heat transfer coefficient
is derived from the Nusselt number Nu, which characterizes convective heat transfer performance. Nu is calculated based on Equation (10), incorporating the thermal conductivity k and characteristic length
. To obtain Nu, dimensionless numbers such as the Reynolds number Re, Prandtl number Pr, and Grashof number Gr are required, which are defined by Equations (11)–(13), respectively. These numbers reflect the effects of flow inertia, viscosity, and buoyancy on the convection process [
47]:
In the single-phase turbulent flow region, the Nusselt number is calculated using the Gnielinski correlation, as shown in Equation (14) [
47]:
Under two-phase flow conditions, the convective heat transfer coefficient in the condensation region is defined in the simulation based on Equation (15) [
47]:
where
represents the convective heat transfer coefficient for the fully liquid state of the working fluid and is calculated by Equation (16) [
47]:
In the boiling region, the convective heat transfer coefficient is defined by Equations (17) and (18), reflecting the heat transfer characteristics in the liquid–vapor mixture region [
47]:
The terms
and
, which quantify heat transfer characteristics under two-phase flow conditions, are calculated using Equations (19) and (20) [
47]:
The variables
and
are calculated using Equations (21) and (22) [
47]:
Additionally,
is calculated by Equation (23), and
is derived using Equation (24) [
47]:
The variable
in the function
in Equation (n) represents the molar mass of the working fluid. Constants such as
,
,
,
, and
are determined based on the properties of the working fluid. These parameters are incorporated into the relevant equations used to calculate
within the AMESim software environment [
47].
Table 5 describes the initial and boundary conditions used to develop the AMESim software model in this study. The material properties of SUS304 were used for the battery, as it matches the material used in both the actual battery and the experimental apparatus, thereby enhancing the reliability of the simulation results. Each porosity value was calculated under the assumption that the internal structure of the wick follows a packed spherical particle configuration, based on SEM images of the wicks used in the experiments [
39]. Water was used as the coolant in the cooling plate, with a maintained temperature of 20 °C and a flow rate of 0.5 L/min. Additionally, as in the experiment, filling ratios ranging from 30% to 70% were considered, and heat loads of 15 W, 50 W, and 85 W were applied based on the heat generation rates corresponding to 1 C, 2 C, and 3 C charging rates of the Tesla 4680 battery cells [
38].
3.3. Initial Modeling Parameters
Simulation Workflow and Modeling Logic in AMESim
As shown in
Figure 10, the simulation system was modeled based on the Wick 2 structure used in the experiment.
The entire system consists of three key components: a heater block, a wick module, and a capillary circulation unit, as detailed in
Appendix B [
47]. The heater block and wick structure are represented by red and yellow boxes, respectively, and were designed to allow physical interaction within the simulated environment that reflects the experimental setup. The blue box in the center represents the capillary system, which plays a crucial role in sustaining the circulation of the working fluid. The mass flow rate was calculated by converting the average pressure of the vapor and liquid lines using a thermodynamic state converter component. The resulting pressure information was then transformed into specific enthalpy values, which were mapped to the fully evaporated state (x = 1) and the saturated liquid state (x = 0). These calculations were executed according to the flow chart-based system constructed within AMESim to simulate the behavior of the working fluid, as illustrated in
Figure 11.
A constant heat input ranging from 15 to 85 W was applied to each of the three heater blocks. Each block was vertically divided into three equal sections to allow for precise measurement of temperature distribution along the height.
As shown in
Figure 12, the wick module was constructed by vertically stacking three layers of wick and enclosing them within an outer casing. This configuration ensures that the working fluid is uniformly distributed across the entire wick through a series connection. This modular design was optimized based on the physical characteristics observed in the experiment, while also maintaining high consistency in implementation and analysis within the simulation environment.
In particular, the cylindrical wick structure was tightly wrapped around the inner wall of the circular heater block, allowing for uniform heat transfer from both directions. Additionally, a cooling plate was positioned at the top of the system to extract heat from the working fluid. This plate enables the condensation of vapor-phase fluid through its cooled surface, thereby stabilizing the thermal circulation cycle throughout the system. In other words, the vaporized fluid ascends along the vapor line and condenses at the cooling plate, thus realizing a heat pipe-based thermal exchange mechanism.
This configuration accurately reflects the coupled thermal and fluid behavior observed during the experiment and is well-suited for effectively simulating the actual heat transfer flow within the system.
3.4. Validation of Simulation Model
To evaluate the reliability of the simulation results, the difference between the simulated and experimental maximum surface temperatures of the battery, one of the most critical performance indicators in this study, was analyzed. This temperature difference was defined as .
To minimize discrepancies between the simulation and experimental results, a correction factor was applied to the heat transfer gain coefficient of the working fluid. Instead of referencing existing literature values, this study directly derived the correction factors based on experimental data [
32,
33].
First, separate correction factors were obtained independently for wickless and wicked (Wick 1–5) structures. For the wickless structure, which has a single configuration, the correction factor was calculated individually for each filling ratio by minimizing the error between the simulation and experimental results. The final correction factor of 1.014 was determined as the average of these values.
For the wicked structures (Wick 1–5), correction factors were similarly derived for each filling ratio and then averaged for each wick type. To ensure consistency across different wick geometries, the average correction factors of all wicked structures were further averaged, resulting in a unified correction factor of 1.2872, which was commonly applied to all wicked configurations.
By consistently applying these correction factors, the simulation model achieved reliable quantitative comparisons across various wick designs. As shown in
Figure 13, the temperature differences between simulation and experiment were mostly within ±10 °C under all conditions, with some cases showing high agreement within ±1 °C.
Such error margins are generally considered acceptable in heat pipe modeling, and the results obtained after applying the correction factors contributed to enhancing the predictive accuracy and reliability of the simulation model. For instance, Lee et al. [
32] reported a 6.5% error in their heat pipe modeling study but considered it acceptable due to the physical trend agreement between simulation and experiment. Therefore, the present model can also be considered sufficiently reliable in terms of prediction accuracy.
In the case of the wickless structure, a generally negative was observed under the 15 W condition. At 50 W and 85 W, gradually transitioned from positive to negative as the filling ratio increased. Particularly at a filling ratio of 70%, large negative values were found across all heat load conditions. This is attributed to the fact that in the experiment, the hydrothermal siphon effect is activated around a 60% filling ratio, leading to superior cooling performance. In contrast, the AMESim simulation primarily reflects cooling effects driven by total latent heat rather than detailed evaporation and condensation behavior, resulting in underprediction compared to actual experimental outcomes.
From the comprehensive analysis of simulation results for Wick 1 through 5 structures, the at 50% filling ratio generally converged to within ±5 °C across all heat loads. This aligns well with the 40–50% range in which the system showed optimal performance experimentally, indicating that the simulation model can reasonably predict thermal behavior under these conditions. However, some exceptional deviations were observed. For example, Wick 2 and Wick 3 showed abnormally high values at 85 W under 40% filling, likely due to limitations in AMESim’s ability to accurately capture the micro-pore geometry and capillary behavior of the actual wick structures. These findings suggest that in wick configurations with complex internal geometries, simulation accuracy cannot be fully ensured using only a generalized correction factor, and structure-specific thermal modeling is required.
As shown in
Figure 14, the simulation results with correction factors applied exhibited a consistent trend of decreasing maximum temperature as the filling ratio increased across all conditions. This trend can be explained by the increase in total working fluid mass in the system, which in turn increases the amount of latent heat absorbed during the evaporation–condensation process. This enhanced latent heat buffering effect contributes to limiting the maximum surface temperature rise of the battery under the same heat load conditions. However, when the working fluid volume exceeds the experimentally verified optimal range, the heat pipe fails to operate in a stable manner. Therefore, filling ratios beyond that range should be considered outside the valid operating range.
These results demonstrate that the simulation model can significantly improve its agreement with experimental data through the application of correction factors. In particular, the model was enhanced to a degree that it can reflect the complex micro-pore characteristics of wick structures to some extent. However, in certain conditions, deviations exceeding ±10 °C still occurred. This is likely due to simplifications within the numerical model regarding physical variables such as the internal pore geometry of the wick, working fluid distribution, and initial wetting conditions. Therefore, the adoption of advanced modeling techniques that can more precisely capture these microscopic characteristics is needed in future work. In particular, beyond simply incorporating wick structures based on porosity and tortuosity, it is necessary to introduce advanced submodels that accurately describe the internal structure without simplification, thereby enabling a more precise representation of capillary phenomena.
Additionally,
Figure 15 compares the thermal resistance results between the experimental and simulation data under varying filling ratios at a heat input of 85 W. In this study, thermal resistance was calculated based on the transient temperature difference between the coolant and the heat block. According to the experimental results, at 15 W, the variation in thermal resistance with respect to the filling ratio was too small to derive any meaningful trend, while at 50 W, only a relatively linear increase was observed. In contrast, the 85 W condition exhibited the largest variation in thermal resistance across filling ratios, indicating that the heat pipe was operating most actively under this condition. Therefore, the analysis focused on the 85 W condition. Wick 5 was selected as the simulation target structure because it exhibited the best cooling performance in the experimental results. Under the 85 W condition, the AMESim simulation results for Wick 5 showed thermal resistance levels similar to those of the experiment in the 30–50% filling ratio range, suggesting high predictive reliability of the simulation model in this region.
However, in the 60–70% filling ratio range, the trends in thermal resistance between simulation and experiment diverged. This discrepancy aligns with the previously discussed analysis and is considered to stem from the tendency of the AMESim model to overestimate the cooling performance enhancement due to increased heat capacity with higher working fluid content, while underestimating the actual degradation in heat pipe operation under high filling conditions.
3.5. One-Dimensional Simulation Results and Discussion
Figure 16 presents a comprehensive summary of the sensitivity analysis and predictive simulations conducted in this study. The purpose of this analysis was to investigate the thermal behavior of the system under conditions constrained by experimental limitations and to predict the system’s performance across different design configurations. By combining sensitivity assessment with predictive simulations of future design scenarios, this study identifies the relative influence of each design variable and provides practical insights for optimizing wick structures and improving cooling system performance.
Among various configurations, the Wick 5 structure with a 30% filling ratio was selected as the baseline for this analysis, as it demonstrated the best cooling performance during experiments. This condition was thus considered the most significant and meaningful for further predictive evaluation.
Figure 16a illustrates the change in the maximum surface temperature of the Wick 5 structure under different coolant temperatures (10 °C, 20 °C, and 30 °C), with all other conditions held constant. As the coolant temperature decreased, the maximum system temperature also declined.
Taking 20 °C as the reference condition, the maximum surface temperature at 85 W was 58.1 °C. When the coolant temperature was lowered to 10 °C, the temperature decreased to 52.5 °C, resulting in a 5.6 °C reduction or approximately 9.6% improvement. Conversely, increasing the coolant temperature to 30 °C led to a rise in maximum temperature to 64.5 °C, representing a 6.4 °C increase or approximately 11.0% degradation compared to the baseline. This trend is attributed to enhanced condensation performance and an increased thermal gradient, which more effectively drives the evaporation and condensation cycle of the working fluid, thereby demonstrating that coolant temperature is a highly sensitive parameter influencing the system’s thermal performance.
Figure 16b shows the effect of wick thickness (2.5 mm, 3.0 mm, and 3.5 mm). As the thickness increased, the maximum temperature gradually decreased. Using the 3.0 mm configuration as the reference, the maximum surface temperature at 85 W was 58.1 °C. When the wick thickness was reduced to 2.5 mm, the temperature increased to 58.7 °C, showing a 0.6 °C rise or approximately 1.0% degradation. Conversely, increasing the wick thickness to 3.5 mm resulted in a temperature of 57.4 °C, indicating a 0.7 °C decrease or approximately 1.2% improvement. This trend can be explained by the increased pore volume in thicker wicks, which allows more working fluid to be stored, thereby enhancing latent heat absorption and promoting more distributed heat dissipation. Overall, the findings suggest that while wick thickness influences thermal behavior, its impact is moderate compared to other design variables.
Initially, simulations were conducted by varying only the porosity of the wick structure. However, this resulted in negligible differences in maximum temperature. Consequently, a more representative parameter of internal structural complexity—tortuosity—was varied to perform additional analysis.
Figure 16c presents the effect of tortuosity variation on the maximum surface temperature. As the tortuosity increased from 1.2 to 1.52 and 1.8, the maximum temperature consistently rose. Using 1.52 as the reference, the maximum surface temperature at 85 W was 58.3 °C. When the tortuosity was reduced to 1.2, the temperature slightly decreased to 57.9 °C, indicating a 0.4 °C reduction or approximately 0.7% improvement. In contrast, increasing the tortuosity to 1.8 resulted in a temperature of 58.6 °C, showing a 0.3 °C increase or approximately 0.5% degradation. These results indicate that tortuosity has a measurable but limited impact on thermal performance. The trend is attributed to increased internal flow resistance in more complex structures, which impedes the movement of working fluid and reduces heat transfer efficiency.
Figure 16d presents simulation results for different wick materials (blended fabric, aluminum, and SUS304). Although these materials possess distinct effective thermal conductivity, density, and specific heat, the simulation results showed minimal differences. This implies that, within the AMESim model, cooling performance is more significantly influenced by the behavior of the working fluid circulating through the wick rather than the wick material itself.
This analysis extends beyond a basic sensitivity evaluation by offering predictive insights into operating conditions that are difficult to replicate through experiments. It provides practical direction for optimizing wick structures and improving cooling system performance. Based on the results, the filling ratio of the working fluid and the geometric configuration of the wick emerged as the most critical parameters affecting the system’s thermal response. The simulations also confirmed that coolant temperature is a highly influential factor. Reducing the coolant temperature from 20 °C to 10 °C resulted in a 5.6 °C drop in surface temperature, corresponding to a performance improvement of approximately 9.6%. This indicates that adjusting coolant temperature is the most effective approach among the variables studied. Changes in wick thickness and tortuosity produced improvements of approximately 1.2% and 0.7%, respectively, suggesting that their thermal impact is moderate but meaningful. In contrast, material selection for the wick demonstrated relatively minor influence on the overall thermal performance. These results show that the simulation model can reliably predict system behavior across design variations and can be used to guide future design efforts toward configurations that maximize thermal efficiency under constrained conditions.
4. CFD Simulation
To evaluate the thermal management performance of a 4680 cylindrical battery, a CFD simulation was conducted and analyzed by comparing the results with experimental data. The simulation system is based on the finite volume method and solves governing equations, including the continuity, momentum, and energy equations, while incorporating multiphase flow and phase-change models. In this study, ANSYS Fluent V24.2.0 was employed to numerically analyze the internal thermal–fluid characteristics of a heat pipe integrated cooling system. The analysis was carried out under the following conditions and assumptions to enable comparison between the simulated temperature values and the experimental results [
52,
53,
54].
The vapor phase was assumed to be turbulent, and the Realizable k–ε model—commonly used in multiphase flow—was applied [
55].
Based on the experimentally observed boiling point, the saturation temperature was back-calculated, and the internal pressure was set to approximately 10,000 Pa. Accordingly, the saturation temperature was set to 318.7 K.
A low-pressure rarefied air region was assumed to occupy the space not filled by the vapor or liquid phases within the domain.
The system was assumed to be symmetric with respect to a single cell.
All components of the system, except for the heating and cooling sections, were assumed to be adiabatic.
4.1. Governing Equations
This section describes the key numerical theories applied in the analysis and explains how these theories are implemented within ANSYS Fluent.
4.1.1. Continuity Equation
The continuity equation is based on the principle of mass conservation, which states that the mass of a fluid remains constant over time and serves as a fundamental basis for all flow analyses. The variations in fluid density and velocity are expressed by Equation (25) [
52,
53,
54]:
In cases where there is an additional mass source, is assigned a value greater than 0. However, in this study, no external mass source was assumed; therefore, was set to 0.
4.1.2. Momentum Equation
The momentum equation describes the changes in momentum of a fluid element due to external forces such as pressure, viscous forces, and gravity. It represents the numerical implementation of Newton’s second law of motion, as expressed in Equation (26) [
52,
53,
54]:
4.1.3. Energy Equation
The energy equation is based on the principle of energy conservation within the system and is used to calculate the temperature distribution by accounting for internal energy, thermal conduction, and viscous dissipation in the fluid, as expressed in Equation (27) [
52,
53]:
is defined by Equation (28):
The enthalpy
is given by Equation (29):
The species enthalpy
is defined by Equation (30):
The reference temperature used in the enthalpy calculation is 298.15 K when using the pressure-based solver. The energy equation is included in the solver by enabling the Energy Equation option in ANSYS Fluent, allowing the calculation of temperature variations due to heat transfer and phase change within the working fluid.
4.1.4. Multiphase Flow Model (VOF)
Since both vapor and liquid phases coexist inside the heat pipe, the Volume of Fluid (VOF) model in ANSYS Fluent was employed in this study. The VOF model solves a single set of momentum equations while tracking the volume fraction of each phase throughout the computational domain, enabling the modeling of two or more immiscible fluids. This model is available only with the pressure-based solver and does not allow void regions where no fluid exists. The VOF algorithm calculates the volume fraction of the q-th phase using Equation (31), after which the volume fractions of all phases are summed within each computational cell [
53,
54,
56]:
Here, represents the volume fraction of the q-th phase, and it satisfies one of the following three conditions:
: The cell contains none of the q-th phase (empty cell).
: The cell is completely filled with the q-th phase.
: The cell contains the interface between the q-th phase and one or more other phases.
In this study, only the VOF model was employed to simulate the two-phase flow inside the heat pipe. The primary objective of the CFD analysis was to accurately reproduce the interfacial behavior associated with evaporation and condensation.
In ANSYS Fluent, the VOF model is configured through the multiphase model settings, where parameters such as interfacial surface tension and initial volume fractions of each phase can be defined. This enables dynamic tracking of the spatial distribution of liquid and vapor phases within the heat pipe.
The VOF model is particularly well-suited for capturing sharp liquid–vapor interfaces. While the Eulerian–Eulerian model is commonly used for dispersed multiphase flows, it does not distinguish between phases at the interface.
Guerrero et al. [
57] compared both models and found that only the VOF model could replicate actual two-phase flow structures, such as slug and bubbly flow, observed in experiments. Based on these considerations, the VOF model was selected as the most appropriate method in this study.
4.1.5. Evaporation–Condensation Model
In the VOF model, the Lee model is employed to simulate interphase mass transfer resulting from evaporation and condensation, as described by Equation (32)–(34) [
52,
54]:
If
(evaporation),
If
(condensation),
This model calculates mass transfer such that evaporation occurs when the working fluid temperature is higher than the saturation temperature, and condensation occurs when it is lower.
4.2. CFD Simulation Model
The simulation model in this study consists of a wickless heat pipe (thermosyphon type) integrated cooling system, in which three 4680 cylindrical lithium–ion battery cells are arranged in an equilateral triangular configuration. To improve computational efficiency and reduce simulation time, the domain was simplified to one-third of the full geometry based on its inherent symmetry, as illustrated in
Figure 17. The simplified domain represents one-third of the equilateral triangle, corresponding to a single battery cell, and the analysis focuses on the internal flow of the working fluid and the resulting steady-state temperatures under different heat generation conditions.
Water was used as the working fluid, and the filling ratios were set to 30%, 40%, and 50%, based on the internal volume, using the Split Body function in SpaceClaim V 2024.2.0.06032. The simulations were performed using a pressure-based transient solver with the energy equation enabled, the Realizable k-ε turbulence model, and a multiphase VOF model (implicit scheme) incorporating the Lee phase-change model. Each simulation was run with a time step of 0.1 s and continued until the battery temperature reached steady state (flow time of 60 s).
4.2.1. Grid Independence Test
To ensure reliable numerical results, a grid independence test was conducted. The mesh configuration was constructed using unstructured triangular elements. Mesh refinement was applied in the vicinity of wall regions to accurately resolve fluid dynamics and thermal gradients within the simulation domain. While increasing the total number of mesh elements generally improves the precision of the computed results, it also significantly raises the computational cost and may exceed the capacity of available hardware.
To evaluate the influence of mesh resolution on simulation accuracy and identify a practical configuration, four mesh cases were tested under identical boundary conditions. Each case was simulated using a filling ratio of 40%, a heat generation rate of 85 W, and a coolant inlet temperature of 293 K. The simulations were conducted over a physical time duration of 10 s. The results of this assessment are summarized in
Figure 18.
Among the four mesh configurations tested in this study, Case C provided the most balanced outcome in terms of numerical accuracy and computational efficiency.
This configuration consisted of 2,140,049 mesh elements, and the simulation resulted in a maximum temperature of 1033.01 K. Case D, which used a finer mesh with 2,856,949 elements, produced a maximum temperature of 1057.94 K. The difference in temperature between Case C and Case D was 2.36%, which was considered minor. However, the computational time required for Case D was significantly longer. Case A was composed of 1,055,264 mesh elements, and the maximum temperature obtained was 852.58 K. Further reduction in mesh density caused numerical instability, and the simulation could not be completed successfully. Case B included 1,342,004 mesh elements and resulted in a maximum temperature of 1002.67 K, which indicated a temperature deviation of approximately 14.9% compared to Case A. These findings are summarized in
Figure 18. Based on these results, the mesh configuration used in Case C was selected for all subsequent simulations, as it ensured sufficient numerical precision and computational stability.
The choice of mesh resolution was further guided by practical considerations related to computational resources. The selected mesh ensured convergence and stable operation of the solver while remaining within acceptable limits of memory usage and simulation time.
Although adaptive refinement methods are known to improve spatial accuracy, particularly in regions with steep temperature gradients or near the liquid and vapor interface, such techniques were not employed in the present study. This decision was made due to the high computational cost associated with transient multiphase simulations involving the volume of fluid method in combination with the Lee phase change model. Future research will consider the application of local mesh refinement strategies to improve the resolution of interfacial phenomena and capillary-scale transport mechanisms.
4.2.2. Boundary Conditions
As shown in
Table 6, a total of seven CFD cases were constructed for this study. CFD Cases 1–4 were designed to implement a digital twin of the heat pipe integrated immersion cooling system, aiming to validate the consistency between the simulation and experimental results. CFD Cases 5–7 were carried out under high heat generation conditions that could not be tested experimentally in order to conduct an extended performance analysis of the system and achieve the intended purpose of the digital twin approach.
In CFD Cases 1–3, the filling ratios were set to 30%, 40%, and 50%, respectively, with a constant heat input of 85 W applied to the battery surface. CFD Cases 4–7 focused on the 40% filling ratio, which showed the best performance in the experiments, and applied heat inputs of 50 W, 120 W, 155 W, and 190 W, respectively. The condenser temperature
was set to 293 K, consistent with the experimental setup. The detailed boundary conditions are summarized in
Table 7.
4.3. CFD Simulation Results and Discussion
As expected from the CFD simulation results, condensation occurred along the inner walls of the system.
Figure 19 corresponds to the conditions of CFD Case 2 and illustrates the variation in liquid volume fraction near the upper cooling plate between flow times 58.1 and 58.4 s. The evaporated fluid loses heat through the cooling plate, subsequently condenses, and flows downward along the wall due to gravity, returning to the bottom of the system. This process repeats, resulting in continuous circulation of the working fluid.
Figure 20a–c presents the vapor volume fraction distributions within the system at filling ratios of 30%, 40%, and 50%, respectively (corresponding to CFD Cases 1–3), under a heat input of 85 W. The figure illustrates only the vapor phase. Regions with high vapor volume fractions, typically corresponding to localized hotspots, are displayed in red. The remaining blue regions do not represent a complete absence of liquid but instead indicate areas with low vapor concentration, which may contain either liquid or air.
This distinction arises because any region not initially filled with liquid is automatically assigned as rarefied air within the simulation. Due to the difference in density between air and liquid, air tends to accumulate in the upper portion of the system, while the denser liquid phase tends to migrate downward along the walls and settle at the bottom. When comparing the results with the vapor volume fraction scale fixed between 0 and 1, both cases (a) and (b) show vigorous vapor generation and clearly observable evaporation–condensation circulation, with condensed liquid flowing downward along the wall. However, in case (c), vapor initially appears but suddenly disappears, suggesting a breakdown in the circulation pattern. This behavior can be explained by the increased volume of water, which leads to a higher heat capacity and a rapid decrease in system temperature, reducing the time spent above the saturation point. Moreover, the reduced vapor space due to the increased liquid volume may result in rapid pressure buildup within a confined area, leading to abrupt vapor condensation. These conditions deviate from the typical latent heat exchange cycle of a functioning heat pipe, where continuous evaporation and condensation occur.
On the other hand, in the 30% filling condition, nearly all the liquid evaporates, indicating the onset of a dry-out condition. At filling ratios lower than this, proper heat pipe operation may not be achieved. In contrast, at 40% filling, sufficient working fluid remains at the bottom, while vapor generation is observed on both sides around the battery. The rising vapor and the descending condensed liquid clearly demonstrate a stable evaporation–condensation circulation.
Additionally, the average surface temperatures for each filling ratio were extracted from the simulation results. At a filling ratio of 30%, the average temperature was 327.2 K, while it slightly decreased to 325.3 K at a filling ratio of 40%. At a filling ratio of 50%, the average temperature further declined to 316.1 K, indicating a general trend of decreasing temperature with increasing liquid volume. However, at a filling ratio of 30%, the rising temperature suggested a tendency toward dry-out, while at 50%, the rapid temperature drop indicated a breakdown in evaporation–condensation circulation.
Therefore, the optimal filling ratio for this thermosyphon-type heat pipe system is determined to be 40%, which is also supported by experimental results showing the best performance under this condition.
Table 8 summarizes the changes in steady-state temperature and flow characteristics of CFD Cases 1–3.
To implement a digital twin of the heat pipe integrated cooling system, the simulation results from CFD Cases 1 through 4 were compared with the experimental data.
Figure 21a presents the simulation result of CFD Case 4, in which the average temperature was 318.5 K, closely matching the experimental average temperature of 316.4 K. Under the low heat generation condition of 50 W, an evaporation–condensation circulation structure was observed; however, due to the reduced vapor generation, the vapor inflow into the condensation region and its recirculation slowed down, resulting in less active circulation compared to higher heat generation conditions.
Figure 21b,c compares the average steady-state surface temperatures of the battery obtained from CFD Cases 1–3 with those from the experiments. The temperature differences between the simulation and experimental results remained within a range of 2–6 K, demonstrating a high level of agreement.
Figure 22 presents the results of CFD Cases 5–7, which were performed under high heat generation conditions that could not be tested experimentally due to safety concerns. These simulations provide an extended analysis of the system’s thermal performance. As the heat input increased, the amount of vapor generated within the system also increased significantly. At 120 W, a stable evaporation–condensation circulation structure was observed, but the amount of vapor exceeded the capacity of the condensation region, leading to near saturation throughout the domain even at the point where condensation occurred. From 155 W onward, delayed recovery of condensed liquid at the condenser was observed, and the vapor volume fraction began to increase abnormally throughout the system. Under the 190 W condition, the system was almost entirely filled with saturated vapor, and the evaporation–condensation circulation collapsed due to insufficient condensation and return flow. These results indicate that, under high heat loads of 155 W or more, excessive vapor saturation, imbalance in condensation and recirculation, internal pressure buildup, and the onset of dry-out conditions prevent the heat pipe from maintaining its normal phase change-driven operation.
The steady-state temperatures under 120 W and 155 W were 327.2 K and 335.9 K, respectively, remaining below the critical temperature threshold for thermal runaway. However, under the 190 W condition, the battery temperature did not converge and instead continued to increase over time. As the heat input increases, the average temperature shows a general upward trend, and when the heat input exceeds 155 W, the heat pipe is no longer able to maintain stable operation based on the phase-change mechanism. As a result, thermal performance may deteriorate, and it becomes necessary to define a heat source limit to ensure system stability.
Table 9 summarizes the changes in steady-state temperature and flow characteristics under various heat loads at a 40% filling ratio. Based on this analysis, it can be inferred that 155 W is the upper limit of heat input for maintaining stable evaporation–condensation circulation under the experimentally verified optimal filling ratio of 40%.
5. Conclusions
This study investigated the thermal performance of a heat pipe-based cooling system specifically developed for cylindrical lithium–ion battery cells in the 4680 format. The system was evaluated through experimental testing, one-dimensional simulation using AMESim, and three-dimensional analysis using computational fluid dynamics. All wick configurations, from Wick 1 to Wick 5, showed improved thermal performance at lower filling ratios within the 30 to 70% range. Among them, Wick 5, which employed a blended fabric material with a crown-shaped structure, demonstrated the most favorable performance under an 85 W heat load at a 30% filling ratio. The recorded maximum temperature was 47 °C, the surface temperature difference was 2.8 °C, and the thermal resistance was 0.26 °C/W. Although Wick 4 exhibited a slightly lower thermal resistance of 0.24 °C/W, Wick 5 was identified as the most effective configuration when all performance criteria were considered. The crown-shaped structure was found to improve phase-change efficiency at reduced fluid volumes, contributing to lower system weight and enhanced packaging efficiency, both of which are essential for automotive battery applications.
To confirm the observed trends and extend the analysis to a broader set of conditions, a one-dimensional AMESim model was constructed. Correction factors derived from experimental results were applied to improve consistency between simulation and measurement. The model produced the most accurate predictions within the 30 to 50% filling ratio range, which corresponded to the optimal conditions identified through experimentation. Parametric evaluations revealed that the filling ratio and wick geometry had the most significant influence on thermal response. A reduction in coolant temperature from 20 °C to 10 °C decreased the surface temperature by 5.6 °C, corresponding to an improvement of 9.6%. Variations in wick thickness and tortuosity yielded smaller improvements of 1.2% and 0.7%, respectively. The selection of wick material was found to have limited impact on system performance.
Three-dimensional simulations using ANSYS Fluent were performed to analyze system behavior under high heat input conditions that could not be safely tested in the laboratory. The simulations focused on a wickless thermosyphon configuration within the 30 to 50% filling ratio range, which represented the most thermally active condition observed experimentally. Under an 85 W heat input and a 40% filling ratio, the average temperature was 325.3 K, and stable internal circulation through evaporation and condensation was maintained. Increasing the heat input to 155 W raised the average temperature to 335.9 K and resulted in early signs of internal instability. At 190 W, further temperature increase and loss of convergence indicated the onset of dry-out conditions and system failure. These results indicate the existence of a thermal limit that must be carefully considered during design to ensure operational stability.
To reduce simulation complexity, the CFD model excluded the porous wick domain. However, future simulations will incorporate detailed wick structures to improve alignment with physical conditions. In addition, further development of the one-dimensional simulation framework will include an expanded range of environmental and operational variables. This enhancement will enable more comprehensive assessment of system performance and battery temperature distribution under realistic usage scenarios. Experiments and simulations with electrically insulating working fluids will also be pursued to improve safety and broaden applicability in automotive systems.
The study proposes a novel thermal management strategy specifically optimized for the thermal and structural characteristics of large cylindrical battery cells. Compared to conventional indirect liquid cooling methods, the proposed system reduced thermal resistance and improved temperature uniformity at the battery surface. These improvements are expected to support higher cell packing densities by minimizing local temperature differences and enabling compact module integration. Similar to immersion cooling methods, the proposed system involves direct contact between the coolant and the battery surface, allowing for suppression of flame propagation during thermal runaway events. However, unlike conventional immersion systems, the proposed method uses a smaller volume of coolant, which supports a lighter and more efficient design without compromising cooling effectiveness.
In conclusion, this work presents a practical and experimentally verified cooling configuration that addresses the thermal regulation and structural requirements of next-generation battery modules. The findings may serve as a foundational reference for the continued development and refinement of advanced thermal management systems in high-power electric vehicle applications.