1. Introduction
Overhead transmission lines are critical components of modern power systems, ensuring the efficient and reliable transfer of electricity over long distances. However, expanding transmission capacity through new line construction is increasingly constrained by social, environmental, regulatory, and economic challenges. Large-scale projects often encounter lengthy approval processes, high capital costs, and public opposition due to their environmental footprint. In parallel, power system operators are confronted with rapidly changing operating conditions: increasing electricity demand, growing peak loads, deregulated energy markets, and, most critically, the large-scale integration of variable Renewable Energy Sources (RES) such as wind and solar [
1,
2]. These developments frequently push transmission networks to operate close to their thermal and stability limits, raising concerns about congestion, reduced reliability, and higher operational risks.
To address these challenges without the need for costly new infrastructure, operators are increasingly adopting advanced strategies to enhance the utilization of existing transmission assets, with reconductoring to increase line capacity and DLR to adjust operational limits based on real-time conditions emerging as particularly promising approaches [
3]. Reconductoring involves replacing existing conductors with high-capacity alternatives, but its implementation is capital-intensive and disruptive. DLR, by contrast, represents a more flexible, adaptive, and cost-effective approach. Unlike static line rating (SLR) methods that rely on conservative worst-case assumptions, DLR continuously adjusts the current-carrying capacity of lines according to real-time environmental conditions such as ambient temperature, wind speed, and solar irradiance [
4,
5,
6]. This dynamic adjustment increases available capacity, facilitates the integration of variable renewable generation, reduces congestion and curtailment, and improves overall grid flexibility [
5,
7,
8]. Recent advances in sensor technologies, weather forecasting, and data analytics have accelerated the deployment of DLR systems, with studies reporting 10–30% additional available capacity and significant improvements in operational reliability [
5,
9,
10].
Recent studies have further emphasized the practical applications and benefits of DLR. It has been shown that DLR-based monitoring enables real-time assessment of operational limits, enhancing safety and reducing the risk of conductor overheating [
11]. The technology has been demonstrated to optimize transmission capacity and increase grid flexibility without requiring additional infrastructure investments [
12]. Moreover, Dynamic Thermal Rating has been successfully applied to increase allowable current on existing lines while maintaining safety and reliability, contributing to congestion reduction and more efficient integration of renewable energy sources [
5]. These findings highlight the importance of investigating DLR methods, particularly for high-voltage lines with bundled conductor configurations.
Despite these advantages, the practical application of DLR still faces significant technical challenges. The most widely used industry standards, IEEE 738 [
13] and CIGRE Technical Brochure 601 [
14], provide thermal rating methods that model conductors as single, isothermal cylinders. While effective for individual conductors, these simplified models fail to account for the aerodynamic and thermal interactions present in multi-conductor bundles, which are commonly deployed in high-voltage transmission systems. In practice, bundle ampacity is often estimated by multiplying the ampacity of a single conductor by the number of conductors in the bundle. However, experimental and numerical studies demonstrate that leeward conductors in a bundle may experience elevated temperatures—sometimes several degrees higher than windward conductors—making such simplifications non-conservative [
15,
16]. This thermal imbalance can reduce the effective ampacity of the entire bundle and lead to underestimated risks of overheating, mechanical degradation, and accelerated aging.
Recent research has attempted to refine the modeling of overhead conductors by considering non-isothermal behavior, turbulence effects, and the influence of environmental conditions. While progress has been made for single-conductor systems [
17,
18,
19,
20], relatively few studies have focused on bundled configurations despite their widespread use in high-voltage networks. Moreover, existing analyses often rely on simplified turbulence models or partial thermal considerations, leaving a gap in high-fidelity multiphysics approaches that couple fluid dynamics with thermal conduction and radiation.
While numerous studies have employed k-ε, k-ω, and SST turbulence models for heat transfer and fluid flow simulations in various engineering applications, nearly all of these investigations focus on domains other than overhead conductors or bundled configurations. However, SST-based approaches specifically targeting the heat transfer in conductors or bundled conductors with consideration of DLR appear to be largely unexplored. Although some relevant studies may exist, the literature still shows a noticeable gap in high-fidelity multiphysics analyses that integrate SST turbulence modeling with conductor thermal behavior and DLR considerations.
Furthermore, existing studies in this area have largely concentrated on the thermal behavior of individual conductors or bundled systems, while paying little attention to how these temperature differences influence DLR [
16,
21]. In many cases, the non-uniform heating among sub-conductors is not considered, and the real-time current-carrying capacity of the line is assumed uniform. Ignoring these factors can result in non-conservative ratings and may reduce the reliability and efficiency of transmission system operation.
To provide a more accurate assessment of conductor ampacity under DLR conditions, this study presents several key contributions, as summarized below:
A comprehensive 2D coupled thermal–fluid model for overhead conductors was developed using COMSOL Multiphysics 4.3b with the SST turbulence model, enabling more accurate representation of convective heat transfer compared to conventional approaches.
The study extends the analysis to bundled conductors (two-, three-, and four-conductor configurations), commonly used in high-voltage transmission systems, which have received limited attention in terms of ampacity and thermal behavior under DLR conditions.
Unlike simplified industry practices that assume isothermal behavior or calculate bundle ampacity by multiplying single-conductor ratings, the proposed model explicitly considers thermal and aerodynamic interactions among subconductors, highlighting the temperature imbalance between windward and leeward conductors.
By comparing the numerical results with IEEE 738 and CIGRE TB601 standards, the study reveals important differences between simplified analytical methods and high-fidelity multiphysics modeling, particularly under DLR conditions.
The findings demonstrate that ignoring non-uniform heating in bundled conductors can lead to non-conservative ratings, which may underestimate the risks of overheating, mechanical degradation, and reduced transmission reliability.
In summary, the study provides a multiphysics-based methodology that enhances the accuracy of DLR estimation, thereby contributing to safer, more reliable, and more efficient operation of transmission systems.
This study develops a comprehensive 2D coupled thermal-fluid model for both single and bundled conductors under steady-state conditions, implemented in COMSOL Multiphysics using SST turbulence model. The model considers the convective and radiative heat transfer mechanisms, as well as aerodynamic interactions among subconductors in multi-conductor bundles. Environmental parameters such as wind speed, wind direction, ambient temperature, and solar irradiance are included as boundary conditions to closely simulate real operating conditions of overhead transmission lines. For bundled conductors, the model accounts for the thermal imbalance between windward and leeward subconductors, providing a more realistic estimation of temperature distribution along the line. Additionally, an analysis was conducted on the two-conductor bundle configuration with reduced spacing from 40 cm to 30 cm to demonstrate how closer conductor spacing could further increase the difference between numerical results and standard IEEE ampacity predictions.
Simulation results are compared with ampacity predictions from IEEE 738 and CIGRE TB601 standards to highlight the differences between simplified analytical approaches and numerical results. The comparison includes not only maximum conductor temperatures but also detailed spatial temperature distributions and local convective heat flux variations, providing insight into the limitations of standard single-conductor approximations. Moreover, the analysis considers how non-uniform heating in bundles affects DLR and how closer spacing in the two-conductor bundle can increase the difference between numerical results and IEEE predictions. The main objective of this study is to provide a more accurate, multiphysics-based methodology for estimating the current-carrying capacity of overhead transmission lines, with particular emphasis on two-conductor, three-conductor, and four-conductor bundles under DLR conditions. This approach allows for a more reliable prediction of line capacity under realistic operating conditions.
The remainder of the paper is organized as follows.
Section 2 provides an overview of IEEE 738 and CIGRE TB601, describes the modeling of stranded conductors and multi-conductor arrangements, and presents the details and assumptions of the computational model.
Section 3 presents and discusses the temperature, wind, and pressure distributions around single and bundled conductors, provides a quantitative comparison with IEEE 738 and CIGRE TB601, and a qualitative discussion in relation to other models reported in the literature, highlighting the differences and advantages of the proposed model. Finally,
Section 4 summarizes the key findings of the study.
3. Results and Discussion
To examine the impact of key environmental and operational parameters on conductor temperature, an example following IEEE methodology was considered, with a load current of 1025 A, ambient temperature of 40 °C, solar radiation of 1027 W/m2, surface emissivity (ε) set to 0.8, and wind speed assumed at 0.61 m/s. Building on this scenario, both single and bundled conductor configurations are analyzed using COMSOL Multiphysics. A comparative assessment is then performed to evaluate their thermal performance and calculate the corresponding DLR under identical environmental and electrical loading conditions.
3.1. Single Conductor Configuration
Understanding the thermal behavior of overhead conductors is fundamental to the reliable and efficient operation of power transmission systems. While internal Joule heating provides a relatively uniform baseline of thermal input, it is the external environmental influence, namely wind direction, convective cooling efficiency, and solar radiation, that sculpt the asymmetric temperature distribution along the conductor’s surface. In this study, wind is assumed to blow from right to left, defining the right side as the windward face and the left as the leeward. This orientation elegantly correlates the root causes of thermal variation to aerodynamic flow patterns and solar loading.
The simulations conducted for the single conductor case also serve as a model validation study. In this validation, the results obtained from the coupled multiphysics model are compared with the reference calculations provided in IEEE 738 and CIGRE TB601. During the simulation, the conductor is represented by 16 outer strands surrounding the core, each experiencing slightly different thermal conditions due to local convection and radiation effects. To obtain a representative conductor surface temperature, the temperatures of all 16 outer strands were numbered from 1 to 16 and average surface temperature (
Tc) was calculated as:
where
Tc is the average surface temperature of the outer strands,
Ti is the surface temperature of strand
i, and
N = 16 is the total number of outer strands. The relative error with respect to the reference standards is defined as:
Under the IEEE 738 reference operating conditions listed in
Table 6, the conductor surface temperature is reported as 100 °C, while the CIGRE TB 601 reference provides a value of 102 °C. The average strand temperature obtained from the simulation is 99.04 °C, corresponding to relative errors of 0.96 % and 2.90 % with respect to IEEE 738 and CIGRE TB 601, respectively.
This close agreement confirms that the numerical model reliably captures the thermal behavior of a single conductor under the defined operating conditions. As shown in
Figure 5, the temperature distribution along the conductor surface is highly non-uniform, reflecting the combined influence of aerodynamic and radiative effects.
Strand 2, located in the upper-leeward quadrant, reaches the highest temperature of approximately 102.2 °C, whereas Strand 11, positioned on the lower-windward side, remains at 95.8 °C. The elevated temperatures observed in the leeward strands result from reduced convective heat removal due to low local airflow and simultaneous exposure to direct solar radiation.
In contrast, windward strands experience stronger airflow and lower solar gain, leading to more effective convective cooling. The spatial temperature variation demonstrates the critical role of strand location and local flow conditions in determining thermal response. The strand-level temperature profile is presented in
Figure 6, and further illustrates the formation of two distinct thermal zones.
The hot zone, comprising Strands 1–9 on the upper-leeward side, exhibits temperatures ranging from 100 °C to 102 °C. These strands are exposed to low airflow velocities in the wake region and receive significant solar radiation, limiting convective heat transfer. The cool zone, covering Strands 10–15, experiences lower temperatures from 95.8 °C to 98 °C, where airflow accelerates around the conductor surface and convective heat removal is enhanced. A pronounced thermal gradient between Strands 9 and 10 coincides with the flow separation point, marking a clear transition between inefficient and efficient cooling regions. Strand 16, located at the leeward lower sector, shows a high temperature (~101 °C), reinforcing the spatial symmetry of thermal patterns dictated by aerodynamic effects and convective interactions. Airflow mechanisms driving this temperature distribution are visualized in
Figure 7, where the wake region is evident around Strands 2–9.
In this region, air velocity drops sharply, often approaching zero, leading to stagnation and reduced convective cooling. Strands 10–12 and 14–16 are exposed to accelerated airflow along the conductor curvature, which increases convective heat transfer and lowers the surface temperature. Strand 13, although experiencing near-zero tangential airflow, is located at the frontal stagnation point. Here, direct wind impingement compresses the thermal boundary layer, enhancing heat transfer despite the absence of significant lateral flow. The figure highlights the interplay between local airflow patterns and strand-specific thermal response.
The strand-level wind velocity distribution is shown in
Figure 8, which quantitatively confirms the heterogeneous nature of airflow along the conductor.
Wake-region strands exhibit velocities below 0.1 m/s, leading to limited convective dissipation and higher temperatures. In contrast, strands located in accelerated flow zones experience velocities exceeding 1.2 m/s, significantly enhancing convective heat removal. The velocity profile closely aligns with the observed temperature gradients, demonstrating the critical impact of local aerodynamic conditions on strand-level thermal behavior.
The strand-level pressure profile is illustrated in
Figure 9, where a local pressure maximum of +0.25 Pa occurs at Strand 13 due to frontal wind impact, while a trough of −0.25 Pa at Strand 10 reflects flow acceleration, consistent with Bernoulli’s principle.
Persistent low pressure and recirculating airflow in the leeward region (Strands 1–9) further reduce convective cooling. These aerodynamic features, combined with differential radiation exposure, provide a comprehensive explanation for the strand-level temperature variations along the conductor.
The results indicate that local airflow, pressure distribution, and solar radiation jointly control the thermal behavior of individual conductor strands. They emphasize the limitations of conventional average-temperature-based rating methods and demonstrate the necessity of high-resolution, strand-level modeling for accurate DLR assessment.
3.2. Two-Conductor Bundle Configuration
The airflow in the bundled conductor configuration moves from right to left, illustrating the complex thermal and aerodynamic interaction between the two conductors. As shown in
Figure 10, the windward conductor (C1) on the right is directly exposed to the ambient wind, allowing effective convective cooling on its windward face. Its leeward face, however, experiences higher temperatures due to the formation of a localized wind shadow. Heat generated by the conductor is transferred to the surrounding air, producing a thermal wake that trails downstream. This thermal wake modifies the local airflow characteristics, increasing turbulence and air temperature, which significantly impacts the cooling performance of the leeward conductor (C2). The leeward conductor receives this pre-heated and disturbed airflow, reducing the temperature gradient between the conductor surface and the air, thus lowering convective heat transfer efficiency. As a result, the downstream conductor accumulates more heat and reaches the highest temperature in the bundle, approximately 107.2 °C. This illustrates that, in bundled configurations, the thermal performance of each conductor is interdependent, and the downstream conductor is subject to less favorable cooling conditions than the windward conductor. The strand-level temperature profiles are shown in
Figure 11.
Both conductors exhibit similar profile shapes, indicating that wind direction is the primary factor controlling strand-level temperature distribution. In the windward conductor (C1), the leeward strands (Strands 1–9) reside in a “wind shadow” region and exhibit higher temperatures between 100 °C and 102.6 °C, while the windward strands (Strands 10–14) experience lower temperatures down to 95.6 °C, reflecting enhanced convective cooling. The leeward conductor (C2) shows an overall upward shift in the temperature profile by roughly 4 °C due to thermal shielding, with the leeward strands reaching up to 106.5 °C and windward strands remaining around 99.9 °C. The temperature differentials—7.0 °C for C1 and 6.6 °C for C2—highlight the effect of airflow pre-heating on the downstream conductor. The hot spot consistently occurs at Strand 2 in both conductors, indicating that the leeward region receives the least convective cooling. The temperature distribution demonstrates that both aerodynamic shielding and wake formation must be accounted for in DLR calculations for accurate ampacity assessment. The strand-level wind velocity is presented in
Figure 12.
The windward conductor (C1) experiences undisturbed ambient airflow, generating a characteristic velocity distribution with accelerated flow along the conductor sides (peaking at approximately 0.21 m/s at Strand 15) and stagnation at the frontal point (Strand 13) where the flow splits. The downstream conductor, located within the turbulent thermal wake of the upstream conductor, encounters reduced airflow across all strands, with a peak velocity of only about 0.14 m/s. This diminished airflow reduces convective heat transfer efficiency, explaining the higher temperature profile observed on the leeward conductor. Both conductors exhibit zero velocity at the frontal stagnation point, emphasizing that local flow separation and wake formation dictate strand-level cooling efficiency. The figure highlights the interplay between aerodynamic wake effects and strand-specific temperature responses. The strand-level surface pressure distribution is shown in
Figure 13.
Both conductors present a high-pressure stagnation point at Strand 13 and a low-pressure trough at Strand 10, consistent with cylindrical body flow characteristics. However, the magnitude of pressure variation differs: the windward conductor exhibits a peak pressure of +0.26 Pa and a minimum of −0.25 Pa, reflecting direct exposure to undisturbed wind. The leeward conductor (C2) experiences a dampened pressure profile, with a peak of +0.16 Pa and a minimum of −0.18 Pa, due to shielding by the windward conductor (C1).
The reduced pressure gradients on the downstream conductor lead to lower local wind velocities, further diminishing convective heat transfer and contributing to elevated temperatures. This clearly demonstrates how aerodynamic shielding modifies both pressure and velocity fields, directly influencing strand-level thermal behavior in bundled conductors.
These results demonstrate that the thermal behavior of bundled conductors is strongly controlled by aerodynamic interactions between the windward and leeward elements. The upstream conductor generates a thermal wake that both increases local air temperature and introduces turbulence, reducing convective cooling efficiency on the downstream conductor. As a result, the maximum temperature—and thus the ampacity limit—of the bundle is determined by the leeward conductor (C2) rather than the windward conductor. Accurate consideration of these wake-induced thermal and aerodynamic effects is therefore essential for reliable DLR and for optimizing the operational safety and efficiency of bundled transmission lines under varying environmental and loading conditions.
3.3. Three-Conductor Bundle Configuration
Figure 14 depicts the temperature distribution in a triple-bundle arrangement under crosswind.
The airflow, moving from right to left, first interacts with the windward conductors (C1 and C3). These conductors are directly exposed to fresh ambient air, which promotes effective convective heat transfer and keeps their surface temperatures relatively low. In contrast, the leeward conductor (C2) is located entirely within the turbulent wakes generated by C1 and C3. Instead of receiving undisturbed air, C2 is surrounded by preheated, low-velocity, and chaotic flow. Consequently, C2 reaches a surface temperature of about 107 °C, which is 4–5 °C higher than the windward conductors despite identical current, geometry, and boundary conditions. This clearly demonstrates that aerodynamic shielding is the dominant factor responsible for the thermal disadvantage of the leeward conductor and establishes it as the limiting element for ampacity in the system.
Figure 15 presents the strand-level surface temperature profiles for three sub-conductors.
All conductors exhibit a similar circumferential temperature distribution, with a maximum at the stagnation point (Strand 2) and a minimum near Strand 13, consistent with the effect of wind incidence. Compared with C1 and C3, the temperature profile of C2 is uniformly higher across all strands. This vertical shift indicates that the leeward conductor remains hotter along its entire circumference, not only at localized regions. The persistence of this offset demonstrates the continuous influence of wake effects and confirms that conductor position within the bundle is a key determinant of thermal performance.
Figure 16 presents the local wind velocity distribution on the conductor surfaces.
The windward conductors (C1 and C3) reach maximum surface velocities of approximately 0.21 m/s, enabling robust convective cooling. In contrast, the leeward conductor (C2) achieves only ~0.14 m/s at its maximum, a substantial reduction caused by wake-induced turbulence and flow deceleration. This reduction in airflow velocity leads to a lower convective heat transfer coefficient, thereby allowing more heat from Joule losses and solar radiation to accumulate at the surface. The stagnation point at Strand 13 is common to all conductors, indicating uniform wind incidence. Beyond this point, velocity fields diverge due to aerodynamic interaction, leading to asymmetric thermal behavior.
Figure 17 depicts the pressure distributions around the conductor surfaces.
All three conductors display a cylindrical pressure pattern, with a high-pressure stagnation region at Strand 13 and a low-pressure area near Strand 10. C3 exhibits the steepest pressure gradient, reflecting its position at the outer edge of the bundle and direct exposure to undisturbed wind. In contrast, C1 and C2 experience more moderate pressure variations due to partial aerodynamic shielding. Despite this dampened pressure profile, C2 remains the thermal hotspot, highlighting that pressure distribution alone cannot reliably predict conductor heating. Instead, the combined effects of reduced velocity, turbulence, and thermal wake interactions govern the downstream conductor’s elevated temperature.
3.4. Four-Conductor Bundle Configuration
Figure 18 shows the temperature distribution within a four-conductor bundle configuration under crosswind conditions.
The airflow approaches from right to left and first encounters the windward conductors (C1 and C4). These conductors are directly exposed to undisturbed ambient wind, which enables strong convective cooling and prevents excessive temperature rise. In contrast, the leeward conductors (C2 and C3) are located fully within the thermal wakes generated by C1 and C4. The airflow reaching them is slower, turbulent, and preheated, reducing its cooling capacity.
As a consequence, C2 reaches a peak surface temperature of 107.1 °C, while C1 and C4 remain near 102.5 °C. This temperature difference of about 5 °C occurs even though all conductors carry the same current and share identical geometry and boundary conditions.
This finding clearly shows that a conductor’s position within the bundle strongly influences its thermal behavior. In particular, C2 consistently reaches the highest temperature, making it the limiting factor for the system’s overall ampacity.
Figure 19 presents the circumferential temperature variation in each conductor as a function of strand position.
All conductors show a similar temperature trend, with a maximum at Strand 2 on the leeward side and a minimum near Strand 13. This pattern reflects the effect of crosswind on the conductors. The leeward conductors, C2 and C3, have temperature curves that are consistently higher than those of C1 and C4. This upward shift occurs along the entire circumference, showing that aerodynamic shielding affects the full perimeter of the leeward conductors. Among them, C2 reaches the highest temperature at all strand positions, confirming it as the thermal hotspot. Overall, these trends indicate that wake-induced heating is a continuous and circumferentially uniform phenomenon that elevates the temperature of leeward conductors.
Figure 20 shows the local wind velocity at the conductor surfaces.
The windward conductors (C1 and C4) achieve maximum values of approximately 0.21 m/s, providing effective convective cooling. In contrast, the leeward conductors (C2 and C3) reach only ~0.15 m/s, reflecting the reduced airflow momentum within the wake region. Although all velocity profiles intersect at the stagnation point (Strand 13), which confirms uniform wind incidence, they diverge significantly around the conductor circumference. The velocity decay on the leeward conductors leads to a lower convective heat transfer coefficient, which, when combined with identical Joule and solar heating, results in higher operating temperatures. This analysis establishes a direct causal link: wake-induced velocity reduction is the key aerodynamic mechanism driving the thermal disadvantage of leeward conductors.
Figure 21 presents the surface pressure distribution of the four conductors.
Each exhibits the classical cylindrical flow pattern, with a high-pressure stagnation point at Strand 13 and a low-pressure trough near Strand 10. The windward conductors (C1 and C4) exhibit the steepest pressure gradients, with peak values around +0.27 Pa and minima near −0.26 Pa, reflecting their direct exposure to the incoming airflow. In contrast, the leeward conductors (C2 and C3) show gentler pressure gradients, with maximum values near +0.17 Pa and minimum values around −0.18 Pa. This attenuation reflects the shielding effect: the leeward conductors are aerodynamically protected, but at the cost of reduced airflow momentum and weaker convective cooling. Importantly, despite the smaller pressure differentials, C2 and C3 still emerge as the hottest conductors, demonstrating that pressure alone cannot capture thermal risk. The combined evaluation of pressure, velocity, and temperature confirms that leeward conductors suffer from aerodynamic shielding, making them the critical elements that determine bundle ampacity.
3.5. Discussion
This study developed and validated a detailed 2D coupled thermal–fluid model to examine aerodynamic and thermal interactions in single and bundled conductors under DLR conditions. By implementing SST turbulence model in COMSOL Multiphysics, the work systematically evaluated the limitations of conventional standards such as IEEE 738 and CIGRE TB 601, which neglect position-dependent effects within bundled conductors. Validation against these standards for a single conductor demonstrated the reliability of the proposed numerical framework.
The simulation results, summarized in
Table 7, demonstrate that bundled conductors exhibit significant temperature non-uniformities driven by aerodynamic shielding, wake effects, and solar radiation.
As expected, the single conductor exhibited temperatures consistent with IEEE 738 and CIGRE TB601 values. In the two-conductor bundle, the windward conductor (C1) showed nearly identical behavior to the isolated conductor, while the leeward conductor (C2) reached ~103 °C due to wake effects—namely the reduced wind speed and pre-heated airflow caused by the upstream conductor. In the three-conductor arrangement, C1 and C2 behaved similarly to the two-conductor case, whereas C3, positioned beneath them, exhibited a slightly lower temperature (~98.7 °C). Although a similar temperature to C1 might be expected, the oblique solar incidence (76.1° relative to the x-axis) caused C1 to partially shade C3, reducing its absorbed solar radiation and thus limiting its heating. In the four-conductor configuration, C1 and C4 (windward) again matched the single-conductor values, while C2 consistently emerged as the hottest conductor (~103.3 °C), acting as the thermal hotspot. Conductor C3 exhibited an intermediate value, slightly cooler than C2, again due to partial shading. These findings confirm that downstream conductors systematically govern the thermal limit of the bundle, with temperature elevations of up to 3–6 °C compared to windward or isolated conductors.
Considering the maximum temperature observed in the bundled configurations (~103.3 °C at the DLR current of 1025 A according to IEEE 738), an additional analysis was performed to determine the current at which the hottest sub-conductor reaches the critical limit of 100 °C. The simulation results indicate that this occurs at approximately 993.27 A, corresponding to a reduction of about 3.1% compared to the standard IEEE 738 DLR estimate. The calculated DLR currents are very close for all bundled configurations, since the maximum temperatures of the sub-conductors are nearly identical (around 103 °C). Therefore, only minor differences are observed between the bundle cases, which are provided for clarity. These observations highlight that, despite the general reliability of IEEE 738, the allowable current it predicts for bundled conductors may be slightly higher than what is observed under realistic thermal conditions. It should be emphasized that this adjustment is specific to the conditions considered in this study; for other operational scenarios—including variations in sub-conductor spacing, conductor type, wind speed, and solar radiation—separate analyses are necessary, and further investigation is recommended.
To illustrate that the DLR difference can vary when operational parameters change, an additional simulation was performed in which all conditions were kept constant except for the sub-conductor spacing, which was reduced from 40 cm to 30 cm. The results showed a decrease of approximately 4.1% compared to the IEEE 738 estimate, providing a concrete example that even small variations in operational parameters can lead to noticeable changes in the calculated ampacity.
These findings highlight that although IEEE 738 provides generally reliable estimates, its assumptions are limited when applied to bundled configurations. To place these results in a broader context,
Table 8 compares the strengths and limitations of analytical, empirical, and numerical approaches, including the present study.
Table 8 provides a comparative overview of analytical, empirical, and numerical approaches for ampacity and DLR evaluation. Standard analytical methods, such as IEEE 738 and CIGRE TB601, are widely adopted due to their simplicity and ease of application. These methods, however, are limited to single conductors and rely on simplified representations of environmental conditions, neglecting complex interactions such as those between bundled conductors or localized thermal variations. Empirical approaches, as in [
11] make use of real-world measurements and provide reliable localized predictions under actual weather conditions. Nevertheless, their applicability is confined to the specific regions where measurements are conducted, and they cannot fully capture broader variations along the line. Numerical simulations allow for more detailed thermal assessments. For instance, in [
16] the authors employ a 3D RANS k-ε model to analyze multi-conductor bundles, accounting for geometric effects and strong wind conditions. While this approach provides comprehensive thermal insights, it comes at the expense of significant computational effort. In contrast, the present study utilizes a 2D FEM-based approach with the SST k-ω turbulence model to simulate up to four conductors. This methodology enables accurate prediction of local temperature distributions and hotspot locations while maintaining computational efficiency. By combining detailed thermal modeling with a manageable computational cost, the present approach effectively bridges the gap between classical analytical methods and complex 3D numerical simulations, offering a practical and reliable tool for DLR analysis across different conductor configurations.
While this study shows that a 2D simulation provides accurate and computationally efficient results for the thermal analysis of bundled conductors under the considered conditions, some cases may require a 3D simulation. This is particularly relevant for complex conductor geometries, such as spacers, non-uniform strand arrangements, or structural irregularities, and for scenarios with highly non-uniform wind or localized thermal effects. The 2D approach remains advantageous due to its computational efficiency, simpler geometry, and ease of implementation, making it suitable for routine analyses. By clarifying these assumptions and limitations, the study guides when a more detailed 3D simulation may be necessary, allowing a balance between accuracy and computational cost.
The results presented in
Table 7 indicate that the temperatures of leeward conductors in bundled configurations are consistently higher than those of windward or isolated conductors. This leads to a slight reduction in allowable current compared to the classical IEEE 738 estimates. For instance, in the two-, three-, and four-conductor bundles, the IEEE 738-based DLR current is 1025 A, whereas the proposed model predicts approximately 993 A for the hottest sub-conductor, corresponding to a reduction of about 3.1%. In an additional analysis where all operational parameters were kept constant but the sub-conductor spacing was reduced from 40 cm to 30 cm, the IEEE 738 DLR remained 1025 A, while the model predicted 983 A, representing a decrease of approximately 4.1%.
Such refinements in temperature assessment have both technological and economic implications. Technically, neglecting the thermal hotspot may result in overloading the conductor beyond its maximum allowable temperature, potentially accelerating aging, compromising mechanical stability, and reducing line reliability. Clearance distances may also be affected due to increased conductor sag, which can further compromise operational safety. From an operational perspective, an inaccurate understanding of the line’s true ampacity may lead to suboptimal energy management decisions, particularly in systems with high penetration of distributed generation, where load flow and demand balancing rely on accurate line ratings. Economically, these factors can translate into increased maintenance costs, unexpected service interruptions, or early conductor replacement, highlighting the importance of precise thermal modeling. The proposed methodology thus provides a framework that can support practical applications, guiding the development of ampacity adjustment strategies and improving operational decision-making in real-world transmission line management.
4. Conclusions
This study investigated the thermal behavior of bundled conductors under realistic environmental conditions by combining finite element modeling with advanced turbulence closure. A systematic comparison was carried out against analytical standards, highlighting the limitations of conventional methods in predicting local temperature variations within conductor bundles.
In summary, the findings establish that the common practice of estimating bundle ampacity by simply scaling single-conductor ratings is fundamentally non-conservative. The true thermal limit of a bundled line is dictated by its hottest leeward conductor, and neglecting this positional dependency can lead to overestimated ampacity, thermal overloading, and accelerated conductor aging. The study demonstrated that temperature non-uniformities within bundles, driven by wake effects, aerodynamic shielding, and partial solar shading, can elevate local conductor temperatures by up to 3–6 °C compared to windward or isolated conductors. This positional dependency introduces systematic deviations that analytical standards cannot capture.
A direct comparison against IEEE 738 revealed that the proposed model consistently predicts slightly lower allowable currents—approximately 3–4% reductions—highlighting the importance of localized thermal assessment in bundle configurations. Such deviations, while numerically small, can translate into meaningful operational impacts, particularly in terms of sag increase, accelerated aging, and overall system reliability. By bridging the gap between simplified analytical methods and more computationally demanding 3D numerical studies, the 2D FEM-based framework adopted here provides both accuracy and efficiency, making it a practical tool for DLR analysis in real-world applications.
To facilitate practical application in field operations, the model results were further processed to derive ampacity correction factors and functional relationships. These allow standard DLR estimates to be adjusted without performing full model simulations for every scenario. Additionally, data from simulations, supplemented by experimental measurements when necessary, can be used to train machine learning algorithms capable of predicting the required ampacity correction factors for diverse operational conditions. This methodology provides a systematic way to translate complex numerical model outcomes into actionable guidelines for real-world transmission line management while maintaining consistency with the underlying physical and thermal behavior of the conductors.
Future studies should explore transient and 3D simulations, supported by experimental validation to reinforce the findings. Additionally, systematically varying wind, solar incidence, loading, and bundle geometry could generate a large dataset for developing an artificial intelligence framework. Such a model could provide data-driven correction factors to supplement existing standards, offering transmission operators a practical tool to account for bundle-specific effects without relying on computationally intensive simulations in routine operation.