Next Article in Journal
Modeling and Optimization of Natural Gas CCHP System in the Severe Cold Region
Previous Article in Journal
CFD Simulation and Experimental Study on a Thermal Energy Storage–Updraft Solid Waste Gasification Device
Previous Article in Special Issue
Design Aspects of Conical Coaxial Magnetic Gears
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of the Temperature Field Distribution in an EI Type Iron-Cored Coil Using 3D FEM Modeling at Different Load Conditions

by
Denitsa Darzhanova
1 and
Ilona Iatcheva
2,*
1
Department of Electrical Measurements, Technical University of Sofia, 1756 Sofia, Bulgaria
2
Department of Fundamentals of Electrical Engineering, Technical University of Sofia, 1756 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Energies 2023, 16(12), 4581; https://doi.org/10.3390/en16124581
Submission received: 21 April 2023 / Revised: 1 June 2023 / Accepted: 5 June 2023 / Published: 8 June 2023

Abstract

:
In this present paper, results for the temperature distribution in electromagnetic construction in the form of coil coupled with EI-type ferromagnetic core are presented. This construction is typical for small electric transformers (AC operation mode) or chokes—DC or AC operation mode. Investigation has been provided using 3D finite element method computer modeling at various load conditions to simulate power dissipation both in the coil and the core volumes. The results obtained were used to calculate overall thermal resistances toward ambient free air and the thermal resistance between the coil and the core. These results show the important role the thermal resistance between the coil and the core may play for both steady-state and transient device operation.

1. Introduction

Engineering practice often deals with electromagnetic devices in the form of a current-carrying coil, which is surrounded by a ferromagnetic core. These devices vary from the view-point of construction and dimension, starting from low frequency small electric transformers for radio–electronic purpose, medium power air or oil-cooled transformers for public service use [1], some special transformers for medium frequency [2] and ending with high frequency planar transformers for special electronic applications [3]. For all of them, the most important task of the designer is to ensure that the coil average temperature will not exceed certain limits [4], which depends on the selected wire, and more specifically, on the thermal properties of any additional electrical insulation used. This guarantees the reliability and endurance of the relevant device.
An example of that group of devices [5] is presented in Figure 1. It represents the special transformer subgroup family, namely the air-cooled EI laminated transformers. Their assigned power varies mainly in the range of 50 VA up to 1500 VA.
Usually, in steady-state conditions and at a predetermined constant load, calculations for the coil/core temperature of the transformers are based on the thermal equivalent scheme, presented in Figure 2. The concept of thermal resistances [6] has been applied to this scheme, which is also based on the analogy between the heat exchange processes and the electrical circuit theory. The scheme refers to the coil/core temperature rise in relation to the predetermined ambient air temperature T0. Here, P1 is the heat power dissipated in the coil volume, while P2 is the power dissipated in the core volume. The coil temperature rise is denoted by θ1, while θ2 is the corresponding temperature rise of the core. Both of them are considered to be the average calculated values for the corresponding volumes V1 and V2.
In particular, with regard to the temperature rise θ1, it has been observed that for the special transformers group, that the temperature differences within the framework of the coil volume V1 do not exceed 2–3 °C. Thus, at levels θ1 = 80 °C, the calculation accuracy is approximately 3%, which from a practical engineering point of view is considered quite acceptable. From an electrical circuit analyses point of view both P1 and P2 are considered to be “current sources”. R1 and R2 are the thermal resistances responsible for the convection–radiation heat exchange from the coil and the core open surfaces, while R12 is the thermal resistance between the coil and the core. This thermal resistance will include the thermal resistances of the spool walls and those of the inevitable adjacent air layers between the coil and the core.
It is worth noting that different modifications of the equivalent scheme presented in Figure 2 have been reported, which reflect specificity of other design solutions and the author’s preferences. For example, in [7] the ambient medium is expressed through a “voltage (temperature) source”. For that case, the core is enveloped by the transformer windings, and due to the lack of direct contact between the core and ambient air, the thermal resistance R2 has to be omitted. In [8], the authors offer an equivalent scheme for a dry-transformer, where the thermal resistance R12 is omitted. In addition, the two power sources P1 and P2 are combined to form only one (P = P1 + P2), which may lead to calculation inaccuracy when R12 > 0. For the very special design solution of the toroidal inductor [9], an extended equivalent scheme is proposed, with a large number of thermal resistances to reflect the variety of the model areas. Again, a single heat source is included to energize the scheme, which for that case is acceptable due to the unique coil-core construction. That same idea for a single heat source has also been utilized for the investigation of a high MHz frequency planar transformer [3]. For large oil-immersed power transformers, the problem with a single or divided heat source is presented and discussed in [10,11]. Additional considerations for the transformer oil properties as a cooling medium for those transformers are presented and discussed in [12]. In the thermal equivalent scheme proposed to investigate a dry-type 5 kVA transformer [13], a thermal contact resistance R12 between the core and windings exists. This way it separates the core and both windings as heat generating sources. It has to be mentioned here that together with the widely-used-in-engineering thermal equivalent schemes, modern computing tools have already been proposed and introduced [14]. The author’s conclusion in that study is that the implementation of neural networks may help substantially to predict a power transformer’s dynamic thermal behavior.
For the samples such as the one presented in Figure 1, obviously the ideal case from an engineering point of view is when R12 = 0. This will secure maximum cooling efficiency for the coil. Extended analysis provided in [5] has revealed the supplement of the thermal resistance R12 to the temperature rise θ1 and θ2. In [5], calculations of R12 component values and its contribution are based on a one-dimensional [6] “plane wall” simplified for engineering use.
An extended version [15] of the scheme, shown in Figure 2, has been proposed to study the dynamic transient response of the transformer. As presented in Figure 3, thermal capacitances C1/C2 of the coil/core are included in the scheme. This analysis revealed the important role that the resistance R12 may play for the coil temperature time-flow curve, and to what extent this curve will differ from the well-known and widely-accepted-in-engineering practice single exponential curve. In [16], a similar scheme has been proposed to study the transient behaviour of small power pulse transformers, including some nonlinear thermal effects, by converting the current sources into controlled ones. A SPICE subcircuit was created on that ground to help investigations for that special group of transformers.
The main objective in this present paper is, by more accurate calculation of the thermal resistances R1, R2 and R12 using 3D FEM analysis, to reveal the actual distribution of the temperature field in the construction presented in Figure 1, thus validating previously formulated [5,15] conclusions. Next, attention is additionally focused on how and to what extent both θ1 and θ2 depend on the total heat power impact P = (P1 + P2) at the given value of R12. In the broad engineering practice, and some academic teaching courses, a concept [17] is formulated, that there exists a linear functional relationship for θ1(P) and θ2(P). A glance at the equivalent scheme in Figure 2 shows that it will occur only when R12 = 0, otherwise, deviations shall be expected. The rest of the paper is organized as follows. In Section 2, the developed 3D finite element model is described. In Section 3, experimental and numerical results are presented. In Section 4, discussion of the results is carried out.

2. Description of the 3D FEM Model

The 3D FEM model has been developed using COMSOL Multiphysics software [18] and its heat transfer module. Due to the construction symmetry, as illustrated in Figure 4, only 1/8 part of the construction in Figure 1 has been exempted to create the model. A simplified view is also presented in Figure 5, to illustrate some details of the coil placement inside the core and the real construction dimensions in mm. The air gaps δ1 (between the spool and the central core leg) and δ2 (between the winding surface and the external core leg) are also illustrated in this Figure. These two parameters, together with the spool thickness designated as Δs, play an important role for the thermal resistance R12 evaluation.
The number of finite elements in the model is about 700,000 and the number of degrees of freedom is about 1,200,000. The typical error of the solution of the system of linear equations is about 5 × 10−5.
Additionally, Figure 6 helps to clarify how exactly the components of the resistance R12 are specified and then calculated using the model data. The coordinate system X, Y and Z corresponds to the one in Figure 4. In Figure 6, part of the core is omitted and only the core central leg part (volume V2) remains. The main model surfaces used to evaluate the heat flow and temperature are described as follows:
  • SA1 (b,e1,e,b4,b2,b)—free air coil thermal contact (Z-direction) through the spool;
  • SA2 (b,e1,e,f,f1,c)—free air coil thermal contact (XY-direction);
  • SA3 (a,b,b2,a2)—contact between the coil and upper core part (Z-direction);
  • SA4 (b2,c2,d2,a2)—contact between the coil and the core central leg via spool wall (X-direction) and the air gap δ1;
  • SA5 (a,b,c,d)—contact between the coil and external core leg via air gap δ2 (X-direction);
  • SA6 (b2,c2,c4,b4)—contact between the coil and central core leg via spool frame wall (Y-direction).
Based on the above designations, the thermal contact resistances of interest have been calculated as a ratio between the average (by integration provided) temperature of the respective surface and the total heat flow through that same surface:
  • R1—thermal contact via surfaces SA1 and SA2. The thermal resistance R1 is formed by two components R1_1 and R1_2, connected in parallel:
R1_1 = (Tav_SA1 − T0)/qn_SA1;
R1_2 = (Tav_SA2 − T0)/qn_SA2;
R1 = 1/(1/R1_1 + 1/R1_2);
where Tav is the average temperature of the respective surface, qn is the normal thermal flow through the surface and T0 is the ambient temperature;
  • R12—thermal contact via surfaces SA3, SA4, SA5 and SA6. The thermal resistance R_12 is formed by four components R12_1 … R12_4, connected in parallel:
R12_1 = (Tav_SA3 − Tav_SA3_core)/qn_SA3;
R12_2 = (Tav_SA4 − Tav_SA4_core)/qn_SA4;
R12_3 = (Tav_SA5 − Tav_SA5_core)/qn_SA5;
R12_4 = (Tav_SA5 − Tav_SA6_core)/qn_SA6;
R12 = 1/(1/R12_1 + 1/R12_2 + 1/R12_3 + 1/R12_4);
where Tav_SA1,2,3,4 is the average temperature of the respective coil surface; Tav_SA1,2,3,4_core is the average temperature of the respective core surface and qn is the normal thermal flow through the surface. Different materials are present in the gaps between the respective coil and core surfaces: for R12_1 and R12_4 it is only spool material, for R12_2 it is spool material and air and for R12_5—only air;
  • R2—thermal contact between the external core surface and air. The thermal resistance R2 is determined by the expression
R2 = (Tav_core_out − T0)/qn_core_out,
where Tav_core_out is the average temperature of the outer core surface and qn_core_out is the total normal thermal flow through this surface.

3. Experimental and Modeling Results

There is a large variety of AC and DC investigation cases for the sample presented in Figure 1 and its 3D FEM model presented in Figure 4. The DC case meets, e.g., a situation, when a choke of that construction is included in series with the DC welding arc to prevent a sudden severe arc short circuit. Attention has been mainly focused on some specific cases to meet different combinations of heat powers P1 dissipated in the coil and P2 dissipated in the core, together with variation of the air gap δ2. It comes naturally that other (additional) required model parameters, including constructive ones and those reflecting thermal properties of the materials used, have to be fixed. These are specified below:
  • The convection coefficient from the open surfaces of the core and coil: fixed to 10 W/m2·°C;
  • The equivalent thermal conductivity λ in the coil volume: fixed constant λ = 4 W/m·°C;
  • The thermal conductivity in the core volume (considered anisotropic): fixed constant λx,y,z = (44.5;2.5;44.5) W/m·°C;
  • Ambient temperature: T0 = 20 °C;
  • The thermal conductivity of the spool material λ = 0.06 W/m·°C;
  • The thickness of the plastic spool EI 120 wall Δs = 1.8 mm [19];
  • The air gap between spool wall and central core leg (X direction) δ1 = 0.2 mm.
Figure 7 shows the electrical scheme for the experimental measurements provided. The experimental sample, presented in Figure 1, is a small electric transformer with a rated power of 230 VA. A copper enameled wire with a diameter of 0.78 mm has been used for the winding, which consists of two equal sections S1 and S2, each one having 420 turns. The winding together with the plastic spool forms the transformer coil as an individual constructive unit. The total resistance of the two sections connected in series (Figure 7a) at 25 °C was measured to be 6.31 Ω. A thermocouple K type NiCr-Ni is located in the very center between the two sections for temperature rise measurement purposes, and the coil is not impregnated. The same second thermocouple was installed to control the ferromagnetic core surface temperature. The thermocouple sensitivity used in the range up to 100 °C was 40.9 µV/°C, and the readings were digitally measured.
Measures have been taken to keep the heat power impact constant, reflecting the inevitable change of the section S1 and S2 conductor’s resistance.
For verification purpose, experiments have been provided to cover both DC and AC heat impact cases on the sample. The model temperature rise readings are at the points where the experimental thermocouples were located.
  • For the DC experiment (Figure 7a), the two sections S1 and S2 were connected in series. The power P1 was measured by wattmeter W1 and controlled by adjusting the voltage V1. Here at P1 = 16 W, P2 = 0 W, the steady-state experimental data, obtained after 2.5 h and indexed by (E), were θ1E = 53.3 °C and θ2E = 25.6 °C. The corresponding model results, at δ1 = 0.6 mm, and δ2 = 3.5 mm, indexed by (M) are θ1M = 52.35 °C and θ2M = 26.06 °C. Thus, adequate agreement was found.
  • For the AC experiment (Figure 7b), a preliminary no-load test for the power loss in the core was provided in a range of the primary voltage V1 up to 250 V. Table 1 shows the data obtained for the no-load transformer current I0 and the power loss P2 in the ferromagnetic core.
Table 1. Experimental data for AC case at no load.
Table 1. Experimental data for AC case at no load.
U1 [V]5171.490110.5129.8149.9170.2190.6210.8230
I0 [mA]303540496184115156218332
P2 [W]0.901.702.493.694.996.789.2612.4216.1520.25
Under load at rated voltage U1 = 200 V, the difference between the readings of the digital wattmeters W1 and W2 was recorded to get the total heat power impact P = (P1 + P2) on the sample. The digital voltmeter V2 reading was used to recalculate the power P2 based on the relationship P2 = f (V2), having in mind that sections S1 and S2 have an equal number of turns. The test was carried out for 4 h, keeping the total power at the level P = 23 ± 0.5 W. Its components were close to P1 = 8.7 W and P2 = 14.3 W, which were used to find the computer model response. The steady-state temperature rise’s recorded experimental values are θ1E = 56.9 °C and θ2E = 45.6 °C, while the model results are θ1M = 55.0 °C and θ2M = 47.3 °C, so again adequate agreement is considered to exist.
The model load conditions were varied to meet various typical DC and AC load cases, together with different air gap δ2 variation.
In Figure 8 and Figure 9, results for the temperature distribution for two cases, (P1 = 10 W, P2 = 0 W) and (P1 = 20 W, P2 = 20 W), are given for two values of the air gap δ2 (1 and 4 mm).
Several typical cases have been studied by varying the coil power P1, the core power P2 and the air gap δ2. The combination of the studied parameters, together with the results for both average temperature rises θ1 and θ2, are given in Table 2. In the last three columns, the thermal resistance R1, R12 and R2 calculated values are presented.
Temperature distribution for the cases Cs 12, Cs 22 and Cs 33 is shown in Figure 10.
For the Cs 2 case series, at a load typical for a transformer presented in Figure 1, the variation of the temperature rise along the x axis at z = 0 and y = 0 is displayed in Figure 11. The coordinate x = 0 corresponds to the geometry centre of the core central leg, while x = 60 mm meets the free air open surface of the core external leg. The four curves presented cover air gap values δ2 = 1, 2, 3 and 4 mm. The average temperature rises fixed at a gap value δ2 = 3 mm are θ1 = 67.88 °C and θ2 = 51.79 °C.
Some additional data in Table 3 illustrate the temperature rise θ at key points shown in Figure 6.
In Table 4, additional results are shown to check the model response at different combinations of AC load. For this extended Cs 22 case, both air gaps are kept constant (δ1 = 0.2 mm, δ2 = 3 mm). The total heat impact P on the construction increases, but its components P1 and P2 take a different distribution. This approach is illustrated by two specific sub-cases, namely:
Cs 22 A—the ratio between P1 and P2 maintained constant—P1/P2 = 1.66.
Cs 22 B—the coil power P1 remains practically constant, and only the core power P2 increases. This variant meets two load cases that may happen practically: DC choke case (P1 = 8 W, P2 = 0), and AC case—designed under different quality of the selected transformer core steel (P2 = 6; 13; 22; 38 W).

4. Discussion

The construction in Figure 1 presents a complex case to study from a heat transfer point of view. It is a non-homogeneous 3-dimensional body, containing:
  • Two main heat-generating parts: the winding itself (including enameled copper wire, dielectric insulation paper etc.) and the iron core around it (produced by use of ferromagnetic electro technical steel lamellae);
  • Additional materials: the coil plastic spool;
  • Technological airgaps between the coil and the core.
Convection is responsible for that complex body cooling from all open external surfaces. Conduction heat transfer exists between the coil, the spool and the core. The coil temperature evaluation is vital in order to observe standard recommendations of ISO 2578 [4] for the copper conductors’ electrical insulation. This temperature is considered as the average value for the entire coil volume V1 (Figure 2) due to small temperature differences observed in the volume V1 (of about 2 ÷ 3 °C)—a result of the coil’s usually high thermal equivalent conductivity, as mentioned in the introduction and based on the authors’ experiences.
The equivalent scheme for engineering calculation in Figure 2 may serve to study not only the steady-state behavior [5] of the electromagnetic devices such as the one in Figure 1. If extended by adding thermal capacitances of the coil and the core [15], the modified equivalent scheme presented in Figure 3 may also be used to study different dynamic mode operations. However, here, again, a question is raised for the calculation accuracy of the thermal resistance R1, R2 and R12 values. Results for those resistances extracted from the 3D FEM analysis and presented in Table 2 can be considered as supervision of the simplified 1D thermal field approach reported in [15]. These results for R1, R2 and R12 show:
  • As expected, they are not influenced by the heat dissipated in the coil and core volumes.
  • Thermal resistance R1 only slightly depends on the air gap δ2. It grows by 15% when the gap δ2 substantially increases from 1 to 4 mm.
  • Thermal resistance R12 increases, as expected, when the gap δ2 changes from 1 to 4 mm. For the case Cs 1, this change reaches almost 30%.
  • Comparing the values of R1, R2 and R12 in Table 2 with the corresponding values reported in [5] shows:
    • No difference observed for resistance R2.
    • The R1 resistance evaluated in [5] is R1 = 9.1 °C/W, i.e., is overestimated by some 15–20%.
    • The R12 resistance evaluated in [5] is R12 = 2.0 °C/W, i.e., is underestimated by 14% for the case Cs 0, and by 33% for the case Cs 03.
  • Comparing the obtained values of R12 with the values of the thermal resistances between the core and the two windings of a planar transformer obtained in [16] shows that the values are of the same order, even though the construction of the planar transformer is different from the one studied in this present paper.
The temperature distribution in the coil volume, as presented in Table 3, looks quite reasonable. For the case Cs 22, the powers P1 = 16 W and P2 = 10 W are uniformly distributed in the corresponding volumes of the coil and the core. The temperature hot spot is at the centre of the coil (X direction—points d and d2). The difference of 0.5 °C between these two points can be explained by a slightly better cooling condition through the central core leg with respect to the one through the air gap δ2 = 3 mm. A more substantial difference (more than 3 °C) is observed between the points (d, d2) and the free (open-to-air) coil surface SA2. It can be explained by a more favored path for the heat flow from the coil volume enclosed between the core and the free surface SA2.
The diagram depicted in Figure 12 illustrates the results presented in Table 4. For sub-case A, including the point θ1 = 0 and θ2 = 0 at P = 0, the run of the curves θ1(P) and θ2(P) is practically linear. It is not so for sub-case B. The model data obtained were additionally processed mathematically, using the Least Square Method. Results show deflections from a virtual approximation straight line of the order up to 40%. The two curves θ1(P) and θ2(P) even cross each other at the point P = 32 W.

5. Conclusions

  • The data presented in Table 2 show that, expressed in %, the thermal resistance R12 may reach values of about 35–38% of the resistance R1 value. Thus, results drawn in [5,15], but based on simplified engineering calculation of the resistances R12 and R1, are additionally and precisely supported by the use of modern and more accurate calculations. In this way, without any doubt, it is claimed that the role ratio R12/R1 plays for both the steady-state and dynamic behavior of the iron-cored coils under analyses above may become significant.
  • Regarding the assumed linear flow concept suggested in [17] for the curves θ1(P), FEM analysis provided proof that serious deviations may occur when the thermal resistance R12 takes figures about 20–30% of the thermal resistance R1 value. Referring to Figure 2, theoretically, linearity may take place only when:
    (a)
    R12 = 0; this is an ideal heat exchange case between the coil and the core;
    (b)
    Heat powers dissipated in the coil P1 and the core P2 simultaneously increase, but the ratio P1/P2 remains practically the same (dotted lines presented in Figure 12).

Author Contributions

Conceptualization, D.D. and I.I.; methodology, D.D. and I.I.; validation, D.D. and I.I.; formal analysis, D.D. and I.I.; investigation, D.D. and I.I.; resources, D.D.; data curation, D.D. and I.I.; writing—original draft preparation, D.D.; writing—review and editing, I.I.; visualization, D.D. and I.I.; supervision, D.D. and I.I.; project administration, D.D. and I.I.; funding acquisition, D.D. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by the Research and Development Sector at the Technical University of Sofia.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the Research and Development Sector at the Technical University of Sofia for the financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Haritha, V.V.S.S.; Rao, T.R.; Jain, A.; Ramamoorty, M. Thermal Modeling of Electrical Transformers. In Proceedings of the 16th National Power Systems Conference, Hyderabad, India, 15–17 December 2010. [Google Scholar]
  2. Mogorovic, M.; Dujic, D. Thermal modeling and experimental verification of an air cooled medium frequency transformer. In Proceedings of the 2017 19th European Conference on Power Electronics and Applications (EPE’17 ECCE Europe), Warsaw, Poland, 11–14 September 2017. [Google Scholar] [CrossRef] [Green Version]
  3. Shen, Z.; Xu, B.; Liu, C.; Hu, C.; Liu, B.; Xu, Z.; Jin, L.; Chen, W. The Modeling and Simplification of a Thermal Model of a Planar Transformer Based on Internal Power Loss. Sustainability 2022, 14, 11915. [Google Scholar] [CrossRef]
  4. ISO 2578:1993; Plastics—Determination of Time-Temperature Limits After Prolonged Exposure to Heat. International Organization for Standardization: Geneva, Switzerland, 1993.
  5. Darzhanova, D.; Darjanov, P. Extended Analysis for the Static Thermal Response of an Iron-cored Winding at AC Mode of Operation. In Proceedings of the 2020 21st International Symposium on Electrical Apparatus and Technologies SIELA, Bourgas, Bulgaria, 3–6 June 2020. [Google Scholar] [CrossRef]
  6. Holman, J.P. Heat Transfer; McGraw-Hill Book Company: Columbus, OH, USA, 1989. [Google Scholar]
  7. Purushothaman, S.; de Leon, F. Heat-Transfer Model for Toroidal Transformers. IEEE Trans. Power Deliv. 2012, 27, 813–820. [Google Scholar] [CrossRef]
  8. Ghareh, M.; Sepahi, L. Thermal Modeling of Dry-Transformers and Estimating Temperature Rise. Int. J. Electr. Comput. Eng. 2008, 2, 9. [Google Scholar]
  9. Kyaw, P.A.; Delhommais, M.; Qiu, J.; Sullivan, C.R.; Schanen, J.-L.; Rigaud, C. Thermal Modeling of Inductor and Transformer Windings Including Litz Wire. IEEE Trans. Power Electron. 2019, 35, 867–881. [Google Scholar] [CrossRef]
  10. Susa, D. Dynamic Thermal Modelling of Power Transformers. Ph.D. Dissertation, Department of Electrical and Communications Engineering, Helsinki University of Technology, Espoo, Finland, 2005. [Google Scholar]
  11. Susa, D.; Lehtonen, M. Dynamic Thermal Modeling of Power Transformers: Further Development—Part II. IEEE Trans. Power Deliv. 2006, 21, 1971–1980. [Google Scholar] [CrossRef]
  12. Swift, G.; Molinski, T.; Lehn, W. A Fundamental Approach to Transformer Thermal Modeling—Part I: Theory and Equivalent Circuit. IEEE Trans. Power Deliv. 2001, 16, 171–175. [Google Scholar] [CrossRef]
  13. Mamizadeh, A.; Iskender, I. Winding Thermal Analyzing of Dry Type Transformers. Int. J. Tech. Phys. Probl. Eng. 2010, 2, 80–84. [Google Scholar]
  14. Bragone, F.; Morozovska, K.; Hilber, P.; Laneryd, T.; Luvisotto, M. Physics-informed neural networks for modelling power transformer’s dynamic thermal behaviour. Electr. Power Syst. Res. 2022, 211, 108447. [Google Scholar] [CrossRef]
  15. Darjanov, P.; Darzhanova, D. Dynamic Performance Analysis of an Iron-cored Winding at AC Mode of Operation. In Proceedings of the 2020 21st International Symposium on Electrical Apparatus and Technologies SIELA, Bourgas, Bulgaria, 3–6 June 2020. [Google Scholar] [CrossRef]
  16. Górecki, K.; Detka, K.; Górski, K. Compact Thermal Model of the Pulse Transformer Taking into Account Nonlinearity of Heat Transfer. Energies 2020, 13, 2766. [Google Scholar] [CrossRef]
  17. Balian, R.H. Transformers for Radioelectronics; Soviet Radio: Moscow, Russia, 1971; pp. 224–229. (In Russian) [Google Scholar]
  18. COMSOL Inc. COMSOL Multiphysics 5.5 User’s Guide; COMSOL Inc.: Burlington, MA, USA, 2019. [Google Scholar]
  19. CARTOPLAST SRL. EI 120 Lamination 100 × 120. Available online: www.cartoplastsrl.it/cp_english/monofase.html (accessed on 20 May 2023).
Figure 1. Studied transformer.
Figure 1. Studied transformer.
Energies 16 04581 g001
Figure 2. Equivalent scheme for the steady-state thermal analyses.
Figure 2. Equivalent scheme for the steady-state thermal analyses.
Energies 16 04581 g002
Figure 3. Thermal equivalent scheme for transient mode of operation.
Figure 3. Thermal equivalent scheme for transient mode of operation.
Energies 16 04581 g003
Figure 4. Model of the whole device, together with studied 1/8 part of it.
Figure 4. Model of the whole device, together with studied 1/8 part of it.
Energies 16 04581 g004
Figure 5. Schematic presentation of the core/coil dimensions of the studied system.
Figure 5. Schematic presentation of the core/coil dimensions of the studied system.
Energies 16 04581 g005
Figure 6. Main coil heat exchange surfaces.
Figure 6. Main coil heat exchange surfaces.
Energies 16 04581 g006
Figure 7. Electrical experimental study measurement schemes: (a) DC case and (b) AC case.
Figure 7. Electrical experimental study measurement schemes: (a) DC case and (b) AC case.
Energies 16 04581 g007
Figure 8. Temperature distribution for the case P1 = 10 W, P2 = 0 W: (a) air gap δ2 = 1 mm; (b) air gap δ2 = 4 mm.
Figure 8. Temperature distribution for the case P1 = 10 W, P2 = 0 W: (a) air gap δ2 = 1 mm; (b) air gap δ2 = 4 mm.
Energies 16 04581 g008
Figure 9. Temperature distribution for the case case P1 = 20 W, P2 = 20 W: (a) air gap δ2 = 1 mm; (b) air gap δ2 = 4 mm.
Figure 9. Temperature distribution for the case case P1 = 20 W, P2 = 20 W: (a) air gap δ2 = 1 mm; (b) air gap δ2 = 4 mm.
Energies 16 04581 g009
Figure 10. Temperature distribution for the cases from Table 1: (a) Cs 12; (b) Cs 22; (c) Cs 32.
Figure 10. Temperature distribution for the cases from Table 1: (a) Cs 12; (b) Cs 22; (c) Cs 32.
Energies 16 04581 g010aEnergies 16 04581 g010b
Figure 11. Variation of the temperature rise θ(x) at y = 0, z = 0, x = 0–60 mm for the case series Cs 2.
Figure 11. Variation of the temperature rise θ(x) at y = 0, z = 0, x = 0–60 mm for the case series Cs 2.
Energies 16 04581 g011
Figure 12. Temperature rise θ1 and θ2 vs total power P dissipated, Table 4, case Cs22 A and B.
Figure 12. Temperature rise θ1 and θ2 vs total power P dissipated, Table 4, case Cs22 A and B.
Energies 16 04581 g012
Table 2. Average coil and core temperature rise and thermal resistances for typical cases.
Table 2. Average coil and core temperature rise and thermal resistances for typical cases.
CaseP1 [W]P2 [W]δ2 [mm]θ1 [°C]θ2 [°C]R1 [°C/W]R12 [°C/W]R2 [°C/W]
Cs 0100128.8717.356.762.323.13
Cs 01230.4917.277.082.703.13
Cs 02331.6417.427.422.873.13
Cs 03432.6517.657.792.973.13
Cs 11010146.1540.616.752.453.12
Cs 11247.7140.887.062.893.13
Cs 12349.0041.307.403.083.13
Cs 13450.2541.777.773.173.12
Cs 21610163.3351.086.762.343.13
Cs 21265.8851.297.072.763.13
Cs 22367.8851.797.412.953.12
Cs 23469.7352.407.783.053.13
Cs 32010174.7558.076.762.303.13
Cs 31277.9758.247.082.723.13
Cs 32380.4358.807.422.903.12
Cs 33482.6959.497.783.003.12
Cs 42020192.0781.316.752.353.13
Cs 41295.2181.847.062.813.12
Cs 42397.8282.667.403.003.13
Cs 434100.3183.607.773.103.12
Table 3. Temperature rise θ [degC] at some key points from Figure 6 for the case Cs 22.
Table 3. Temperature rise θ [degC] at some key points from Figure 6 for the case Cs 22.
Key Point (Figure 6)
aa1a2a3a4bb1b2b3b4ee1
68.151.767.552.752.766.150.066.565.965.064.664.8
Key point (Figure 6)
cc1c2c3c4dd1d2d3d4ff1
68.449.768.668.267.370.251.669.753.253.166.967.2
Table 4. Average coil and core temperature rise θ1 and θ2 at various total AC load P.
Table 4. Average coil and core temperature rise θ1 and θ2 at various total AC load P.
Case Cs 22 ACase Cs 22 BTotal Power P = P1 + P2 [W]
P1[W]P2 [W]θ1 [°C]θ2 [°C]P1[W]P2 [W]θ1 [°C]θ2 [°C]
5321.115.98025.313.98
10642.031.710642.031.716
14.38.760.345.7101354.248.523
201283.963.6102269.669.432
3018125.695.5103898.0108.048
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

Darzhanova, D.; Iatcheva, I. Investigation of the Temperature Field Distribution in an EI Type Iron-Cored Coil Using 3D FEM Modeling at Different Load Conditions. Energies 2023, 16, 4581. https://doi.org/10.3390/en16124581

AMA Style

Darzhanova D, Iatcheva I. Investigation of the Temperature Field Distribution in an EI Type Iron-Cored Coil Using 3D FEM Modeling at Different Load Conditions. Energies. 2023; 16(12):4581. https://doi.org/10.3390/en16124581

Chicago/Turabian Style

Darzhanova, Denitsa, and Ilona Iatcheva. 2023. "Investigation of the Temperature Field Distribution in an EI Type Iron-Cored Coil Using 3D FEM Modeling at Different Load Conditions" Energies 16, no. 12: 4581. https://doi.org/10.3390/en16124581

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