Next Article in Journal
Photoelasticity for Stress Concentration Analysis in Dentistry and Medicine
Previous Article in Journal
Elastic Properties of Open Cell Metallic Foams—Modeling of Pore Size Variation Effect
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Model to Calculate the Current–Temperature Relationship of Insulated and Jacketed Cables

Electrical Engineering Department, Universitat Politècnica de Catalunya, Rambla Sant Nebridi 22, 08222 Terrassa, Spain
*
Author to whom correspondence should be addressed.
Materials 2022, 15(19), 6814; https://doi.org/10.3390/ma15196814
Submission received: 23 August 2022 / Revised: 19 September 2022 / Accepted: 27 September 2022 / Published: 30 September 2022

Abstract

:
This paper proposes and validates using experimental data a dynamic model to determine the current–temperature relationship of insulated and jacketed cables in air. The model includes the conductor core, the inner insulation layer, the outer insulating and protective jacket and the air surrounding the cable. To increase its accuracy, the model takes into account the different materials of the cable (conductor, polymeric insulation and jacket) and also considers the temperature dependence of the physical properties, such as electrical resistivity, heat capacity and thermal conductivity. The model discretizes the cable in the radial direction and applies the finite difference method (FDM) to determine the evolution over time of the temperatures of all nodal elements from the temperatures of the two contiguous nodes on the left and right sides. This formulation results in a tri-diagonal matrix, which is solved using the tri-diagonal matrix algorithm (TDMA). Experimental temperature rise tests at different current levels are carried out to validate the proposed model. This model can be used to simulate the temperature rise of the cable when the applied current and ambient temperature are known, even under short-circuit conditions or under changing applied currents or ambient temperatures.

1. Introduction

The demand for electrical energy is currently growing worldwide [1,2,3], so power systems load levels are increasing. It is of paramount importance to ensure that power cables operate within their thermal limits to not compromise their safe operation. The ampacity of insulated cables can be calculated by applying the method detailed in the IEC 60287 standard [4]. However, this standard only provides formulas to determine the current rating or maximum permissible current under steady-state conditions and a maximum temperature increase, but it does not develop the heat transfer equation. The same applies for the IEC 60853 [5], which develops methods for determining the cyclic and emergency current ratings of power cables, but does not provide a method for determining their temperature evolution. The IEC 60986 [6] standard, which is related to the short-circuit temperature limits of insulated cables, provides tables with the maximum permissible short-circuit temperatures to limit the I2t heating based on the consideration of the range of limits used by various authorities, but it does not allow the temperature evolution of the cable to be determined. The IEC 60986 recognizes that the values in the tables are safe, but they are not necessarily ideal because there are very little available experimental data on actual cables. Given the limitations of the aforementioned standards, this work contributes to this field, since it develops and validates with experimental tests a dynamic model to determine the current–temperature relationship between insulated and jacketed cables in air. The proposed model includes the core, the inner insulation layer, the outer protective jacket and the air surrounding the cable. The model considers the different cable materials (conductor, insulation and jacket) and the temperature dependencies of their physical properties.
Current rating calculations on power cables require determining the temperature of the different cable layers for a specified current or determining permissible current for a specified cable temperature [7]. Therefore, in order to perform these calculations, it is necessary to determine the heat generated due to the Joule effect within the conductor and the rate of its dissipation away from the cable, which depends on the current level, cable size, composition and laying method. To this end, the heat conduction equation must be solved by applying numerical approaches [7].
Two- and three-dimensional finite element analysis (FEA) approaches have been widely used to address this problem [8,9,10,11], because FEA simulations are widely accepted as a powerful and realistic approach to determine the electromagnetic and thermal performance of power cables and other devices intended for power systems [12]. For example, the IEC TR 62095 standard [7] suggests applying FEA methods, when the methods discussed in the IEC 60287 (steady-state conditions) and IEC 60853 (cyclic conditions) cannot be applied. However, FEA approaches involve preparing the geometry and mesh of the problem. They are often memory-intensive and time-consuming due to the number of discretized elements they require, especially when solving coupled multiphysics problems. FEA approaches also require the purchase of specific software and the periodic maintenance of expensive licenses, as well as the involvement of qualified technicians [13].
In [14], a method to calculate the temperature rise of cable systems under steady-state conditions is presented, which the authors recognize as a simple procedure compared to the transient case. In [15], an empirical transient cable model is presented, which neglects radiation losses and changes in conductor resistance, among others. In [16], a thermal model of bare and insulated conductors is presented, but it does not include the jacket. In [17], a lumped parameter thermal network for direct current medium voltage cables is presented. Although this model allows the temperature rise of the cable to be determined, its accuracy is limited, in part because of the poor level of discretization, and in part, because it neglects the temperature dependency of the main parameters of the cable (thermal conductivity, specific heat capacity and electric resistivity). Similar approaches based on lumped parameter thermal networks are presented in [18,19,20,21], with similar limitations. In [22], an approach to calculate the transient temperature of a single-core insulated cable using an analytic approach is presented, but the model does not include the jacket and different simplifications are performed in order to solve the differential equations arising from the model.
Due to the limitations of FEM-based methods (geometry preparation, meshing, computational burden or software licenses among others) or the limitations of lumped parameter methods (poor discretization or not considering the temperature dependency of the cable parameters), to overcome these drawbacks, it is highly appealing to develop fast and accurate transient models [23], if possible based on model reduction methods [24].
This paper describes a numerical method developed to determine the radial temperature distribution (from the center of the core to the outer surface) of stranded insulated and jacketed cables, which is affected by the steady or time-varying electrical current flowing through the cable and the ambient temperature. The proposed model considers the temperature dependency of the main cable parameters, such as the electric resistivity of the conductor and the volumetric densities, thermal conductivities and specific heat capacities of the different cable materials. The method proposed in this paper may be used to determine the radial temperature distribution when the current is known or to determine the temperature of the cable yielding a maximum allowable temperature. Therefore, it allows the current corresponding to pre-established temperature limits of the cable to be determined, since it solves the non-stationary heat transfer equation. The proposed model also allows steady-state, transient and dynamic problems to be solved. The steady-state problem is applied when the electric current and the ambient and cable temperatures are a constant value and independent of time. The transient problem occurs when the ambient temperature remains constant but the current undergoes a step change, thus affecting the radial temperatures, following an exponential change. Finally, the dynamic problem can be applied when the ambient temperature and/or current change over time follows any pattern. Due to the increase in load levels, cables are often pushed to their limits. The model developed in this paper can also be a useful tool for cable maintenance tasks, which can also disclose the relationship between the current, thickness and material of the insulation and jacket layers, ambient temperature and cable temperature. Finally, the comprehensive method here proposed, which fully develops the physical equations that govern the heat transfer problem, can be easily adapted to other cable configurations.

2. Transient Thermal Model of the Insulated and Jacketed Cable

Figure 1 shows a cross section of the cable, which includes the conductor core (orange color), the inner insulation layer (blue color) and the outer insulating and protecting jacket (gray color). It is noted that both the inner insulation and the outer jacket are layers of the cable that protect the conductor. Whereas the insulation layer isolates the current flow, the jacket is the outermost layer, which protects the conductor core and the insulation from chemical deterioration and external elements.
The temperature of the cable mainly depends on two factors, that is, the self-generated heating due to the Joule effect and the ambient temperature. The electric current is confined within the metallic conductor core, which is usually made of copper or aluminum strands, so the heat flows from the conductor to the air passing through the inner insulation layer and the outer jacket to the surrounding air.
The transient thermal behavior of bare conductors, that is, without insulation and jacket layers, can be modeled by applying the methods found in the CIGRE [25] and IEEE Std. 738 [26] standards, which state that for bare conductors, the heat balance equation results from the balance between the heat gain and the heat loss terms as,
I R M S 2 r ( T ) = P c + P r P s + m c p ( T ) d T d t   [ W / m ]
where IRMS (A) is the root mean square value of the current flowing through the conductor, r (Ω/m) is the per unit length electric resistance of the conductor, Pc (W/m) and Pr (W/m) are, respectively, the per unit length convection and radiation heat loss terms, Ps (W/m) is the term due to the per unit length solar heat gain, m (kg/m) the per unit length conductor mass, cp(T) (J/(kgK)) the specific heat capacity of copper or aluminum (conductor material) and T (K) the mean temperature of the conductor and t (s) the time. Since in this paper the tests are performed indoors, the solar heat gain term Ps is not considered.
It is worth noting that the resistance per unit length depends on the temperature as,
r ( T ) = r T 0 [ 1 + α ( T T 0 ) ]
where T0 (K) is the reference temperature (usually 20 °C or 293.15 K), T (K) is the mean temperature of the analyzed conductor node and α (K−1) is the temperature coefficient of the resistance. It is known that with ac supply the current density throughout the cross-section of the conductor cannot be uniform due to the eddy current effects [27], but the measured value of r T 0 already includes such effects.
When dealing with insulated and jacketed cables, the analysis is more complex than that shown in (1), so a more detailed study is required, which is detailed in the next sections. In this case, three different materials are involved in the three layers of the cable, for example, copper in the conductor core, XLPE in the inner insulation layer and PVC in the outer jacket. For more accuracy, the temperature dependencies of the specific heat capacities and the thermal conductivities of such materials are considered.
Figure 2 shows the temperature dependency of the specific heat of copper [28], XLPE [29] and PVC [30], while Figure 3 shows the temperature dependency of the thermal conductivity of copper [31], XLPE [32] and PVC [33]. It is worth noting that these properties are needed to solve the heat transfer equations.

2.1. Domain Discretization and TDMA Formulation

As explained, the temperature of an insulated cable depends on two main factors, which are the self-generated Joule (I2r) heating and the externally ambient temperature. Thus, in general, heat flows radially from the central part of the conductor to the outer part of the jacket. The whole domain of the cable is discretized into many discrete elements along the radial dimension of the cable, as shown in Figure 4.
The heat transfer problem is solved along the radial axis because the heat is conducted through this axis. This approach assumes that the central point of any node has the mean temperature of the discrete element where it is placed. This assumption is accurate since a small spatial step Δx is chosen. The finite difference method (FDM) determines the nodal temperature from the temperatures of the two contiguous left-hand and right-hand side nodes. A tri-diagonal matrix describing the nodal temperature arises, so the temperatures can be determined by applying the tri-diagonal matrix algorithm (TDMA) [34], which has the form [12],
a i j T , i 1 j + 1 + b i j T , i j + 1 + c i j T , i + 1 j + 1 d i j = 0
i and j being, respectively, the indices related to the spatial and temporal steps, so that Tij = T(iΔx,jΔt) corresponds to the average temperature of the i-th node calculated at the j-th time step. Finally, aij, bij, cij and dij are constant coefficients. Equation (4) calculates the nodal temperatures from the temperatures of the neighboring right-hand side and left-hand side nodes.
[ b 1 j c 1 j a 2 j b 2 j c 2 j . a 3 j b 3 j c 3 j . c t o t j a t o t j a 1 j c t o t 1 j b t o t j ] · [ T 1 j + 1 T 2 j + 1 T 3 j + 1 . T t o t j + 1 ] = [ d 1 j d 2 j d 3 j . d t o t j ]   with   a 1 j = c t o t j = 0

2.2. Conductor Discretization

The conductor part (orange color) designated with subscript A includes NA nodal elements of thickness ΔxA (m) each, as shown in Figure 5, whereas the elements of the insulation layer are shown in blue color. An additional node is placed at the boundary between the conductor and the insulation layer.
To properly distribute the electric current and the electric resistance along each division, the electric resistivity ρe (Ω·m) and the current density Ji = Ii/Si (A/m2) in each element are required, Si and Ii being, respectively, the cross section and the current corresponding to the i-th element, so the resistance per unit length of the i-th element is ri = ρe /Si (Ω/m).
First node (i = 1) analysis. At the first node, heat is generated by the Joule effect and conducted to the adjacent outer node (right-hand side). The discretized heat transfer equation can be expressed as,
ρ A S 1 , A c p , A T 1 j + 1 T 1 j Δ t = I i 2 r i ( T 1 j ) k A T 1 j + 1 T 2 j + 1 Δ x A p 1  
with,
{ S 1 , A = π [ x 1 + 0.5 Δ x A ] 2 p 1 = 2 π [ x 1 + 0.5 Δ x A ]
where ρA (kg/m3) is the mass density of the conductor material, cp,A (J/(kgK)) is the specific heat capacity of the conductor material, ri (Ω/m) is the per unit length resistance of the i-th element of the conductor, kA (W/(mK)) is the thermal conductivity of the conductor material, and Tij (K) is the temperature of the i-th element calculated at time t = jΔt (s).
After reorganizing the terms in (5) to match with the TDMA formulation, the TDMA coefficients of the first node result in,
{ a 1 = 0 b 1 = ρ A S 1 , A c p , A Δ t + k A p 1 Δ x A c 1 = k A p 1 Δ x A d 1 = ρ A S 1 , A c p , A T 1 j Δ t + I i 2 r i ( T 1 j )
Generic node (i) analysis. A generic conductor node includes the heat generation and the heat conduction from the inner (left) to the outer (right) element, thus resulting in,
ρ A S i , A c p , A T i j + 1 T i j Δ t = I i 2 r i ( T i j ) + k A T i 1 j + 1 T i j + 1 Δ x A p i 1 k A T i j + 1 T i + 1 j + 1 Δ x A p i + 1
where
{ S i , A = π [ x i + 0.5 Δ x A ] 2 π [ x i 0.5 Δ x A ] 2 p i 1 = 2 π [ x i 0.5 Δ x A ] p i + 1 = 2 π [ x i + 0.5 Δ x A ] with   2 i N A
After reorganizing the terms above to match with the TDMA formulation, the TDMA coefficients of the generic node result in,
{ a i = k A p i 1 Δ x A b i = ρ A S i , A c p , A Δ t + k A p i 1 Δ x A + k A p i + 1 Δ x A c i = k A p i + 1 Δ x A d i = ρ A S i , A c p , A T i j Δ t + I i 2 r i ( T i j ) with   2 i N A
Boundary node (i = NA + 1). In the boundary node between the conductor and the insulating layer (see Figure 5), the thermal conductivity is different on each side of the node, and the Joule effect heat generation is only in the conductor, not in the insulation.
ρ A S i , A c p , A T i j + 1 T i j Δ t + ρ B S i , B c p , B T i j + 1 T i j Δ t = I i 2 r i ( T , i j ) + k A T i 1 j + 1 T i j + 1 Δ x A p i 1 k B T i j + 1 T i + 1 j + 1 Δ x B p i + 1
where
{ S i , A = π [ x i ] 2 π [ x i 0.5 Δ x A ] 2 S i , B = π [ x i + 0.5 Δ x B ] 2 π [ x i ] 2 p i 1 = 2 π [ x i 0.5 Δ x A ] p i + 1 = 2 π [ x i + 0.5 Δ x B ] with   i = N A + 1
In the boundary node, the resistance per unit length and current only affect the region of material A (conductor), so that the current density Ji = Ii/Si (A/m2) is required in each element, Si and Ii being, respectively, the cross section and the current of the i-th element. The resistance per unit length of the i-th element is ri = ρe/Si,A (Ω/m), where Ii = Si,AJi.
The TDMA coefficients of the boundary node result in,
{ a i = k A p i 1 Δ x A b i = ρ A S i , A c p , A Δ t + ρ B S i , B c p , B Δ t + k A p i 1 Δ x A + k B p i + 1 Δ x B c i = k B p i + 1 Δ x B d i = ρ A S i , A c p , A T i j Δ t + ρ B S i , B c p , B T i j Δ t + I i 2 r i ( T , i j ) with   i = N A + 1
Although the conductor has already been discretized, as the thermal conductivity of copper is approximately three orders of magnitude larger compared to that of the surrounding insulating materials (see Figure 3), a single temperature value can be assigned to the conductor, as the temperature drop over it is almost negligible.

2.3. Inner Insulation Discretization

The elements in the inner insulation layer (blue color) are expressed with subscript B, which includes NB nodal elements, as shown in Figure 6, whereas the elements of the outer jacket are represented in gray color. An additional node is placed at the boundary between the insulation layer and the outer jacket.
Regarding the insulation layer, no internal heat source exists, so the temperature only varies along the radial axis [35].
Generic node (i). Since no heat is generated within the generic node of the insulation layer, the heat transfer equation results in,
ρ B S i , B c p , B T i j + 1 T i j Δ t = k B T i 1 j + 1 T i j + 1 Δ x B p i 1 k B T i j + 1 T i + 1 j + 1 Δ x B p i + 1
where
{ S i , A = π [ x i + 0.5 Δ x B ] 2 π [ x i 0.5 Δ x B ] 2 p i 1 = 2 π [ x i 0.5 Δ x B ] p i + 1 = 2 π [ x i + 0.5 Δ x B ] with   N A + 2 i N A + N B
The TDMA coefficients of a generic node of the inner insulation layer result in,
{ a i = k B p i 1 Δ x B b i = ρ B S i , B c p , B Δ t + k B p i 1 Δ x B + k B p i + 1 Δ x B c i = k B p i + 1 Δ x B d i = ρ B S i , B c p , B T i j Δ t with   N A + 2 i N A + N B
Boundary node (i = NA + NB + 1). The heat transfer equation of the boundary node between the inner insulation layer and the outer jacket can be expressed as,
ρ B S i , B c p , B T i j + 1 T i j Δ t + ρ C S i , C c p , C T i j + 1 T i j Δ t = k B T i 1 j + 1 T i j + 1 Δ x B p i 1 k C T i j + 1 T i + 1 j + 1 Δ x C p i + 1
where
{ S i , B = π [ x i ] 2 π [ x i 0.5 Δ x B ] 2 S i , C = π [ x i + 0.5 Δ x C ] 2 π [ x i ] 2 p i 1 = 2 π [ x i 0.5 Δ x B ] p i + 1 = 2 π [ x i + 0.5 Δ x C ] with   i = N A + N B + 1
The TDMA coefficients of the boundary node between the inner insulation layer and the outer jacket result in,
{ a i = k B p i 1 Δ x B b i = ρ B S i , B c p , B Δ t + ρ C S i , C c p , C Δ t + k B p i 1 Δ x B + k C p i + 1 Δ x C c i = k C p i + 1 Δ x C d i = ρ B S i , B c p , B T i j Δ t + ρ C S i , C c p , C T i j Δ t with   i = N A + N B + 1

2.4. Outer Jacket Discretization

The elements in the outer jacket (gray color) are expressed with subscript C. As shown in Figure 7, it includes NC nodal elements. It can be observed that an additional node has been added to the boundary between the outer jacket and the air.
Generic node (i). Since no heat is generated in the generic node of the outer jacket, the heat transfer equation results in,
ρ C S i , C c p , C T i j + 1 T i j Δ t = k C T i 1 j + 1 T i j + 1 Δ x C p i 1 k C T i j + 1 T i + 1 j + 1 Δ x C p i + 1
where
{ S i , C = π [ x i + 0.5 Δ x C ] 2 π [ x i 0.5 Δ x C ] 2 p i 1 = 2 π [ x i 0.5 Δ x C ] p i + 1 = 2 π [ x i + 0.5 Δ x C ] with   N A + N B + 2 i N A + N B + N C
The TDMA coefficients of a generic node of the outer jacket are as follows,
{ a i = k C p i 1 Δ x C b i = ρ C S i , C c p , C Δ t + k C p i 1 Δ x C + k C p i + 1 Δ x C c i = k C p i + 1 Δ x C d i = ρ C S i , C c p , C T i j Δ t with   N A + N B + 2 i N A + N B + N C
Last node (i = NA + NB + NC + 1). The last node is placed in the boundary between the outer jacket and air, so indoors, only convection and radiation must be considered,
ρ C S i , C c p , C T i j + 1 T i j Δ t = k C T i 1 j + 1 T i j + 1 Δ x C p i 1 h p i + 1 ( T i j + 1 T a i r ) ε σ p i + 1 [ ( T i j ) 4 ( T a i r ) 4 ]
where
{ S i , C = π [ x i ] 2 π [ x i 0.5 Δ x C ] 2 p i 1 = 2 π [ x i 0.5 Δ x C ] p i + 1 = 2 π R C with   i = N A + N B + N C + 1
where RC (m) is the outer radius of the cable, σ (W/(m2K4)) is the Stefan–Boltzmann constant, and ε = 0.85 (-) is the emissivity coefficient [36,37], whereas h (W/(m2K)) is the heat transfer coefficient. Assuming that there is no wind, i.e., the worst condition, the heat transfer coefficient due to natural convection can be calculated as [26]:
h = 3.645 π ρ a i r 0.25 D 0.25 ( T - T a i r ) 0.25
Air density ρair (kg/m3) changes with the cable elevation H (m), air temperature Tair (K) and cable temperature T (K) as [26],
ρ a i r = 1.293 1.525 10 4 H + 6.379 10 9 H 2 1 + 0.00367 ( T a i r + T ) / 2  
Finally, the TDMA coefficients of the boundary node between the outer jacket and air result in,
{ a i = k C p i 1 Δ x C b i = ρ C S i , C c p , C Δ t + h p i + 1 + k C p i 1 Δ x C c i = 0 d i = ρ C S i , C c p , C T i j Δ t + h p i + 1 T a i r ε σ p i + 1 [ ( T i j ) 4 ( T a i r ) 4 ] with   i = N A + N B + N C + 1
It is worth noting that (1)–(27) have been programmed and solved in the MATLAB® environment by the authors of this work.

3. Experimental

3.1. Experimental Setup

A high-current transformer (variable output voltage 0–3 V, variable output current 0–1 kA, and 50 Hz alternating current) was used to conduct the experimental tests carried out at AMBER high-current laboratory of the Universitat Politècnica de Catalunya. To regulate the output current, the transformer has an input stage that includes an autotransformer. A loop formed by the analyzed insulated cable was directly connected to the output terminals of the transformer, as shown in Figure 8.
Figure 8b also shows the insertions made in the cable to place the thermocouples. Table 1 details the main characteristics of the insulated cable used in the experiments.
To eliminate any hot spot or heat sink, the length of the cable was 4 m, which is in agreement with the recommendations found in different international standards [38,39,40].
The current circulating through the cable loop was measured using a Rogowski coil (ACP1000 GMC-I, 1 mV/A, ±1%, DC to 10 kHz, PROsyS, Skelmersdale, UK), which provides a voltage that is linear with the electric current in the loop. The temperature in the different parts of the insulated cable (conductor-insulation boundary, insulation-jacket boundary and jacket-air boundary) was measured using low-thermal inertia welded-tip T-type thermocouples with a diameter of 0.2 mm. T-type thermocouples were selected because they are among the most accurate thermocouples, with an accuracy up to 0.5 °C. An OMEGA USB-2400 acquisition card was used to acquire the temperatures with a sampling frequency of 10 Hz. The ambient temperature was maintained at a constant value during the tests. Before the tests, it was ensured that the cable was at room temperature.
Experimental errors mainly depend on the accuracy of the Rogowski coil used to measure the current and the T-type thermocouples used to measure the temperature. Special care must be taken when performing the insertions made in the cable (conductor-insulation and insulation-jacket) to place the thermocouples, which are shown in Figure 8b. Finally, the values of different geometric conductor parameters (diameter and wall thickness of the insulation and jacket layers) as well as the physical properties of the materials (volumetric mass density, resistivity, specific heat capacity or thermal conductivity) also influence the accuracy of the simulation results.

3.2. Experimental Results

Experimental temperature rise tests were conducted to validate the thermal model of the conductor. These tests consist of applying current steps up to close the rated current to the cable loop during a certain time interval, during which the cable heats up in different stages. The applied current levels and durations of each current level are summarized in Table 2.
Figure 9 shows the experimental and simulated temperature profiles in the three measured points of the cable (conductor-insulation boundary, insulation-jacket boundary and jacket-air boundary) corresponding to the five current levels.
Results presented in Figure 9 show great agreement between experimental and simulated data, since the mean error is less than 1%, thus proving the suitability of the proposed cable model.
The proposed model also allows other parameters of the cable to be determined, such as the radial temperature distribution or the components of the heat balance equation, as shown in Figure 10.
It is worth noting that using an Intel®CoreTM i7-1185G7 CPU@ 3.0 GHz with 32 GB RAM (Intel, Santa Clara, CA, USA) with a time step of 10 s and nine nodes for each material, the software requires 0.7 s to run a simulation consisting of five current steps (see Figure 9) between t = 0 s and t = 8600 s.

3.3. Additional Experimental Results

To further validate the accuracy of the cable model, a second cable of 4 m length was tested, its characteristics are summarized in Table 3.
As with the first cable, experimental temperature rise tests were conducted to validate the proposed model of the conductor. They consist of applying five current steps to the cable loop during a certain time interval. The applied current levels and their durations are summarized in Table 4.
Figure 11 shows the experimental and simulated temperature profiles in the three measured points of the cable (conductor-insulation boundary, insulation-jacket boundary, and jacket-air boundary) for the five current levels.
Results presented in Figure 11 show great agreement between experimental and simulated data, since the mean error is also less than 1%, thus proving the accuracy of the proposed cable model.

4. Conclusions

This paper has presented a model for determining the current–temperature relationship between insulated and jacketed cables in air, which fully develops the physical equations governing the heat transfer problem. The model solves the transient heat transfer equations through the different layers of the cable, namely, the conductor core, the inner insulation layer, the outer insulating and protective jacket and the air surrounding the cable. To this end, the model discretizes the cable in the radial axis and applies a finite difference method approach to calculate the temperatures in all nodes of the discretized domain by considering the contiguous left-hand and right-hand side nodes. Since this approach leads to a tri-diagonal matrix, it is solved by applying the tri-diagonal matrix algorithm (TDMA). Experimental temperature rise tests performed by applying current steps of different magnitudes show the accuracy of the proposed method. The approach presented in this paper can be applied to determine the temperature rise of the cable once the applied current and ambient temperature are known, even under short-circuit conditions or under changing applied currents or ambient temperatures. The method proposed here can be easily adapted to other cable configurations.

Author Contributions

Conceptualization, J.-R.R.; methodology, J.-R.R. and J.L.; software, J.L.; validation, J.-R.R. and J.L.; investigation, J.-R.R. and J.L.; resources, J.-R.R.; writing—original draft preparation, J.-R.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministerio de Ciencia e Innovación de España, grant number PID2020-114240RB-I00 and by the Generalitat de Catalunya, grant number 2017 SGR 967.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to show our gratitude to Mecatraction (SICAME group) for providing the materials required for this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gu, L.; Mo, W.; Wang, Y.; Fan, G.; Liu, H.; Wang, P.; Jia, S.; Xiu, S.; Zhao, L.; Chu, B. Analysis of Short-time Withstand Current and Peak Withstand Current Test of 500kV AC Fault Current Limiter. In Proceedings of the 2021 IEEE 4th International Electrical and Energy Conference (CIEEC), Wuhan, China, 28–30 May 2021. [Google Scholar] [CrossRef]
  2. Dorraki, N.; Niayesh, K. An Experimental Study of Short-Circuit Current Making Operation of Air Medium-Voltage Load Break Switches. IEEE Trans. Power Deliv. 2022. [Google Scholar] [CrossRef]
  3. Elsayed, A.M.; Shaheen, A.M.; Alharthi, M.M.; Ghoneim, S.S.M.; El-Sehiemy, R.A. Adequate operation of hybrid AC/MT-HVDC power systems using an improved multi- objective marine predators optimizer. IEEE Access 2021, 9, 51065–51087. [Google Scholar] [CrossRef]
  4. IEC 60287-1-1:2006; Electric Cables-Calculation of the Current Rating-Part 1-1: Current Rating Equations (100% Load Factor) and Calculation of Losses-General. IEC: Geneva, Switzerland, 2006; pp. 1–136.
  5. IEC 60853-1:1985; Calculation of the Cyclic and Emergency Current Rating of Cables. Part 1: Cyclic Rating Factor for Cables up to and Including 18/30(36) kV. IEC: Geneva, Switzerland, 1985; pp. 1–39.
  6. IEC 60986:2000; Short-Circuit Temperature Limits of Electric Cables with Rated Voltages from 6 kV (Um = 7.2 kV) up to 30 kV (Um = 36 kV). IEC: Geneva, Switzerland, 2000; pp. 1–19.
  7. IEC TR 62095:2003; Electric Cables-Calculations for Current Ratings-Finite Element Method. IEC: Geneva, Switzerland, 2003; pp. 1–69.
  8. Callender, G.; Goddard, K.F.; Dix, J.; Lewin, P.L. A Flexible Model to Calculate Buried Cable Ampacity in Complex Environments. IEEE Trans. Power Deliv. 2021, 37, 2007–2015. [Google Scholar] [CrossRef]
  9. Liu, G.; Xu, Z.; Ma, H.; Hao, Y.; Wang, P.; Wu, W.; Xie, Y.; Guo, D. An improved analytical thermal rating method for cables installed in short-conduits. Int. J. Electr. Power Energy Syst. 2020, 123, 106223. [Google Scholar] [CrossRef]
  10. Bustamante, S.; Mínguez, R.; Arroyo, A.; Manana, M.; Laso, A.; Castro, P.; Martinez, R. Thermal behaviour of medium-voltage underground cables under high-load operating conditions. Appl. Therm. Eng. 2019, 156, 444–452. [Google Scholar] [CrossRef]
  11. Rasoulpoor, M.; Mirzaie, M.; Mirimani, S.M. Thermal assessment of sheathed medium voltage power cables under non-sinusoidal current and daily load cycle. Appl. Therm. Eng. 2017, 123, 353–364. [Google Scholar] [CrossRef]
  12. Abomailek, C.; Capelli, F.; Riba, J.-R.; Casals-Torrens, P. Transient thermal modelling of substation connectors by means of dimensionality reduction. Appl. Therm. Eng. 2017, 111, 562–572. [Google Scholar] [CrossRef]
  13. Abomailek, C.; Riba, J.-R.; Capelli, F.; Moreno-Eguilaz, M. Fast electro-thermal simulation of short-circuit tests. IET Gener. Transm. Distrib. 2017, 11, 2124–2129. [Google Scholar] [CrossRef]
  14. Neher, J.H.; McGrath, M.H. The Calculation of the Temperature Rise and Load Capability of Cable Systems. Trans. Am. Inst. Electr. Eng. Part III Power Appar. Syst. 1957, 76, 752–764. [Google Scholar] [CrossRef]
  15. Yenchek, M.R.; Cole, G.P. Thermal modeling of portable power cables. IEEE Trans. Ind. Appl. 1997, 33, 72–79. [Google Scholar] [CrossRef]
  16. Brito Filho, J.P. Heat transfer in bare and insulated electrical wires with linear temperature-dependent resistivity. Appl. Therm. Eng. 2017, 112, 881–887. [Google Scholar] [CrossRef]
  17. Shekhar, A.; Feng, X.; Hebner, R.; Gattozzi, A.; Strank, S.; Mor, A.; Ramirez-Elizondo, L.; Bauer, P. Thermal modelling and experimental validation for research on medium voltage DC cables. In Proceedings of the 2017 IEEE Power & Energy Society General Meeting, Chicago, IL, USA, 16–20 July 2017; pp. 1–5. [Google Scholar] [CrossRef]
  18. Echavarren, F.; Rouco, L.; Gonzalez, A. Dynamic thermal modeling of insulated cables. In Cigré Session 2012; Cigré: Paris, France, 2012; pp. 1–8. [Google Scholar]
  19. Yang, L.; Qiu, W.; Huang, J.; Hao, Y.; Fu, M.; Hou, S.; Li, L. Comparison of Conductor-Temperature Calculations Based on Different Radial-Position-Temperature Detections for High-Voltage Power Cable. Energies 2018, 11, 117. [Google Scholar] [CrossRef]
  20. Aras, F.; Biçen, Y. Thermal modelling and analysis of high-voltage insulated power cables under transient loads. Comput. Appl. Eng. Educ. 2013, 21, 516–529. [Google Scholar] [CrossRef]
  21. Tkachenko, V.A.; Kropotin, O.V.; Shepelev, A.O.; Kropotin, V.O. Mathematical model of cable power line with XLPE insulation with underground installation. Omsk Sci. Bull. 2018, 137–141. [Google Scholar] [CrossRef]
  22. Henke, A.; Frei, S. Transient Temperature Calculation in a Single Cable Using an Analytic Approach. J. Fluid Flow Heat Mass Transf. 2020, 7, 58–65. [Google Scholar] [CrossRef]
  23. Rezk, K.; Forsberg, J. A fast running numerical model based on the implementation of volume forces for prediction of pressure drop in a fin tube heat exchanger. Appl. Math. Model. 2014, 38, 5822–5835. [Google Scholar] [CrossRef]
  24. Wang, X.; Jiang, Y. Model reduction of discrete-time bilinear systems by a Laguerre expansion technique. Appl. Math. Model. 2016, 40, 6650–6662. [Google Scholar] [CrossRef]
  25. Cigré Working Group 22.12. Thermal Behaviour of Overhead Conductors; Cigré: Paris, France, 2002. [Google Scholar]
  26. IEEE Std 738-2012; IEEE Standard for Calculating the Current-Temperature of Bare Overhead Conductors. IEEE: New York, NY, USA, 2012.
  27. Riba, J.-R. Analysis of formulas to calculate the AC resistance of different conductors’ configurations. Electr. Power Syst. Res. 2015, 127, 93–100. [Google Scholar] [CrossRef] [Green Version]
  28. Banerjee, B. An evaluation of plastic flow stress models for the simulation of high-temperature and high-strain-rate deformation of metals. Acta Mater. 2005, 58, 6810–6827. [Google Scholar] [CrossRef]
  29. Lee, K.Y.; Yang, J.S.; Choi, Y.S.; Park, D.H. Specific heat and thermal conductivity measurement of XLPE insulator and semiconducting materials. In Proceedings of the 2006 IEEE 8th International Conference on Properties & Applications of Dielectric Materials, Bali, Indonesia, 26–30 June 2006; pp. 805–809. [Google Scholar] [CrossRef]
  30. Alford, S.; Dole, M. Specific Heat of Synthetic High Polymers. VI. A Study of the Glass Transition in Polyvinyl Chloride. J. Am. Chem. Soc. 1955, 77, 4774–4777. [Google Scholar] [CrossRef]
  31. Abu-Eishah, S.I. Correlations for the Thermal Conductivity of Metals as a Function of Temperature. Int. J. Thermophys. 2001, 22, 1855–1868. [Google Scholar] [CrossRef]
  32. Qi, X.; Boggs, S. Thermal and mechanical properties of EPR and XLPE cable compounds. IEEE Electr. Insul. Mag. 2006, 22, 19–24. [Google Scholar] [CrossRef]
  33. Kok, M.; Demirelli, K.; Aydogdu, Y. Thermophysical Properties of Blend of Poly (Vinyl Chloride) With Poly (Isobornyl Acrylate). Int. J. Sci. Technol. 2008, 3, 37–42. [Google Scholar]
  34. Datta, B.N. Numerical Linear Algebra and Applications, 2nd ed.; SIAM: Philadelphia, PA, USA, 2010; ISBN 0898717655. [Google Scholar]
  35. Li, X.J.; Yang, J.; Yan, B.Q.; Zheng, X. Insulated Cable Temperature Calculation and Numerical Simulation. MATEC Web Conf. 2018, 175, 03014. [Google Scholar] [CrossRef]
  36. Beňa, Ľ.; Gáll, V.; Kanálik, M.; Kolcun, M.; Margitová, A.; Mészáros, A.; Urbanský, J. Calculation of the overhead transmission line conductor temperature in real operating conditions. Electr. Eng. 2021, 103, 769–780. [Google Scholar] [CrossRef]
  37. Hong, S.S.; Yang, Y.C.; Hsu, T.S.; Tseng, K.S.; Hsu, Y.F.; Wu, Y.R.; Jiang, J.A. Internet of Things-Based Monitoring for HV Transmission Lines: Dynamic Thermal Rating Analysis with Microclimate Variables. In Proceedings of the 2020 8th International Electrical Engineering Congress (iEECON), Chiang Mai, Thailand, 4–6 March 2020. [Google Scholar] [CrossRef]
  38. ANSI/NEMA CC 1-2018; Electric Power Connection for Substation. ANSI: New York, NY, USA, 2009.
  39. IEC 60512-5-2:2002; Connectors for Electronic Equipment-Tests and Measurements-Part 5-2: Current-Carrying Capacity Tests-Test 5b: Current-Temperature Derating. IEC: Geneva, Switzerland, 2002; pp. 1–13.
  40. IEC 61238-1-1:2018; Compression and Mechanical Connectors for Power Cables-Part 1-1: Test Methods and Requirements for Compression and Mechanical Connectors for Power Cables for Rated Voltages up to 1 kV (Um = 1.2 kV) Tested on Non-Insulated Conductors. IEC: Geneva, Switzerland, 2018; pp. 1–89.
Figure 1. Cable layout including the copper core, the inner XLPE insulation layer and the outer PVC jacket.
Figure 1. Cable layout including the copper core, the inner XLPE insulation layer and the outer PVC jacket.
Materials 15 06814 g001
Figure 2. Temperature evolution of the specific heat capacity cp of the different cable materials: (a) Copper [28]. (b) XLPE [29]. (c) PVC [30].
Figure 2. Temperature evolution of the specific heat capacity cp of the different cable materials: (a) Copper [28]. (b) XLPE [29]. (c) PVC [30].
Materials 15 06814 g002
Figure 3. Temperature evolution of the thermal conductivity k of the different cable materials: (a) Copper [31]. (b) XLPE [32]. (c) PVC [33].
Figure 3. Temperature evolution of the thermal conductivity k of the different cable materials: (a) Copper [31]. (b) XLPE [32]. (c) PVC [33].
Materials 15 06814 g003
Figure 4. Cable discretization along the radial axis.
Figure 4. Cable discretization along the radial axis.
Materials 15 06814 g004
Figure 5. Conductor discretization.
Figure 5. Conductor discretization.
Materials 15 06814 g005
Figure 6. Inner insulation layer discretization.
Figure 6. Inner insulation layer discretization.
Materials 15 06814 g006
Figure 7. Outer jacket discretization.
Figure 7. Outer jacket discretization.
Materials 15 06814 g007
Figure 8. Experimental setup: (a) Transformer and cable loop used in the experiments. (b) Details of the insertions (conductor-insulation and insulation-jacket) made in the cable to place the thermocouples.
Figure 8. Experimental setup: (a) Transformer and cable loop used in the experiments. (b) Details of the insertions (conductor-insulation and insulation-jacket) made in the cable to place the thermocouples.
Materials 15 06814 g008
Figure 9. Conductor of 70 mm2. Experimental versus simulated temperature rise profiles in the three measured points of the cable (conductor-insulation boundary, insulation-jacket boundary and jacket-air boundary) and the relative difference between the experimental and simulated temperatures.
Figure 9. Conductor of 70 mm2. Experimental versus simulated temperature rise profiles in the three measured points of the cable (conductor-insulation boundary, insulation-jacket boundary and jacket-air boundary) and the relative difference between the experimental and simulated temperatures.
Materials 15 06814 g009aMaterials 15 06814 g009b
Figure 10. Conductor of 70 mm2: (a) Temporal evolution of the temperature along the radial axis considering 28 nodal elements. (b) Temporal evolution of the components of the heat balance equation.
Figure 10. Conductor of 70 mm2: (a) Temporal evolution of the temperature along the radial axis considering 28 nodal elements. (b) Temporal evolution of the components of the heat balance equation.
Materials 15 06814 g010
Figure 11. Conductor of 150 mm2. Experimental versus simulated temperature rise profiles in the three measured points of the cable (conductor-insulation boundary, insulation-jacket boundary and jacket-air boundary) and the relative difference between the experimental and simulated temperatures.
Figure 11. Conductor of 150 mm2. Experimental versus simulated temperature rise profiles in the three measured points of the cable (conductor-insulation boundary, insulation-jacket boundary and jacket-air boundary) and the relative difference between the experimental and simulated temperatures.
Materials 15 06814 g011
Table 1. Conductor dimensions and characteristics.
Table 1. Conductor dimensions and characteristics.
CharacteristicValue
DesignationBarrinax U-1000 R2V
Rated voltage [kVRMS]0.6/1.0
Max voltage [kVRMS]1.2
Max continuous service temperature [°C]90
Short circuit temperature [°C]250
Inner insulation materialXLPE
Inner insulation wall thickness [mm]1.1
Outer jacket materialPVC
Outer jacket wall thickness [mm]1.5
Effective copper cross section [mm2]70
Outer conductor diameter [mm]9.5
Copper resistivity 20 °C [Ohm·m]1.85 × 10−8
Temperature coefficient of resistivity [K−1]0.0043
Number of strands [-]14
Per unit length mass of the conductor [kg/m]0.584
Ambient temperature [°C]19
Table 2. Conductor of 70 mm2. Realized temperature rise tests.
Table 2. Conductor of 70 mm2. Realized temperature rise tests.
StepCurrent (ARMS)Duration (s)
#11201950
#21451950
#31701500
#41851850
#52121350
Table 3. Dimensions and characteristics of the second conductor.
Table 3. Dimensions and characteristics of the second conductor.
CharacteristicValue
DesignationH07RN-F TITANEX 1 × 150
Rated voltage [kVRMS]0.6/1.0
Max voltage [kVRMS]1.2
Max continuous service temperature [°C]90
Short circuit temperature [°C]250
Inner insulation wall thickness [mm]2.0
Outer jacket wall thickness [mm]2.5
Effective copper cross section [mm2]150
Outer conductor diameter [mm]15
Copper resistivity 20 °C [Ohm·m]1.85 × 10−8
Flexibility class5
Per unit length mass of the conductor [kg/m]1.74
Ambient temperature [°C]23.5
Table 4. Conductor of 150 mm2. Realized temperature rise tests.
Table 4. Conductor of 150 mm2. Realized temperature rise tests.
StepCurrent (ARMS)Duration (s)
#1130766
#2163554
#3220850
#4305530
#5405400
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Riba, J.-R.; Llauradó, J. A Model to Calculate the Current–Temperature Relationship of Insulated and Jacketed Cables. Materials 2022, 15, 6814. https://doi.org/10.3390/ma15196814

AMA Style

Riba J-R, Llauradó J. A Model to Calculate the Current–Temperature Relationship of Insulated and Jacketed Cables. Materials. 2022; 15(19):6814. https://doi.org/10.3390/ma15196814

Chicago/Turabian Style

Riba, Jordi-Roger, and Jordi Llauradó. 2022. "A Model to Calculate the Current–Temperature Relationship of Insulated and Jacketed Cables" Materials 15, no. 19: 6814. https://doi.org/10.3390/ma15196814

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