Next Article in Journal
Spatio-Temporal Characterization of Metal Stocks at a Provincial Scale: The Case of Iron and Steel Industry in Henan Province, China
Previous Article in Journal
Application of Waste Tire Carbon for Iron-Containing Dust Reduction in Industrial Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Analysis for Cost-Effective Temperature Reduction in High-Power Light-Emitting Diodes Using Thermal via Array

1
Department of Electrical and Information Engineering, Seoul National University of Science and Technology, 232 Gongneung-ro, Nowon-gu, Seoul 01811, Republic of Korea
2
Department of Bionic Machinery, KIMM Institute of AI Robot, Korea Institute of Machinery and Materials, 156 Gajeongbuk-ro, Yuseong-gu, Daejeon 34103, Republic of Korea
3
Department of Information Display, Kyung Hee University, Seoul 02447, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2025, 15(12), 6505; https://doi.org/10.3390/app15126505
Submission received: 1 May 2025 / Revised: 4 June 2025 / Accepted: 6 June 2025 / Published: 9 June 2025
(This article belongs to the Special Issue Recent Advances and Applications Related to Light-Emitting Diodes)

Abstract

The dissipation of excessive heat in high-power light-emitting diodes (LEDs) is essential for maintaining luminous efficiency, color stability, and device lifetime. While the incorporation of thermal vias in substrates is commonly used to improve heat dissipation, increasing their number is difficult in the limited area due to fabrication constraints. In this study, we use finite element analysis to investigate the effects of thermal via configurations on LED performance, including variations in the number of vias, spacing between vias, and their misalignment relative to the LED, arising from manufacturing tolerances. We found that the reduction in LED temperature saturated beyond a certain number of vias. Moreover, heat reduction can be further enhanced by optimizing the spacing between vias under a fixed number of vias. Based on these findings, the design of via configurations can achieve both fabrication feasibility and effective heat dissipation in high-power LEDs.

1. Introduction

Light-emitting diodes (LEDs) have largely replaced traditional light sources due to their high energy efficiency and long lifetime. Thermal management in high-power applications is crucial as the majority of input energy—typically 70–85%—is dissipated as heat [1,2]. Inadequate heat dissipation leads to a self-reinforcing temperature rise, significantly degrading LED performance and reliability [3,4,5,6,7,8]. As high-density LED devices have emerged, the importance of thermal management has become increasingly evident [9,10]. For instance, micro-LED array panels generate continuous heat within confined regions, which can severely impact device performance. In flexible LEDs, the inherently low thermal conductivity of substrate materials imposes additional constraints on thermal management. LEDs integrated on polyimide substrates often lack sufficient thermal conduction layers, resulting in substantial reductions in optical output and overall device performance [11]. The design of high-density LED arrays, including the output power of individual chips, the number of LEDs, and their spatial arrangement, plays a crucial role in achieving effective heat distribution, thereby reducing the overall thermal resistance of the system [12].
Various approaches have been explored to enhance heat dissipation, including the use of high thermal conductivity packaging materials [8,13,14] and external thermal management components such as heat sinks [15,16,17,18], heat pipes [18,19], and active cooling systems [14,20,21]. The heat dissipation performance of these methods strongly depends on power output and device form factor.
Heat sinks, typically composed of highly thermally conductive materials (e.g., aluminum or copper), dissipate heat by increasing the effective surface area exposed to the surrounding air. Radial heat sinks, which feature symmetrically arranged fins around a central axis, have been widely adopted, with key design parameters such as base thickness and fin geometry optimized to enhance thermal dissipation efficiency [15,16,17,18]. In addition, various fin designs, such as staggered, elliptic, or twisted geometries, have been developed to further facilitate effective natural convection cooling [22,23].
Heat pipes enable efficient heat transfer to remote cooling surfaces through the phase change of a working fluid involving evaporation and condensation. In LED devices, they have demonstrated significant reductions in thermal resistance, particularly in compact or enclosed environments. Beyond the conventional cylindrical shape, loop-type or flat-type heat pipes with micro-grooves have been proposed for high-power LEDs [24,25,26]. Moreover, transient modeling of heat pipe behavior has enabled accurate prediction of thermal hysteresis and dryout recovery dynamics [27].
Active cooling systems, such as miniature fans or thermoelectric modules, are commonly employed to dissipate heat using external power. While they offer fast response and localized cooling capability, they are often hindered by relatively low energy efficiency. For example, the coefficient of performance of thermoelectric modules remains significantly below that of traditional refrigeration systems, limiting their use in high-precision applications [28]. Moreover, although hybrid designs integrating phase-change materials or immersion cooling techniques have been proposed, their compact, scalable integration in miniaturized LED systems is elusive [29,30].
In another approach, thermal vias are embedded in substrates to provide efficient heat dissipation pathways. These sub-millimeter-scale vertical holes, fully filled or coated with highly thermally conductive materials, facilitate heat transfer from LEDs to underlying heat management components [31,32]. Thermal vias reduce the thermal resistance of devices by providing direct, high-conductivity pathways from the LED junction to the underlying heat-spreading layers. In contrast, heat sinks, heat pipes, and active cooling techniques are primarily designed to enhance heat dissipation to the ambient environment rather than directly controlling the internal thermal resistance of the system. In terms of cost, thermal vias are considered a cost-effective solution, as they can be implemented using standard substrate fabrication processes without requiring additional components. Also, circuit layouts and substrate dimensions remain unchanged with the implementation of vias, enabling their integration into various device form factors. Therefore, the integration of vias is a practical and scalable thermal management solution in compact, high-power LED devices.
The optimization of via configurations is essential for enhancing heat dissipation efficiency [31,32,33,34,35]. For instance, the heat transfer performance of vias improves when positioned close to heat sources, such as directly beneath LEDs rather than in peripheral regions. However, the number of vias is constrained by the minimum manufacturable feature size of the holes and the required spacing between them. In addition, increasing via density can introduce fabrication complexity, imposing further constraints on their practical implementation. Furthermore, localized heat accumulation caused by low substrate thermal conductivity or weak interlayer adhesion may offset the benefits of increasing via number and density.
In detailed thermal via configuration, parameters such as dimensions, location, number, and even misalignment due to manufacturing tolerances are interdependent and thus require comprehensive analysis. However, previous studies have often addressed these factors individually using simplified models. Furthermore, these parameter studies have rarely been evaluated in relation to the associated fabrication cost and complexity. As discussed later, the marginal thermal gains from increasing the number of vias gradually decrease, which reduces efficiency relative to the increased fabrication complexity.
In this study, we analyze the temperature reduction in high-power LEDs under various thermal via designs using three-dimensional (3D) finite element analysis (FEA). We investigate the peak value and distribution of temperature in the LED and the substrate under various via configurations, including their number, pitch (center-to-center distance), and positional misalignment. We found that the LED temperature saturates as the number of vias increases. For a given number of vias, comprehensively adjusting their pitch further reduces the LED temperature. Misalignment of the via array significantly degrades heat dissipation performance, particularly at lower via densities. This study provides a systemic approach to designing thermal via layouts while accounting for fabrication reliability in high-power LED systems. Such an approach can also be applied as a versatile framework for pixel-level thermal management in a wide range of thermally constrained applications.

2. Methods

A transient 3D thermal analysis was conducted using the Heat Transfer in Solids interface in COMSOL Multiphysics (version 5.2). The governing equation for heat conduction is expressed as:
ρ C P T t = k T + Q
where ρ is the density (kg/m3), Cp is the specific heat capacity (J/kg·K), k is the thermal conductivity (W/m·K), T is the temperature (K), and Q is the volumetric heat generation rate (W/m3). The equation for radiation is expressed as:
n q = ε σ ( T a m b 4 T 4 )
where Tamb is the ambient temperature (K), q is the heat flux vector, n is the unit outward normal vector, ε is the surface emissivity, and σ is the Stefan–Boltzmann constant (W/m2⋅K4). In this study, convection was neglected based on the assumption that heat conduction dominates over convection due to the small size of the LED die, as supported by previous studies [21,36,37,38,39,40]. Figure 1a illustrates the LED module model used in this study. The simulation domains include an LED chip, solder layer, substrate, copper layers, thermal interface material (TIM), and heat sink (Figure 1b). The LED chip with a footprint of 1 × 1 mm2, representative of high-power LEDs [32,41], is mounted at the center of the 2.8 × 2.8 mm2 substrate with the solder interlayer. Solid copper layers are laminated on the top and bottom surfaces of the substrate (25 μm thick for the top layer and 20 μm thick for the bottom layer). The substrate is attached to the heat sink via TIM. Thermal vias are vertical holes with a diameter of 0.1 mm, fully filled with copper, a design commonly used in high-density printed circuit boards (PCBs) [32]. The detailed material properties and geometric parameters of each component are summarized in Table 1.
The initial temperature for all domains was set to 25 °C. A uniform heat flux of 1.5 × 105 W/m2 was applied to the top surface of the LED chip to simulate heat generation. Adiabatic boundary conditions (−n·q = 0) were imposed on the lateral surfaces of all domains except for the heat sink to prevent heat loss. The outer surfaces were subject to surface-to-ambient radiation boundary conditions with an emissivity of 0.85. Two-dimensional temperature distributions were extracted from the top surface of the LED chip (Monitor 1) and the substrate (Monitor 2), as shown in Figure 1b.
To balance computational efficiency and accuracy, a physics-controlled mesh (normal setting) was applied, with locally refined elements in regions of high thermal gradients, particularly around the LED chip and the substrate (Figure 1c). Insufficient mesh refinement in these critical areas can lead to errors in temperature estimation and diminished local resolution. The mesh size in the reference model, including the heat sink, consisted of over 30,000 domain elements and 1000 edge elements. The minimum element quality, based on skewness, was 0.0078, and the average quality was 0.5235, both of which are suitable for stable numerical calculations. Mesh refinement had no noticeable effect on the resulting temperature distribution.
In this study, TLED is defined as the maximum temperature within the LED chip, and ΔTLED represents the temperature difference across the chip. Similarly, Tsub and ΔTsub denote the maximum temperature and the temperature difference within the substrate, respectively. Temperature data were collected after the system reached a steady thermal state. As shown in Figure 2a, TLED rapidly increased during the initial stage and began to plateau around 600 s, indicating that the thermal equilibrium had been reached. A time point of 1500 s was chosen for analysis to ensure that the system had fully stabilized. Regardless of the presence of thermal vias, TLED and Tsub were consistently measured on the Monitor 1 and Monitor 2 planes. The lowest temperature within the substrate was observed at the bottom surface of the substrate, remaining at approximately 63–64 °C across all simulation cases, independent of the thermal via configurations (Figure 1b). This indicates that the heat sink serves as the primary heat dissipation pathway under steady-state conditions [27,42]. As a reference case, TLED and Tsub reached 129 °C in the absence of thermal vias (Figure 2b,c).

3. Results and Discussion

Figure 3 presents a comparative thermal analysis with varying numbers of thermal vias. Thermal via arrays with uniform pitch were placed beneath the LED chip. For via numbers ranging from 1 to 9, the pitch was fixed at 0.4 mm. For 16 and 25 vias, the pitch was reduced to 0.3 mm and 0.2 mm, respectively, considering the LED footprint and via dimensions (Figure 3a).
As shown in Figure 3b,c, the temperatures of both the LED and substrate decreased with the addition of thermal vias. The spatial distribution of the LED temperature on the Monitor 1 plane was represented by the isothermal contours of 97.9−98.4 °C, 75.8−76.1 °C, 70.6−70.9 °C, and 66.3−66.3 °C for 1, 5, 9, and 25 vias, respectively (Figure 3b). Similarly, the spatial distribution of the substrate temperature on the Monitor 2 plane was 97.6−98.4 °C, 75.3−75.7 °C, 70.5−70.8 °C, and 66.2−66.3 °C, respectively (Figure 3c). The minimum temperature was consistently observed at the center of both the LED and substrate.
The full trends in TLED, ΔTLED, Tsub, and ΔTsub as a function of the number of thermal vias are shown in Figure 4. Even a single thermal via substantially reduced the temperature (TLED: 129 °C → 98.4 °C; Tsub: 129 °C → 98.4 °C). However, heat dissipation saturates beyond a certain number of vias. For instance, increasing the number of vias from 5 to 9 (80% increase) reduced TLED by approximately 7.3% (76.1 °C → 70.9 °C), whereas increasing from 9 to 16 (77% increase) resulted in only a 3.95% reduction, and from 16 to 25 (56% increase), the reduction was merely 2.5%. For configurations with 16 or 25 vias, the reduced pitch increased via density, and the greater number of vias expanded the total heat transfer area. However, these changes led to a negligible reduction in temperature. The saturation behavior can be understood by considering the effective cross-sectional area available for heat conduction. In the substrate, parallel heat conduction paths are formed by vias and surrounding substrate material in the vertical direction of the module [14,18,32,43,44,45]. Analytical models describing thermal resistance in such parallel heat conduction paths have been established in prior studies [43,44,45,46]. In our case, the effective thermal resistance (Rsub) across the substrate can be expressed as:
1 R s u b = k v i a A v i a L + k s u b A s u b L
where kvia and ksub are the thermal conductivities of thermal via (copper) and substrate material (FR4), respectively (in W/m·K); Avia and Asub are their respective cross-sectional areas (in m2); and L is the heat path length (i.e., substrate thickness, in meters). The large disparity between kvia (400 W/m·K) and ksub (0.3 W/m·K) results in a significant reduction in Rsub when even a single via is introduced. While the copper fraction continues to increase with the addition of more vias, its incremental contribution to reducing Rsub becomes marginal due to the already-dominant high-conductivity paths. As a result, the rate of decrease in Rsub gradually diminishes. This diminishing, in turn, emphasizes a fundamental trade-off between thermal performance and manufacturing efficiency. Adding vias becomes impractical due to the increasing PCB complexity and fabrication costs, so the via configuration should remain minimal to ensure feasibility. Building on the analytical model that explains the overall saturation trend, finite element simulations can further evaluate local temperature distributions according to specific geometric parameters of vias.
We analyzed the temperature response with respect to the number of thermal vias using a high thermal conductivity substrate, aluminum nitride (AlN) (180 W/m·K), as shown in Figure 5. Due to the significantly higher thermal conductivity of AlN compared to FR4, the presence or number of vias has a relatively minor influence on the overall thermal performance. Both TLED and Tsub remain around 63 °C, which is comparable to the performance of a 25-via configuration on an FR4 substrate. This result implies that the inherent thermal spreading capability of the AlN substrate effectively dissipates heat even without a large number of vias, indicating minimal thermal bottlenecking within the substrate.
Mechanical fatigue, such as delamination or cracking, typically occurs at interfaces between materials with mismatched coefficients of thermal expansion [46,47]. These defect sites can propagate rapidly under repeated thermal cycling conditions [48]. Therefore, minimizing temperature fluctuations is essential for enhancing reliability. As shown in Figure 6, during the transient heating stages at 100 s and 400 s, the presence of thermal vias substantially reduces the temperature rise compared to the no-via case. Specifically, TLED decreases from over 110 °C to below 50 °C at 100 s, and from over 130 °C to below 70 °C at 400 s when more than five vias are implemented. The saturation behavior observed in the quasi-steady-state (at 1500 s) is similarly reflected during these transient stages. This reduction in temperature variation is expected to alleviate thermomechanical stress and enhance reliability under thermal cycling conditions.
Temperature distribution as a function of pitch of the thermal vias was evaluated. Here, five- and nine-via configurations were selected as representative cases for optimized design (Figure 7a,c). For each case, the pitch was varied from 0.2 mm to 0.6 mm. In the five-via configuration, the spatial distribution of the LED temperature on the Monitor 1 plane was 75.5−76.1 °C, 75.4−75.8 °C, 75.8−76.1 °C, and 75.8−76.1 °C for 0.2, 0.3, 0.4, and 0.6 mm pitch, respectively (Figure 7b). In the nine-via configuration, this distribution was 70.8−71.3 °C, 70.7−70.8 °C, 70.6−70.9 °C, and 71.0−71.5 °C, respectively (Figure 7d). The general trend in temperature variation across the Monitor 1 plane became smaller with a higher number of vias. The smallest temperature variations occurred at 0.4 mm pitch for the five-via case (0.3 °C) and at 0.3 mm for the nine-via case (0.1 °C). Those values are far improved from the highest values, 0.6 °C and 0.5 °C at a pitch of 0.2 mm and 0.1 mm, respectively.
The full trends in TLED, ΔTLED, Tsub, and ΔTsub as a function of the thermal via pitch are shown in Figure 8. In the five-via configuration, the temperature variations on both the LED and substrate were minimal across all pitches (Figure 8a). Specifically, Tsub remained nearly constant at 76.1 °C for pitches of 0.2 mm, 0.4 mm, and 0.6 mm, with a slight decrease to 75.7 °C at a pitch of 0.3 mm. The corresponding ΔTsub values were 13.1 °C for 0.2 mm, 0.4 mm, and 0.6 mm and 12.7 °C for 0.3 mm. These results suggest that adjusting the pitch alone does not significantly improve heat dissipation when the overall via density is low.
In contrast, with nine thermal vias, thermal performance became more sensitive to pitch variation (Figure 8b). At pitches of 0.3 mm and 0.4 mm, Tsub reached the lowest value of 70.8 °C, with ΔTsub decreasing to 7.8 °C, indicating improved heat spreading and more uniform temperature distribution. At a pitch of 0.2 mm, Tsub slightly increased to 71.3 °C, with ΔTsub of 8.3 °C. Thermal vias located outside the heat-generating area contributed less to heat conduction, despite a higher number of vias. When the pitch increased to 0.6 mm, the outermost vias were positioned outside the LED chip, leading to degraded thermal performance. Tsub rose to 71.6 °C, and ΔTsub returned to 8.3 °C, offsetting the benefits of additional vias.
The optimal pitch condition varied depending on the number of vias. For example, the five- and nine-via configurations exhibited the best thermal performance at different pitches (0.4 mm and 0.3 mm, respectively). Therefore, the number and pitch of vias are interdependent design parameters, and adjusting one without considering the other may lead to suboptimal thermal behavior.
The thermal sensitivity to misalignment of thermal vias was investigated by laterally shifting the via array away from the center of the LED chip (Figure 9). Two configurations were examined: five vias with a pitch of 0.3 mm and nine vias with a pitch of 0.2 mm. The center of the via array was offset by (0.1 mm, −0.1 mm), (0.2 mm, −0.2 mm), and (0.4 mm, −0.4 mm) relative to the LED center (Figure 9a,c). In the five-via configuration, the spatial distribution of the LED temperature on the Monitor 1 plane ranged from 75.4−75.8 °C, 75.4−76.3 °C, 75.3−76.6 °C, and 75.3−78.1 °C for offset of (0 mm, 0 mm; no offset), (0.1 mm, −0.1 mm), (0.2 mm, −0.2 mm), and (0.4 mm, −0.4 mm), respectively (Figure 9b). In the case of the nine-via configuration, the corresponding distributions were 70.6−70.9 °C, 70.8−71.8 °C, 70.4−72.4 °C, and 70.4−73.9 °C (Figure 9d).
Notably, the nine-via configuration exhibited greater temperature variation under misalignment than the five-via case. Although the variation was smaller with no offset (0.4 °C and 0.3 °C for the five- and nine-via cases, respectively), it increased significantly even with slight misalignment, reaching 0.9 °C and 1.0 °C, respectively at an offset of (0.1 mm, −0.1 mm). This trend deviates from the more uniform temperature distributions observed in the pitch variation results (Figure 6 and Figure 7), where configurations with more vias generally showed better thermal uniformity. This implies that the effective area occupied by the via array, rather than via density, plays a dominant role in determining misalignment tolerance.
Figure 10 shows a comprehensive comparison between the two configurations. For the five-via case, TLED increased from 75.8 °C (no offset) to 76.3 °C, 76.6 °C, and 78.1 °C as the offset increased (Figure 10a). A similar trend was observed for the nine-via configuration, where TLED rose from 70.9 °C to 71.8 °C, 72.4 °C, and 73.9 °C (Figure 10b). The overall TLED value was smaller in the nine-via configuration due to the higher number of vias. However, the rate of ΔTLED increase was more noticeable in the smaller pitch configurations. In both cases, even slight misalignments, as small as 0.1 mm, significantly degraded thermal performance. Such levels of misalignments fall within the typical tolerance range of surface mount technology processes considering factors such as slippage during placement, solder flow, or thermal expansion during reflow. In such cases, increasing the number of vias can provide a buffer against fabrication tolerances despite the accompanying complexity. It is also important that the effective working area of the via array closely matches the footprint of the LED chip to ensure efficient thermal conduction. Misaligned vias, particularly those placed outside the heat-generating area, resulted in greater degradation of temperature uniformity. This thermal nonuniformity induces asymmetric temperature distributions across the substrate, concentrating thermal stress in localized regions. In addition, thermal shock in the presence of such uneven temperature distributions can significantly elevate local mechanical stress [47]. Consequently, degradation mechanisms such as solder joint fatigue, delamination, or component detachment may be accelerated [48]. It has also been reported that vias with inner void structures due to partial filling exhibit higher sensitivity to such thermal localization [46].

4. Conclusions

We performed a comprehensive thermal analysis of high-power LED modules to evaluate heat dissipation performance as a function of thermal via design, including the number of thermal vias, pitch, and misalignment. The temperatures of the LED and substrate exhibited saturation behavior as the number of thermal vias increased. The optimal via pitch for heat dissipation performance varied depending on the number of thermal vias, emphasizing the need for integrated consideration of pitch and via density. Misalignment of the thermal via array relative to the LED center also resulted in significant increases in both temperature and thermal nonuniformity, particularly for a smaller effective working area of the via array.
We propose a cost-effective and fabrication-reliable design strategy for thermal vias, offering practical guidelines for the scalable design of high-power LED modules. In practical applications, the minimum via diameter and via-to-via spacing are often constrained by standard fabrication limits [49]. Our proposed thermal via design guidelines are fully compatible with these manufacturing constraints. Notably, they achieve effective thermal performance without requiring via configurations that approach the limits of process capabilities. Our results demonstrate that even with relaxed via configurations, substantial heat dissipation is possible due to the saturation behavior in thermal performance. Therefore, the proposed strategy provides a balanced solution between thermal efficiency and fabrication feasibility using conventional, cost-effective processes.
Experimental validation of our proposed concept would be highly desirable for its practical application. Although this study primarily relies on numerical analysis, previous studies on similar thermal via implementations have reported good agreement between simulation and experimental measurements, typically within a 10% error margin when considering manufacturing tolerances [34,35,50].
Importantly, the proposed via configuration strategy can be effectively applied to a wide range of heat dissipation modules, offering consistent thermal performance improvements across various device layouts. Our approach is applicable to a wide range of LED layouts and substrate configurations.

Author Contributions

Conceptualization, Y.J.H., K.W. and S.-U.K.; methodology, S.-C.P., K.W. and S.-U.K.; software, B.-Y.L. and S.-C.P.; validation, B.-Y.L. and S.-C.P.; investigation, M.J.K. and S.-C.P.; data curation, Y.J.H., M.J.K. and S.-C.P.; writing—original draft preparation, Y.J.H., B.-Y.L. and S.-U.K.; writing—review and editing, B.-Y.L., K.W. and S.-U.K.; visualization, Y.J.H., K.W. and S.-U.K.; supervision, K.W. and S.-U.K.; project administration, S.-U.K.; funding acquisition, S.-U.K., Y.J.H. and B.-Y.L. contributed equally to this study. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors upon request.

Acknowledgments

This study was financially supported by Seoul National University of Science and Technology.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
LEDLight-emitting diode
FEAFinite element analysis

References

  1. Arik, M.; Becker, C.A.; Weaver, S.E.; Petroski, J. Thermal Management of LEDs: Package to System. In Proceedings of the Third International Conference on Solid State Lighting, San Diego, CA, USA, 5–7 August 2003; Volume 5187, p. 64. [Google Scholar]
  2. Ye, H.; Sau, K.; Van Zeijl, H.; Gielen, A.W.J.; Zhang, G. A Review of Passive Thermal Management of LED Module. J. Semicond. 2011, 32, 014008. [Google Scholar] [CrossRef]
  3. Tong, X.C. Advanced Materials for Thermal Management of Electronic Packaging; Springer Series in Advanced Microelectronics; Springer: New York, NY, USA, 2011; Volume 30, pp. 233–276. ISBN 9781441977588. [Google Scholar]
  4. Narendran, N.; Gu, Y. Life of LED-Based White Light Sources. J. Disp. Technol. 2005, 1, 167–171. [Google Scholar] [CrossRef]
  5. Pan, X.; Wei, M.; Yang, C.; Xiao, H.; Wang, C.; Wang, X. Growth of GaN Film on Si (111) Substrate Using AlN Sandwich Structure as Buffer. J. Cryst. Growth 2011, 318, 464–467. [Google Scholar] [CrossRef]
  6. Birchmore, S. Book Reviews: Technology Condemned. Expo. Times 1992, 103, 126–127. [Google Scholar] [CrossRef]
  7. Nam, S.R.; Jung, C.W.; Choi, C.H.; Kang, Y.T. Cooling Performance Enhancement of LED (Light Emitting Diode) Packages with Carbon Nanogrease. Energy 2013, 60, 195–203. [Google Scholar] [CrossRef]
  8. Kim, C.M.; Kang, Y.T. Cooling Performance Enhancement of LED (Light Emitting Diode) Using Nano-Pastes for Energy Conversion Application. Energy 2014, 76, 468–476. [Google Scholar] [CrossRef]
  9. Li, H.; Zhu, X.; Zhang, D.; Gao, Y.; Feng, Y.; Ma, Z.; Huang, J.; He, H.; Ye, Z.; Dai, X. Thermal Management towards Ultra-Bright and Stable Perovskite Nanocrystal-Based Pure Red Light-Emitting Diodes. Nat. Commun. 2024, 15, 6561. [Google Scholar] [CrossRef]
  10. Li, Z.; Tan, J.; Li, J.; Ding, X.; Tang, Y. A Review on Thermal Management of Light-Emitting Diodes: From Package-Level to System-Level. Appl. Therm. Eng. 2024, 257, 124145. [Google Scholar] [CrossRef]
  11. Asad, M.; Li, Q.; Sachdev, M.; Wong, W.S. Thermal and Optical Properties of High-Density GaN Micro-LED Arrays on Flexible Substrates. Nano Energy 2020, 73, 104724. [Google Scholar] [CrossRef]
  12. Lobo Ploch, N.; Rodriguez, H.; Stolmacker, C.; Hoppe, M.; Lapeyrade, M.; Stellmach, J.; Mehnke, F.; Wernicke, T.; Knauer, A.; Kueller, V.; et al. Effective Thermal Management in Ultraviolet Light-Emitting Diodes with Micro-LED Arrays. IEEE Trans. Electron Devices 2013, 60, 782–786. [Google Scholar] [CrossRef]
  13. Liu, Z.; Li, J.; Liu, X. Novel Functionalized BN Nanosheets/Epoxy Composites with Advanced Thermal Conductivity and Mechanical Properties. ACS Appl. Mater. Interfaces 2020, 12, 6503–6515. [Google Scholar] [CrossRef]
  14. Baldry, M.; Timchenko, V.; Menictas, C. Optimal Design of a Natural Convection Heat Sink for Small Thermoelectric Cooling Modules. Appl. Therm. Eng. 2019, 160, 114062. [Google Scholar] [CrossRef]
  15. Yu, S.H.; Lee, K.S.; Yook, S.J. Optimum Design of a Radial Heat Sink under Natural Convection. Int. J. Heat Mass Transf. 2011, 54, 2499–2505. [Google Scholar] [CrossRef]
  16. Jang, D.; Yu, S.H.; Lee, K.S. Multidisciplinary Optimization of a Pin-Fin Radial Heat Sink for LED Lighting Applications. Int. J. Heat Mass Transf. 2012, 55, 515–521. [Google Scholar] [CrossRef]
  17. Sundar, S.; Song, G.; Zahir, M.Z.; Jayakumar, J.S.; Yook, S.-J. Performance Investigation of Radial Heat Sink with Circular Base and Perforated Staggered Fins. Int. J. Heat Mass Transf. 2019, 143, 118526. [Google Scholar] [CrossRef]
  18. Costa, V.A.F.; Lopes, A.M.G. Improved Radial Heat Sink for Led Lamp Cooling. Appl. Therm. Eng. 2014, 70, 131–138. [Google Scholar] [CrossRef]
  19. Mitterhuber, L.; Defregger, S.; Hammer, R.; Magnien, J.; Schrank, F.; Horth, S.; Hutter, M.; Kraker, E. Investigation of the Temperature-Dependent Heat Path of an LED Module by Thermal Simulation and Design of Experiments. In Proceedings of the 2016 22nd International Workshop on Thermal Investigations of ICs and Systems, Budapest, Hungary, 21–23 September 2016; pp. 194–200. [Google Scholar]
  20. Ben Salah, S.; Ben Hamida, M.B. Heat Transfer Enhancement of Circular and Square LED Geometry. Int. J. Numer. Methods Heat Fluid Flow 2019, 29, 1877–1898. [Google Scholar] [CrossRef]
  21. Ben Abdelmlek, K.; Araoud, Z.; Ghnay, R.; Abderrazak, K.; Charrada, K.; Zissis, G. Effect of Thermal Conduction Path Deficiency on Thermal Properties of LEDs Package. Appl. Therm. Eng. 2016, 102, 251–260. [Google Scholar] [CrossRef]
  22. Hou, F.; Yang, D.; Zhang, G. Thermal Analysis of LED Lighting System with Different Fin Heat Sinks. J. Semicond. 2011, 32, 014006. [Google Scholar] [CrossRef]
  23. Yang, T.; Zhang, X.; Chang, Z.; Xu, R.; Ma, J.; Xu, L.; Xi, L. A Review on Application of Pin-Fins in Enhancing Heat Transfer. Energies 2024, 17, 4305. [Google Scholar] [CrossRef]
  24. Lu, X.-Y.; Hua, T.C.; Liu, M.J.; Cheng, Y.X. Thermal Analysis of Loop Heat Pipe Used for High-Power LED. Thermochim. Acta 2009, 493, 25–29. [Google Scholar] [CrossRef]
  25. Chang, C.; Han, Z.; He, X.; Wang, Z.; Ji, Y. 3D Printed Aluminum Flat Heat Pipes with Micro Grooves for Efficient Thermal Management of High Power LEDs. IEEE Trans. Power Electron. 2021, 11, 8255. [Google Scholar] [CrossRef] [PubMed]
  26. Huang, D.S.; Chen, T.C.; Tsai, L.-T.; Lin, M.T. Design of Fins with a Grooved Heat Pipe for Dissipation of Heat from High-Powered Automotive LED Headlights. Energy Convers. Manag. 2019, 180, 550–558. [Google Scholar] [CrossRef]
  27. Baraya, K.; Weibel, J.A.; Garimella, S.V. A Transient Heat Pipe Model Considering Wick Saturation Effects That Predicts Dynamic Evaporator Dryout and Recovery. Int. J. Heat Mass Transf. 2025, 242, 126837. [Google Scholar] [CrossRef]
  28. Chen, W.-Y.; Shi, X.-L.; Zou, J.; Chen, Z.-G. Thermoelectric Coolers: Progress, Challenges, and Opportunities. Small Methods 2022, 6, e2101235. [Google Scholar] [CrossRef]
  29. Cai, Y.; Hong, B.-H.; Wu, W.-X.; Wang, W.-W.; Zhao, F.-Y. Active Cooling Performance of a PCM-Based Thermoelectric Device: Dynamic Characteristics and Parametric Investigations. Energy 2022, 254, 124356. [Google Scholar] [CrossRef]
  30. Azarifar, M.; Cengiz, C.; Arik, M. Thermal and Optical Performance Characterization of Bare and Phosphor Converted LEDs through Package Level Immersion Cooling. Int. J. Heat Mass Transf. 2022, 189, 122607. [Google Scholar] [CrossRef]
  31. Shin, H.W.; Lee, H.S.; Jung, S.B. Analysis on Thermal Resistance of LED Module with Various Thermal Vias. In Proceedings of the 18th IEEE International Symposium on the Physical and Failure Analysis of Integrated Circuits, Incheon, Republic of Korea, 4–7 July 2011; pp. 1–4. [Google Scholar]
  32. Juntunen, E.; Tapaninen, O.; Sitomaniemi, A.; Jamsa, M.; Heikkinen, V.; Karppinen, M.; Karioja, P. Copper-Core MCPCB with Thermal Vias for High-Power COB LED Modules. IEEE Trans. Power Electron. 2014, 29, 1410–1417. [Google Scholar] [CrossRef]
  33. Yuan, Z.; Ding, D.; Zhang, W. Effect of Thermal via Design on Heat Dissipation of High-Lead QFN Packages Mounted on PCB. Appl. Sci. 2023, 13, 12653. [Google Scholar] [CrossRef]
  34. Liang, R.; Zhang, J.; Wang, S.; Ding, T.; Dai, J.; Chen, C. Experimental Study on the Effects of Eutectic Voids on the Thermal Performance within Flip-Chip Ultraviolet Light-Emitting Diodes. IEEE Trans. Compon. Packag. Manuf. Technol. 2016, 6, 1488–1492. [Google Scholar] [CrossRef]
  35. Kapusta, C.; Yin, L.; Nagarkar, K.; Walwaikar, A.; Srihari, K. Thermal Modeling and Experimental Analysis of High Power LED Using POL-Kw Packaging Technology. Int. Symp. Microelectron. 2019, 2019, 306–311. [Google Scholar] [CrossRef]
  36. Pape, H.; Schweitzer, D.; Chen, L.; Kutscherauer, R.; Walder, M. Development of a Standard for Transient Measurement of Junction-to-Case Thermal Resistance. Microelectron. Reliab. 2012, 52, 1272–1278. [Google Scholar] [CrossRef]
  37. Fu, H.-K.; Wang, C.-P.; Chiang, H.-C.; Chen, T.-T.; Chen, C.-L.; Chou, P.-T. Evaluation of Temperature Distribution of LED Module. Microelectron. Reliab. 2013, 53, 554–559. [Google Scholar] [CrossRef]
  38. Hu, J.; Yang, L.; Hwang, W.J.; Shin, M.W. Thermal and Mechanical Analysis of Delamination in GaN-Based Light-Emitting Diode Packages. J. Cryst. Growth 2006, 288, 157–161. [Google Scholar] [CrossRef]
  39. Li, J.; Yang, Q.; Niu, P.; Jin, L.; Meng, B.; Li, Y.; Xiao, Z.; Zhang, X. Analysis of Thermal Field on Integrated LED Light Source Based on COMSOL Multi-Physics Finite Element Simulation. Phys. Procedia 2011, 22, 150–156. [Google Scholar] [CrossRef]
  40. Kim, I.; Cho, S.; Jung, D.; Lee, C.R.; Kim, D.; Baek, B.J. Thermal Analysis of High Power LEDs on the MCPCB. J. Mech. Sci. Technol. 2013, 27, 1493–1499. [Google Scholar] [CrossRef]
  41. Ben Abdelmlek, K.; Araoud, Z.; Charrada, K.; Zissis, G. Optimization of the Thermal Distribution of Multi-Chip LED Package. Appl. Therm. Eng. 2017, 126, 653–660. [Google Scholar] [CrossRef]
  42. Park, D.H.; Lee, D.B.; Seo, E.R.; Park, Y.J. A Parametric Study on Heat Dissipation from a LED-Lamp. Appl. Therm. Eng. 2016, 108, 1261–1267. [Google Scholar] [CrossRef]
  43. Ha, M.; Graham, S. Development of a Thermal Resistance Model for Chip-on-Board Packaging of High Power LED Arrays. Microelectron. Reliab. 2012, 52, 836–844. [Google Scholar] [CrossRef]
  44. Lin, Y.; Lu, Y.; Gao, Y.; Chen, Y.; Chen, Z. Measuring the Thermal Resistance of LED Packages in Practical Circumstances. Thermochim. Acta 2011, 520, 105–109. [Google Scholar] [CrossRef]
  45. Kim, L.; Shin, M.W. Thermal Resistance Measurement of LED Package with Multichips. IEEE Trans. Compon. Packag. Technol. 2007, 30, 632–636. [Google Scholar] [CrossRef]
  46. Shao, S.; Liu, D.; Niu, Y.; O’Donnell, K.; Sengupta, D.; Park, S. A Study on the Thermomechanical Reliability Risks of Through-Silicon-Vias in Sensor Applications. Sensors 2017, 17, 322. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, A.; Lou, J.; Wang, B.; Huang, W.M. Thermal Shock Fracture in a Cracked Strip: Incorporating Convective Heat Transfer between Lateral Surfaces and Ambient Environment. J. Therm. Stress. 2024, 47, 672–694. [Google Scholar] [CrossRef]
  48. Depiver, J.A.; Mallik, S.; Harmanto, D. Solder Joint Failures under Thermo-Mechanical Loading Conditions—A Review. Adv. Mater. Process. Technol. 2021, 7, 1–26. [Google Scholar] [CrossRef]
  49. Li, R.S. Optimization of Thermal via Design Parameters Based on an Analytical Thermal Resistance Model. In Proceedings of the Sixth Intersociety Conference on Thermal and Thermomechanical Phenomena in Electronic Systems, Seattle, WA, USA, 27–30 May 1998; pp. 475–480. [Google Scholar]
  50. Shao, J.; Wei, F.; Zhao, X.; Solovey, J. Thermal Solutions for Surface Mount Power Devices. PCIM Eur. Conf. Proc. 2020, 1, 981–984. [Google Scholar]
Figure 1. Geometry of the LED module used in the FEA. (a) Perspective view and (b) cross-sectional view of the LED module, illustrating the LED chip, solder layer, substrate, copper layers, and heat sink. The locations of Monitor 1 and Monitor 2 are indicated with red lines in (b). (c) Finite element mesh applied to the LED module.
Figure 1. Geometry of the LED module used in the FEA. (a) Perspective view and (b) cross-sectional view of the LED module, illustrating the LED chip, solder layer, substrate, copper layers, and heat sink. The locations of Monitor 1 and Monitor 2 are indicated with red lines in (b). (c) Finite element mesh applied to the LED module.
Applsci 15 06505 g001
Figure 2. Reference simulation of the LED module without thermal vias. (a) Transient temperature response of TLED. (b) Three-dimensional temperature distribution across the full body of the LED and substrate. (c) Temperature distribution measured at Monitor 1, showing TLED reaching 129 °C at the center. Black lines represent isothermal contours.
Figure 2. Reference simulation of the LED module without thermal vias. (a) Transient temperature response of TLED. (b) Three-dimensional temperature distribution across the full body of the LED and substrate. (c) Temperature distribution measured at Monitor 1, showing TLED reaching 129 °C at the center. Black lines represent isothermal contours.
Applsci 15 06505 g002
Figure 3. Thermal analysis of the LED module with varying numbers of thermal vias. (a) Schematic diagrams of simulation models with 1, 5, 9, and 25 thermal vias. (b) Temperature distributions on the LED surface (Monitor 1). Solid lines indicate isothermal contours, and dashed circles indicate the positions of thermal vias. The labeled values indicate the lowest temperature represented in the isothermal contours. (c) Temperature distributions on the substrate surface (Monitor 2). Dashed boxes denote the projected location of the LED chip.
Figure 3. Thermal analysis of the LED module with varying numbers of thermal vias. (a) Schematic diagrams of simulation models with 1, 5, 9, and 25 thermal vias. (b) Temperature distributions on the LED surface (Monitor 1). Solid lines indicate isothermal contours, and dashed circles indicate the positions of thermal vias. The labeled values indicate the lowest temperature represented in the isothermal contours. (c) Temperature distributions on the substrate surface (Monitor 2). Dashed boxes denote the projected location of the LED chip.
Applsci 15 06505 g003
Figure 4. The trend of LED and substrate temperature with varying numbers of thermal vias. (a) TLED and ΔTLED as a function of the number of thermal vias. (b) Tsub and ΔTsub as a function of the number of thermal vias.
Figure 4. The trend of LED and substrate temperature with varying numbers of thermal vias. (a) TLED and ΔTLED as a function of the number of thermal vias. (b) Tsub and ΔTsub as a function of the number of thermal vias.
Applsci 15 06505 g004
Figure 5. The trend of LED and substrate temperature with varying numbers of thermal vias implemented in the AlN substrate. (a) TLED and (b) Tsub as a function of the number of thermal vias.
Figure 5. The trend of LED and substrate temperature with varying numbers of thermal vias implemented in the AlN substrate. (a) TLED and (b) Tsub as a function of the number of thermal vias.
Applsci 15 06505 g005
Figure 6. The trend of LED temperature at transient states with varying numbers of thermal vias. TLED at (a) 100 s and (b) 400 s.
Figure 6. The trend of LED temperature at transient states with varying numbers of thermal vias. TLED at (a) 100 s and (b) 400 s.
Applsci 15 06505 g006
Figure 7. Thermal analysis of the LED module with varying pitches of thermal vias. (a) Schematic diagrams of simulation models with five thermal vias and (b) corresponding temperature distributions on the LED surface (Monitor 1). (c) Schematic diagrams of simulation models with nine thermal vias, and (d) corresponding temperature distributions on the LED surface (Monitor 1). In (b,d), solid lines indicate isothermal contours, and dashed circles indicate the positions of thermal vias. The labeled values indicate the lowest temperature represented in the isothermal contours.
Figure 7. Thermal analysis of the LED module with varying pitches of thermal vias. (a) Schematic diagrams of simulation models with five thermal vias and (b) corresponding temperature distributions on the LED surface (Monitor 1). (c) Schematic diagrams of simulation models with nine thermal vias, and (d) corresponding temperature distributions on the LED surface (Monitor 1). In (b,d), solid lines indicate isothermal contours, and dashed circles indicate the positions of thermal vias. The labeled values indicate the lowest temperature represented in the isothermal contours.
Applsci 15 06505 g007
Figure 8. The trend of LED and substrate temperature with varying pitches of thermal vias. TLED, Tsub, ΔTLED, ΔTsub as a function of the pitch of thermal vias for (a) five thermal vias and (b) nine thermal vias.
Figure 8. The trend of LED and substrate temperature with varying pitches of thermal vias. TLED, Tsub, ΔTLED, ΔTsub as a function of the pitch of thermal vias for (a) five thermal vias and (b) nine thermal vias.
Applsci 15 06505 g008
Figure 9. Thermal analysis of the LED module with varying positions of thermal vias. (a) Schematic diagrams of simulation models with five thermal vias and (b) corresponding temperature distributions on the LED surface (Monitor 1). (c) Schematic diagrams of simulation models with nine thermal vias and (d) corresponding temperature distributions on the LED surface (Monitor 1). In (b,d), solid lines indicate isothermal contours, and dashed circles indicate the positions of thermal vias. The labeled values indicate the lowest temperature represented in the isothermal contours.
Figure 9. Thermal analysis of the LED module with varying positions of thermal vias. (a) Schematic diagrams of simulation models with five thermal vias and (b) corresponding temperature distributions on the LED surface (Monitor 1). (c) Schematic diagrams of simulation models with nine thermal vias and (d) corresponding temperature distributions on the LED surface (Monitor 1). In (b,d), solid lines indicate isothermal contours, and dashed circles indicate the positions of thermal vias. The labeled values indicate the lowest temperature represented in the isothermal contours.
Applsci 15 06505 g009
Figure 10. The trend of LED temperature with varying positions of thermal vias. TLED and ΔTLED as a function of offset for (a) five thermal vias with a pitch of 0.3 mm and (b) nine thermal vias with a pitch of 0.2 mm.
Figure 10. The trend of LED temperature with varying positions of thermal vias. TLED and ΔTLED as a function of offset for (a) five thermal vias with a pitch of 0.3 mm and (b) nine thermal vias with a pitch of 0.2 mm.
Applsci 15 06505 g010
Table 1. Material properties and geometrical parameters of each domain in the simulation.
Table 1. Material properties and geometrical parameters of each domain in the simulation.
CategoryMaterialWidth × Depth [mm2]Height [mm]Thermal Conductivity
[W/(m·K)]
LED ChipGaN1 × 10.05130
Solder60Sn-40Pb1 × 10.02550
Top Cu layerCu2.8 × 2.80.025400
SubstrateFR42.8 × 2.80.10.3
Bottom Cu layerCu2.8 × 2.80.02400
TIMResin-bonded
glass fiber board
2.8 × 2.80.014
Heat SinkAl6 × 84.5238
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hwang, Y.J.; Lee, B.-Y.; Kim, M.J.; Park, S.-C.; Won, K.; Kim, S.-U. Numerical Analysis for Cost-Effective Temperature Reduction in High-Power Light-Emitting Diodes Using Thermal via Array. Appl. Sci. 2025, 15, 6505. https://doi.org/10.3390/app15126505

AMA Style

Hwang YJ, Lee B-Y, Kim MJ, Park S-C, Won K, Kim S-U. Numerical Analysis for Cost-Effective Temperature Reduction in High-Power Light-Emitting Diodes Using Thermal via Array. Applied Sciences. 2025; 15(12):6505. https://doi.org/10.3390/app15126505

Chicago/Turabian Style

Hwang, Yong Jin, Bo-Yeon Lee, Min Ji Kim, Seung-Chul Park, Kanghee Won, and Se-Um Kim. 2025. "Numerical Analysis for Cost-Effective Temperature Reduction in High-Power Light-Emitting Diodes Using Thermal via Array" Applied Sciences 15, no. 12: 6505. https://doi.org/10.3390/app15126505

APA Style

Hwang, Y. J., Lee, B.-Y., Kim, M. J., Park, S.-C., Won, K., & Kim, S.-U. (2025). Numerical Analysis for Cost-Effective Temperature Reduction in High-Power Light-Emitting Diodes Using Thermal via Array. Applied Sciences, 15(12), 6505. https://doi.org/10.3390/app15126505

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