Next Article in Journal
Construction of Nitrogen Content Observer for Fuel Cell Hydrogen Circuit Based on Anode Recirculation Mode
Next Article in Special Issue
Performance Evaluation of Stator/Rotor-PM Flux-Switching Machines and Interior Rotor-PM Machine for Hybrid Electric Vehicles
Previous Article in Journal
Location of the Interurban Fast Charging Infrastructure for Electric Vehicles Using the Methodology for Calculating the Maximum Distance between Fast Charges (MDFC) and Simulation: A Case Study in Ecuador
Previous Article in Special Issue
Designing High-Power-Density Electric Motors for Electric Vehicles with Advanced Magnetic Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal Analysis of a Flux-Switching Permanent Magnet Machine for Hybrid Electric Vehicles

School of Electrical Engineering, Southeast University, Nanjing 210096, China
*
Author to whom correspondence should be addressed.
World Electr. Veh. J. 2023, 14(5), 130; https://doi.org/10.3390/wevj14050130
Submission received: 24 April 2023 / Revised: 12 May 2023 / Accepted: 17 May 2023 / Published: 19 May 2023

Abstract

:
This paper investigates the loss and thermal characteristics of a three-phase 10 kW flux-switching permanent magnet (FSPM) machine, which is used as an integrated starter generator (ISG) for hybrid electric vehicles (HEVs). In this paper, an improved method considering both DC-bias component and minor hysteresis loops in iron flux-density distribution is proposed to calculate core loss more precisely. Then, a lumped parameter thermal network (LPTN) model is constructed to predict transient thermal behavior of the FSPM machine, which takes into consideration various losses as heat sources determined from predictions and experiments. Meanwhile, a simplified one-dimensional (1D) steady heat conduction (1D-SHC) model with two heat sources in cylindrical coordinates is also proposed to predict the thermal behavior. To verify the two methods above, transient and steady thermal analyses of the FSPM machine were performed by computational fluid dynamics (CFD) based on the losses mentioned above. Finally, the predicted results from both LPTN and 1D-SHC were verified by the experiments on a prototyped FSPM machine.

1. Introduction

With the energy dilemma and environmental pollution becoming worse, electric-powered vehicles have attracted considerable attention due to their lower green gas emission and oil consumption [1]. Recent hybrid electric vehicles (HEVs) provide long-distance operation due to the optimal match between an engine powered by oil and electric motors powered by electricity stored in batteries. HEVs are widely researched by academics and have become dominant commercial products in EV markets [2,3]. For micro-hybrid EVs, an integrated starter generator (ISG) is the key component of both driving and generating systems, since it plays two major roles, namely, as a starter and a generator. Due to space and weight limitations, the rotor of an ISG is normally directly coupled to the flywheel of the engine in HEVs, where high torque (power) density, large overload torque capability, and high efficiency are expected. Hence, the flux-switching permanent magnet (FSPM) machine is considered as a promising candidate to be applied in electric vehicles, and aerospace and ship propulsion due to its high torque (power) density, high efficiency, and compact structure [4].
With the ever-increasing demand for power (torque) density, research on machine loss and temperature has become a hot topic. An accurate thermal model is an essential tool not only at the machine design stage but also for online prediction of temperature distribution [5]. While finite element method (FEM)-based and computational fluid dynamics (CFD)-based thermal models can achieve high accuracy, a lumped parameter thermal network (LPTN)-based model is often preferred thanks to its lower computational requirement and good accuracy [6,7,8].
In [9], the thermal influence of vehicle integration on the thermal load of an ISG was discussed by a FEM-based thermal model. In [10], an axially segmented FEM model of a FSPM machine was proposed to analyze the coupled electromagnetic–thermal performances. A thermal resistance network was established based on a nine-node model for an interior PM (IPM) machine and the transient temperature characteristics were obtained [11]. In [12], a numerical approach for estimation of convective heat transfer coefficient in the end region of an ISG was proposed, and both the local and averaged heat transfer coefficients were estimated. A systematic procedure to study the impact of each thermal phenomenon in IPM machines used for ISG was presented in [13]. In [14], a reduced model in a multi-physical electric machine optimization procedure was proposed.
The contribution of this paper is to propose two temperature prediction models for a 10 kW FSPM machine as an ISG for micro-hybrid vehicles; namely, a LPTN thermal model, and a one-dimensional (1D) steady heat conduction (SHC) (1D-SHC) model. The two methods can both quickly predict the internal temperature distribution of the FSPM machine. The results were verified by CFD and experiments to prove their accuracy.
Section 2 will propose an improved core loss model considering both the DC-bias component and minor hysteresis loops in iron flux-density, and the core loss of the FSPM machine is calculated and verified by experiments. Then, in Section 3 a LPTN model is proposed firstly to predict transient thermal behavior. After that, a simplified 1D-SHC thermal model is proposed to reveal the relationship between design parameters of a cooling jacket and thermal distribution of stator, and verified by experiments under different cooling conditions. In Section 4, both steady and transient thermal predictions are compared with those from ANSYS fluent-based CFD. Experiments with rising temperatures were conducted on a prototyped FSPM machine and are detailed in Section 5, followed by conclusions in Section 6.

2. Loss Prediction Model

The key dimensions of the studied FSPM machine are listed in Table 1. In addition to electromagnetic parameters, the thermal conductivities, the specific heats, and densities of materials used for transient thermal analysis are presented in Table 2. The employed PM material was N35 and the silicon steel sheet was 35WW310.
The loss of the FSPM machine includes winding joule loss, core loss, eddy current loss in PMs, housing and frictional loss, and excess loss. According to the Bertotti G. model [15], the core loss of PM machines PFe consists of hysteresis loss, eddy current loss and excess loss, and the core loss yields:
P F e = P h + P c + P e = k h f B m α + k c f 1.5 B m 1.5 + k e f 2 B m 2
where Ph is hysteresis loss in W, Pc is the classical eddy current loss in W, Pe is the excess loss in W, kh, kc, and ke are the corresponding coefficient of the above losses, respectively, f is the fundamental frequency of a magnetizing flux in Hz, and Bm is the maximum flux density in core in T.
However, Equation (1) only works given a purely sinusoidal magnetizing flux. To exactly obtain the magnetizing flux-density characteristics in the FSPM machine core, eight key points located in stator and rotor respectively are selected, as shown in Figure 1a. Correspondingly, the resultant loci of the flux-density radial and tangential components (Bgr/Bgt) are predicted, as shown in Figure 1b. Clearly, for the stator points 1 and 2, the surrounded areas by the Bgr/Bgt loci are to be almost zero, which means the averaged Bgr/Bgt values are nearly zero. However, for points 3 and 4, the corresponding areas (the blue one and the pink one) are not centrosymmetric, which means a DC-biased component exists. For the points 5~8 in the rotor, the Bgr/Bgt loci are all centrosymmetric and the averaged values are close to zero. A typically DC-biased component and a minor hysteresis loop are shown in Figure 1c,d, respectively.
Unfortunately, the influence of magnetized DC-biased components and minor hysteresis loops are not well recognized in the commercial FEM software packages [16]. Hence, to predict the core loss of FSPM machines more precisely, an improved model considering both DC-biased component and minor hysteresis loops is proposed as follows. Assuming that the minor loop is similar to the major loop, the core loss can be divided into radial and tangential components. The total core loss, including the hysteresis loss [17], eddy current loss [18], and excess loss [19], can be obtained as follows:
P h = k h f L a i = 1 N e l e m Δ A i [ ( j = 1 N p r i B r m i j 2 ) ε ( Δ B r ) + ( j = 1 N p t i B t m i j 2 ) ε ( Δ B t ) ]
P c = K c 2 π 2 L a N s e t p k = 1 N s t e p i = 1 N e l e m Δ A i [ ( B r m i k + 1 B r m i k Δ t ) 2 + ( B t m i k + 1 B t m i k Δ t ) 2 ]
P e = K e L a N s e t p k = 1 N s t e p i = 1 N e l e m Δ A i [ ( B r m i k + 1 B r m i k Δ t ) 1.5 + ( B t m i k + 1 B t m i k Δ t ) 1.5 ]
P F e = P h + P c + P e
where Nelem is the finite elements number, ΔAi is the ith finite element area in m2, Npri and Npti are the radial and tangential minor loops numbers of the ith element during one period, respectively, Nstep is the calculation steps number, Brmij and Btmij are the maximum radial and tangential flux-densities of the jth hysteresis loop in the ith element in T, respectively, Brmik and Btmik are the maximum radial and tangential flux-density of the ith element in the kth calculation step in T, and Δt is the time step in s.
According to Equations (2)–(5), the core loss can be obtained by a combination platform of ANSYS and MATLAB, where based on ANSYS the detailed Bgr/Bgt results of each meshed iron element can be acquired, and based on MATLAB the core loss versus rotating speeds under different conditions can be assessed by a series of data processing calculations according to Equations (2)–(5). The no-load core loss density distribution of the stator and rotor is shown in Figure 2. Consequently, the predicted core losses versus rotor speed are compared with those obtained by commercial software, e.g., by JMAG and ANSYS EM as shown in Figure 3. It can be seen that the core losses obtained by the improved method are slightly higher than those from software, which validates the influence of the DC-biased component and minor hysteresis loop, and also validates the feasibility of the improved core loss prediction method.
In addition, the no-load eddy current density distribution derived by 3D-FEM is shown in Figure 4. The resulting no-load eddy loss in PMs Ppmc and housing Phc versus rotating speeds are shown in Figure 5. It can be found that with the increase of the speed, the eddy current losses in PMs and housing increase gradually, which is caused by the air-gap harmonic fields and can be calculated by Equation (6) [20],
P e d d y = 1 T t c i = 1 k J e 2 Δ A i σ r 1 L a d t
where Peddy is the eddy current loss in PM and housing in W, Je is the current density in each element in A/m2, ΔAi is the ith element area in m2, σr is the conductivity of the eddy current zone in S/m, and tc is the time corresponding to a period in each element in s.
The frictional loss Pfri of the FSPM machine yields [21]:
P f r i = 16 N r ( v r 40 ) 3 L a 19 × 10 3
where vr is the rotor peripheral speed in m/s. It is found that Pfri = 10.7 W under the rated speed of 1000 r/min.
Finally, a no-load test under different rotation speeds is conducted on a prototyped FSPM machine to verify the predicted results, where the machine is controlled by a DSP-based controller and the input power is obtained by a power analyzer. The frictional loss is so small that it can be neglected. Therefore, the input power is equal to the total loss. The total losses versus rotor speeds by different methods are compared in Figure 6. Compared with the results obtained by commercial software, the predicted core losses derived by the improved method agree with the measurements with the smallest deviations.

3. Two Thermal Models

For the prototyped FSPM machine, a circumferential water jacket with one cooling duct is introduced in the stator housing as shown in Figure 7. The coolant channel of the casing adopts a single-layer water jacket cooling structure. Figure 7a shows the schematic diagram of the FSPM machine structure. Figure 7b–d show the housing, coolant flow path, and machine assembly, respectively. Figure 7b shows the cross-sectional diagram of the machine, and Figure 7c shows the cross-sectional diagram of the cooling duct. Arrows are used in Figure 7c to indicate the flow path of the coolant. The blue arrow represents the low temperature coolant near the inlet, while the red arrow represents the coolant that has been heated through heat exchange. The fluid running inside the cooling duct can be modeled as the movement of fluid in a rectangular channel using dimensionless numbers. Consequently, the convection coefficient can be obtained in the following stages.
Firstly, with the cooling jacket cross-section in Figure 7c, the Prandtl number of the fluid Prf yields
P r f = c f η f λ f
where cf is the specific heat capacity of fluid in J/(kg·°C), ηf is fluid dynamic viscosity in N·s/m2, and λf is the fluid thermal conductivity in W/(m·°C).
Secondly, the Reynolds number of fluid Re yields
R e = υ d e v f
where υ is the velocity of fluid in m/s, vf is the fluid kinetic viscosity in m2/s, and de is the hydraulic radius in m by Equation (10).
d e = 4 A c s s = 4 b h 2 ( b + h )
where Acs is a cross-section area of a single cooling duct in m2. s, b, and h are the wetted perimeter, width, and height of cooling duct in m, respectively.
Approximately, in a circumferential cooling duct, the velocity of fluid can be figured out as
υ = Q A c s
where Q is the fluid quantity in kg/s.
According to the value of Re, the fluid flow can be divided into turbulence flow and laminar flow. The Nusselt number of the laminar flow Nufl yields [22],
N u f l = 0.644 ( R e 0.5 ) P r f 1 3
For turbulence flow, the Nusselt number Nuft yields
N u f t = 0.023 ( R e 0.8 ) P r f 0.4 ( η f η w ) 0.14
where ηw is the dynamic viscosity of housing in N·s/m2.
Based on the similarity criterion of fluid [22], the convection heat transfer coefficient hf0 yields
h f 0 = N u f λ f d e
Considering that turbulence flow shows a better heat dissipation than laminar flow, the former is employed in the cooling jacket, where the convection heat transfer coefficient hf of the cooling jacket is affected by the geometric parameters and the velocity of the fluid, and can be given by
h f = λ f b + h 2 b h ( 2 Q v f ( b + h ) ) 0.8 P r f 0.4 ( η f η w ) 0.14
From Equation (15), when the fluid quantity keeps constant, as the cross-sectional area of the cooling ducts decreases, the convection coefficient of the cooling jacket increases. However, a small cross-section cooling duct is not only difficult to manufacture, but also may lead to high inlet velocity and high hydraulic pressure, which causes the corrosion of cooling duct and deteriorates the operation stability. In the following, two thermal models are proposed, one a LPTN model and the other a 1D-SHC model.

3.1. Lumped Parameter Thermal Network Model

A LPTN model enables the heat flow and the temperature distribution inside the machine by means of an equivalent thermal circuit, which is composed of heat sources, thermal resistances, and thermal capacitances. For the convenience of calculation and improved accuracy, three assumptions are made as follows [23,24,25]:
  • Symmetrical temperature distribution and the same cooling conditions along the circumference;
  • Uniformly distributed thermal capacity and heat generation;
  • Independent heat flow in radial and axial directions
To simplify the calculation load, only 1/24 of the FSPM machine is modeled as shown in Figure 8 due to symmetry, where the heat sources including stator/rotor core losses, PM/housing eddy current losses, and windings joule loss are considered. The thermal resistances and capacitances can be determined according to the machine geometry and the physical properties of materials. Table 3 and Table 4 list the corresponding resistances and capacitances of the LPTN model. A preliminary selection of the resistances and capacitances can be determined according to the machine geometry and physical properties of the materials used [26].
With the convection heat transfer coefficient, the thermal resistance Rconvi (i = 1, 2, 3) representing the heat dissipation by cooling medium convection between the housing external surface and ambient can be calculated by Equation (16) [25],
R c o n v i = 1 h c o n v i A c o n v i
where Rconvi is the thermal resistance due to convection heat transfer in °C/W, hconvi is the convection heat transfer coefficient in W/(m2·°C), and Aconvi is the convective area in m2. Here, the area of the end-part winding is considered.
Since the heat exchange between stator and rotor through the air-gap is assumed to be only by convection, Equation (16) is also used for the calculation of the thermal resistances Rairi.
In addition to heat convection, heat conduction is also an important way for heat dissipation. The resistance representing the heat flow in the radial direction is modeled by using Equation (17),
R r a d i a l = l n ( r o / r i ) 2 π λ L
where ro and ri are the outer and inner diameter of the cylinder in m, λ is the thermal conductivity of the material in W/m/°C, and L is the cylinder length in m.
In the tangential direction, the thermal resistance due to conduction heat transfer is given by,
R t a n g e n t i a l = l λ A c o n d
where l is a portion length of the path considered in m, and Acond is the area for the conduction in m2.
Figure 8a,b show the 3D module structure and modular stator element of the FSPM machine. Based on the thermal resistances above, a thermal resistance network of the 1/24 machine is constructed as shown in Figure 8c, where Rsh1-Rsh8, Rsy1-Rsy6, Rair1-Rair9, Rst1-Rst3, Rcoil1-Rcoil3, and Rpm1-Rpm5 represent the thermal resistances of the housing, stator yoke, air-gap, stator tooth, stator winding coils, and PMs. Rrt1, Rry1, and Rshaft1 represent the thermal resistances of rotor tooth, rotor yoke, and shaft, respectively. Csh1-Csh3, Csy1-Csy2, Cst1, Ccoil1, Cpm1-Cpm2, and Crt represent the thermal capacitances of the housing, stator yoke, stator tooth, winding coils, PMs, and rotor tooth. Here, since the heat dissipated by forced convection is much larger than that by radiation, the radiation heat dissipation is ignored.
Under two typical operation conditions, i.e., the speed of 1000 r/min and the phase current of 30.7 A (RMS) and 60 A (RMS), the transient temperature rises of different components under forced water cooling are obtained by the LPTN model, and the results are shown in Figure 9. It takes around 70 min for the machine under water cooling to reach a thermal steady state where the armature windings have the highest temperatures (45.7 °C@30.7 A and 82.8 °C@60 A under water cooling). In addition, since the PMs are mounted on the stator, the temperature of the PMs is very close to that of the stator core, exhibiting the advantage of FSPM machines.

3.2. One-Dimensional Steady Heat Conduction Model

Generally, LPTN and FEM have been widely employed in thermal analysis of electrical machines. However, these two methods are normally time-consuming and require complicated modeling. For water-cooling machines, in order to select a reasonable flow rate of coolant, a 1D-SHC approach is proposed based on heat transfer and fluid mechanics as shown in Figure 10, and the relationship between the internal temperature of the stator and the coolant flow rate and coolant temperature is obtained.
It should be noted that the 1D-SHC model is based on the following assumptions. (1) The loss is uniformly distributed in each component of the machine. (2) The heat generated by joule loss and stator core loss is only dissipated by the housing. (3) The stator laminations, windings, and PMs are simplified to a homogeneous heating unit and the equivalent averaged thermal conductivity λavg yields [26,27]:
λ a v e = A s + A w i n d + A p m A s λ s + A w i n d λ c u + A p m λ p m
where As, Awind, and Apm are the cross-section area of the stator, windings, and PMs in m2, respectively, λs, λwind, and λpm are the thermal conductivity of stator, windings, and PMs in W/(m·°C).
The winding and insulation layering is used to calculate the thermal conductivity of the stator windings. Figure 11 shows the equivalent diagram of the winding structure, where the insulation and windings are arranged with intervals. The slot filling factor is set as 0.35 according to the prototyped machine. The equivalent winding thermal conductivity yields:
λ w i n d = i = 1 n δ i / i = 1 n δ i λ i
where δI is the thickness of the ith layer in m, λi is the thermal conductivity of the ith layer in W/(m·°C).
Then, the equivalent volumetric heat generation of stator qVs and rotor qVr can be obtained by
{ P e a v e = P s V s + P c u V c u + P p m V p m V s + V c u + V p m q V s = P e a v e V e q u   q V r = P r V r  
where Peave is equivalent average loss in W, Ps, Pcu, Ppm, and Pr are the stator core loss, winding joule loss, eddy current loss in PMs, and rotor core loss in W, respectively, Vs, Vcu, Vpm, Vequ, and Vr are the volume of stator, winding, PM, equivalent stator, and rotor in m3, respectively.
Since the thermal model is simplified into a 1D-SHC model, the heat flux density of stator/rotor (qs/qr) yields
{ q s = q V s / S s q r = q V r / S r
where Ss/Sr is the cross-section area of stator/rotor lamination in m2.
The inlet and outlet temperature of the cooling fluid can be detected by a hand-held infrared thermometer. Thus, the temperature of the fluid Tf equals
T f = T i + T o 2
where Ti/To is the inlet/outlet temperature of the fluid in °C.
According to the 1D thermal circuit in Figure 10b, the temperature of housing Tsh can be derived by
T s h = T f + q r π h f ( R s h o R s h a o ) + q s π h f ( R s h o R s i )
where hf is the fluid convection coefficient in W/(m2·K), Rsho is the housing outer radius in m, Rshao is the shaft outer radius in m, and Rsi is the stator inner radius in m.
The temperature of stator yoke Tsy yields
T s y = T s h + ( q r + q s ) l n ( R s h o / R s o ) 2 π h s h
where Rso is the stator outer radius in m, and hsh is the housing thermal conductivity in W/(m·°C).
The differential equations of the heat conduction and the boundary conditions for a cylinder with uniform heat generation are as follows [23]:
{ 1 r d d r ( r d t d r ) + q r λ r + q r + q s λ a v g = 0 T = T s y , r = R s o ; d t d r = 0 , r = 0
where λr is the rotor thermal conductivity in W/(m·°C).
The thermal distribution of the machine can be given by
{ h f = λ f b + h 2 b h ( 2 v v f ) 0.8 P r f 0.4 ( η f η w ) 0.14 T ( r ) = R s o 2 4 ( q r λ r + q r + q s λ a v g )
Hence, the stator teeth temperature can be obtained by
T s t = T i + T o 2 + ( q r + q s ) ( l n ( R s h o / R s o ) 2 π h s h + R s o 2 R s i 2 4 λ a v g ) + 1 π h f ( q r R s h o R s h a o + q s R s h o R s i ) + ( R s o 2 R s i 2 ) q r 4 λ r
According to Equations (11), (15) and (28), as the average velocity of fluid increases, the convection coefficient of the cooling jacket increases and the stator temperature decreases. According to the prototype dimensions, when the no-load machine is running at the speed of 1000 r/min, the relationship between the temperature of the equivalent stator core (marked in Figure 10) and the inlet velocity can be obtained as shown in Figure 12. Obviously, as the inlet velocity of water increases to 0.6 m/s, the equivalent stator core temperature decreases almost linearly. When the inlet velocity increases to 1 m/s, the stator core temperature varies nonlinearly and slowly, so 1 m/s is set as the rated cooling inlet velocity of the FSPM machine, corresponding to a pump flow of 1800 L/h.

4. CFD-Based 3D Temperature Field Verification

In order to verify the proposed LPTN and 1D-SHC models, based on the loss calculated by FEM, a 3D-CFD thermal model is built as shown in Figure 13a. Figure 13b corresponds to water cooling.
When the cooling jacket is injected with an total inlet flow of 1800 L/h, a phase current of 30.7 A and 60 A, as well as a speed of 1000 r/min, the transient temperature rises of different components under forced water cooling are obtained as shown in Figure 14. It can be seen that the armature windings achieve the highest temperature under forced water-cooling conditions, whereas the armature windings temperature difference is 31 °C when the armature current is 30.7 A and 60 A, respectively. In addition, compared with the results obtained by the LPTN model shown in Figure 9, both the steady-state and transient results agree well.
On the other hand, to verify the 1D-SHC model, the predicted equivalent stator core temperatures vs. inlet velocity by 1D-SHC and CFD are compared in Figure 15, where the operation status and cooling conditions are consistent with the 1D-SHC model. In addition to predicted temperatures, the time consumed by the three methods is compared in Table 5. Obviously, both the LTPN and 1D-SHC methods can save considerable time, which is favorable for the optimal design of machines.

5. Experiment Verification

To validate the proposed thermal prediction models, a prototyped FSPM machine was manufactured and tested as shown in Figure 16. The prototyped FSPM machine was driven by an inverter supplied by a DC power source and the output shaft was directly connected with a dynamometer machine, in which a torque transducer and a resolver were equipped to measure torque and (rotor position) speed, respectively. Figure 17 compares the FEA-predicted and experimental results of torque versus phase currents. It can be seen that good agreements can be achieved with a deviation below 8%.
To verify the LPTN model, experiments on transient temperature rise were performed. Under the phase current of 30.7 A and rated speed of 1000 r/min, the transient temperature rises of different components under forced water cooling were obtained as shown in Figure 18, where the temperature was detected with a hand-held infrared thermometer. Under forced water-cooling conditions, the measured highest temperature was 47.1 °C. Compared with the results obtained by LPTN (Figure 9) and CFD (Figure 14), it was found that the steady-state and transient results of the three methods were very close. Figure 19 shows the experimental steady-state temperatures under forced water cooling. Obviously, agreement between the experiments and LPTN was achieved, validating the effectiveness of the LPTN model.
To verify the 1D-SHC model, experiments with temperature rising and various fluid inlet velocities of the no-load machine at the rated speed of 1000 r/min were conducted, as shown in Figure 20. It can be seen that when the inlet velocity was bigger than 1.1 m/s, the temperature of the equivalent stator core decreased slowly, which agrees with the simulations giving both 1D-SHC and CFD results.
Overall, satisfactory agreement was achieved between the calculation and measured results, considering manufacturing and testing tolerances.

6. Conclusions

In this paper, a LPTN model is constructed to predict transient thermal behavior of the FSPM machine. Meanwhile, a simplified 1D-SHC model is also proposed to obtain the relationship between the internal temperature of the stator and the coolant flow rate and coolant temperature. The time consumption of the LPTN and 1D-SHC models was significantly less than that of the CFD model, which has advantages in machine design and optimization with large amounts of data. Based on the housing water jacket cooling FSPM machine studied in this manuscript, the LPTN and 1D-SHC methods have accelerated the steady-state temperature calculation speed by 330 and 1100 times, respectively, compared to the CFD method, and have accelerated the transient calculation speed by 857 and 4285 times, respectively. The static and transient temperatures under different conditions were verified by the CFD calculations and experiments. The predicted results from the models agree well with experimental results. This work will be useful in further investigation of thermal analysis of FSPM machines.

Author Contributions

Conceptualization, W.H.; methodology, W.Y. and Z.W.; software, W.Y.; investigation, W.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Fund for Distinguished Young Scholars of China under Grant 51825701 and the Major Program of National Natural Science Foundation of China under Grant 51991381.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to the project being still in progress.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, C.; Chau, K.T.; Lee, C.H.T.; Song, Z. A critical review of advanced electric machines and control strategies for electric vehicles. Proc. IEEE 2021, 109, 1004–1028. [Google Scholar] [CrossRef]
  2. Lee, C.H.T.; Hua, W.; Long, T.; Jiang, C.; Iyer, L.V. A critical review of emerging technologies for electric and hybrid vehicles. IEEE Open, J. Veh. Technol. 2021, 2, 471–485. [Google Scholar] [CrossRef]
  3. Husain, I.; Ozpineci, B.; Islam, M.S.; Gurpinar, E.; Su, G.; Yu, W.; Chowdhury, S.; Xue, L.; Rahman, D.; Sahu, R. Electric Drive Technology Trends, Challenges, and Opportunities for Future Electric Vehicles. Proc. IEEE 2021, 109, 1039–1059. [Google Scholar] [CrossRef]
  4. Wang, P.; Hua, W.; Zhang, G.; Wang, B.; Cheng, M. Principle of flux-switching PM machine by magnetic field modulation theory part II: Electromagnetic torque generation. IEEE Trans. Ind. Electron. 2022, 69, 2437–2446. [Google Scholar] [CrossRef]
  5. Sheng, Z.; Wang, D.; Fu, J.; Hu, J. A computationally efficient spatial online temperature prediction method for PM machines. IEEE Trans. Ind. Electron. 2022, 69, 10904–10914. [Google Scholar] [CrossRef]
  6. Cao, L.; Fan, X.; Li, D.; Kong, W.; Qu, R.; Liu, Z. Improved LPTN-based online temperature prediction of permanent magnet machines by global parameter identification. IEEE Trans. Ind. Electron. 2023, 70, 8830–8841. [Google Scholar] [CrossRef]
  7. Zhang, H.; Giangrande, P.; Sala, G.; Xu, Z.; Hua, W.; Madonna, V.; Gerada, D.; Gerada, C. Thermal model approach to multisector three-phase electrical machines. IEEE Trans. Ind. Electron. 2021, 68, 2919–2930. [Google Scholar] [CrossRef]
  8. Hwang, S.W.; Ryu, J.Y.; Chin, J.W.; Park, S.H.; Kim, D.K.; Lim, M.S. Coupled electromagnetic-thermal analysis for predicting traction motor characteristics according to electric vehicle driving cycle. IEEE Trans. Veh. Technol. 2021, 70, 4262–4272. [Google Scholar] [CrossRef]
  9. Paar, C.; Muetze, A. Influence of dry clutch and ICE transmission integration on the thermal load of a PM based integrated starter-generator. In Proceedings of the IEEE Energy Conversion Congress and Exposition (ECCE), Montreal, QC, Canada, 20–24 September 2015; pp. 106–111. [Google Scholar]
  10. Zhang, G.; Hua, W.; Cheng, M.; Zhang, B.; Guo, X. Coupled magnetic-thermal fields analysis of water cooling flux-switching permanent magnet motors by an axially segmented model. IEEE Trans. Magn. 2017, 53, 8106504. [Google Scholar] [CrossRef]
  11. Du, J.; Liu, Q.; Xue, Y.; Wang, S. Fast Thermal analysis of an ISG in hybrid electric vehicle drive system. In Proceedings of the International Conference on Electrical Machines and Systems (ICEMS), Sydney, NSW, Australia, 11–14 August 2017; pp. 1–6. [Google Scholar]
  12. Nachouane, A.B.; Abdelli, A.; Friedrich, G.; Vivier, S. Numerical Study of Convective Heat Transfer in the End Regions of a Totally Enclosed Permanent Magnet Synchronous Machine. IEEE Trans. Ind. Appl. 2017, 53, 3538–3547. [Google Scholar] [CrossRef]
  13. Assaad, B.; Benkara, K.E.; Vivier, S.; Friedrich, G.; Michon, A. Thermal design optimization of electric machines using a global sensitivity analysis. IEEE Trans. Ind. Appl. 2017, 53, 5365–5372. [Google Scholar] [CrossRef]
  14. Li, R.; Cheng, P.; Lan, H.; Ren, Y.; Hong, Y. Analytic Guided Magnetic-Thermal Kriging Surrogate Model and Multi-Objective Optimization of Synchronous Generator. In Proceedings of the IECON 2022—48th Annual Conference of the IEEE Industrial Electronics Society, Brussels, Belgium, 17–20 October 2022; pp. 1–6. [Google Scholar]
  15. Bertotti, G. General properties of power losses in soft ferromagnetic materials. IEEE Trans. Magn. 1987, 24, 621–630. [Google Scholar] [CrossRef]
  16. Zhu, S.; Cheng, M.; Dong, J.; Du, J. Core loss analysis and calculation of stator permanent-magnet machine considering DC-biased magnetic induction. IEEE Trans. Ind. Electron. 2014, 61, 5203–5212. [Google Scholar] [CrossRef]
  17. Yamazaki, K. Torque and efficiency calculation of an interior permanent magnet motor considering harmonic iron losses of both the stator and rotor. IEEE Trans. Magn. 2003, 39, 1460–1463. [Google Scholar] [CrossRef]
  18. Sadowski, N.; Lajoie-Mazenc, M.; Bastos, J.P.A.; Ferreira da Luz, M.V.; Kuo-Peng, P. Evaluation and analysis of iron losses in electrical machines using the rain-flow method. IEEE Trans. Magn. 2000, 36, 1923–1926. [Google Scholar] [CrossRef]
  19. Simao, C.; Sadowski, N.; Batistela, N.J.; Bastos, J.P.A. Evaluation of hysteresis losses in iron sheets under DC-biased inductions. IEEE Trans. Magn. 2009, 45, 1158–1161. [Google Scholar] [CrossRef]
  20. Yu, W.; Hua, W.; Wang, P.; Tang, C.; Zhang, G.; Cao, R. Coupled electromagnetic-thermal analysis of a 130 kW interior-PM machine for electric vehicles based on field-circuit coupling method. In Proceedings of the International Conference on Electrical Machines and Systems (ICEMS), Harbin, China, 11–14 August 2019; pp. 1–5. [Google Scholar]
  21. Chen, S. The Design of Motors; China Machines Press: Beijing, China, 2000; pp. 94–95. [Google Scholar]
  22. Li, H.; Shen, Y. Thermal analysis of the permanent-magnet spherical motor. IEEE Trans. Energy Convers. 2015, 30, 991–998. [Google Scholar] [CrossRef]
  23. Mellor, P.H.; Roberts, D.; Turner, D.R. Lumped parameter thermal model for electrical machines of TEFC design. IEE Proc. B-Electr. Power Appl. 1991, 138, 205–218. [Google Scholar] [CrossRef]
  24. Sciascera, C.; Giangrande, P.; Papini, L.; Gerada, C.; Galea, M. Analytical thermal model for fast stator winding temperature prediction. IEEE Trans. Ind. Electron. 2017, 64, 6116–6126. [Google Scholar] [CrossRef]
  25. Ahmed, F.; Ghosh, E.; Kar, N.C. Transient thermal analysis of a copper rotor induction motor using a lumped parameter temperature network model. In Proceedings of the IEEE Transportation Electrification Conference and Expo (ITEC), Dearborn, MI, USA, 27–29 June 2016; pp. 1–6. [Google Scholar]
  26. Yunus, A.C. Heat and Mass Transfer: A Practical Approach; McGraw-Hill: New York, NY, USA, 2007; pp. 133–137, 150–159, 410–411, 482–490, 517–521. [Google Scholar]
  27. Chen, M.; Hua, W.; Zhao, G. Thermal performance of a flux-switching permanent magnet machine for an integrated starter generator in hybrid electric vehicles. In Proceedings of the International Conference on Electrical Machines and Systems (ICEMS), Sydney, NSW, Australia, 11–14 August 2017; pp. 1–6. [Google Scholar]
Figure 1. The flux-density loci of key points in the FSPM machine. (a) The key stator and/rotor core points in the FSPM machine. (b) The Bgr/Bgt loci of key stator and rotor core points. (c) The DC-biased components of Bgr/Bgt. (d) The local minor hysteresis loop.
Figure 1. The flux-density loci of key points in the FSPM machine. (a) The key stator and/rotor core points in the FSPM machine. (b) The Bgr/Bgt loci of key stator and rotor core points. (c) The DC-biased components of Bgr/Bgt. (d) The local minor hysteresis loop.
Wevj 14 00130 g001
Figure 2. No-load core loss density distribution of the machine (n = 1000 r/min).
Figure 2. No-load core loss density distribution of the machine (n = 1000 r/min).
Wevj 14 00130 g002
Figure 3. No-load core loss versus speed of the FSPM machine by three methods.
Figure 3. No-load core loss versus speed of the FSPM machine by three methods.
Wevj 14 00130 g003
Figure 4. No-load 3D eddy current density distribution @1000 r/min. (a) PMs; (b) housing.
Figure 4. No-load 3D eddy current density distribution @1000 r/min. (a) PMs; (b) housing.
Wevj 14 00130 g004
Figure 5. No-load eddy current losses in PMs and housing versus rotating speeds.
Figure 5. No-load eddy current losses in PMs and housing versus rotating speeds.
Wevj 14 00130 g005
Figure 6. Total loss versus speed by different prediction methods and experiments.
Figure 6. Total loss versus speed by different prediction methods and experiments.
Wevj 14 00130 g006
Figure 7. The cooling system of the FSPM machine. (a) FSPM machine structure, (b) housing, (c) coolant flow path, and (d) machine assembly. (e) Cross-section of the cooling jacket, and (f) schematic diagram of cooling duct.
Figure 7. The cooling system of the FSPM machine. (a) FSPM machine structure, (b) housing, (c) coolant flow path, and (d) machine assembly. (e) Cross-section of the cooling jacket, and (f) schematic diagram of cooling duct.
Wevj 14 00130 g007
Figure 8. The LPTN model of the FSPM machine: (a) 3D module structure; (b) stator module; (c) the LPTN model of the 1/24 FSPM machine.
Figure 8. The LPTN model of the FSPM machine: (a) 3D module structure; (b) stator module; (c) the LPTN model of the 1/24 FSPM machine.
Wevj 14 00130 g008
Figure 9. The predicted temperature rises by the LPTN model under water-cooling conditions @ nN = 1000 r/min and Iph = 30.7 A and 60 A (RMS).
Figure 9. The predicted temperature rises by the LPTN model under water-cooling conditions @ nN = 1000 r/min and Iph = 30.7 A and 60 A (RMS).
Wevj 14 00130 g009
Figure 10. One-dimensional steady heat conduction model of the FSPM machine. (a) Equivalent heat conduction model; (b) equivalent heat flow path.
Figure 10. One-dimensional steady heat conduction model of the FSPM machine. (a) Equivalent heat conduction model; (b) equivalent heat flow path.
Wevj 14 00130 g010
Figure 11. Equivalent diagram of the winding structure.
Figure 11. Equivalent diagram of the winding structure.
Wevj 14 00130 g011
Figure 12. The temperature of equivalent stator core vs. inlet velocity by 1D-SHC.
Figure 12. The temperature of equivalent stator core vs. inlet velocity by 1D-SHC.
Wevj 14 00130 g012
Figure 13. Three-dimensional CFD thermal model of the FSPM machine and steady-state temperatures @30.7 A and 60 A; (a) 3D-CFD thermal model; (b) water cooling.
Figure 13. Three-dimensional CFD thermal model of the FSPM machine and steady-state temperatures @30.7 A and 60 A; (a) 3D-CFD thermal model; (b) water cooling.
Wevj 14 00130 g013
Figure 14. The 3D-CFD predicted temperature rises of different components of the FSPM machine.
Figure 14. The 3D-CFD predicted temperature rises of different components of the FSPM machine.
Wevj 14 00130 g014
Figure 15. The predicted equivalent stator core temperatures vs. inlet velocity by 1D-SHC and CFD.
Figure 15. The predicted equivalent stator core temperatures vs. inlet velocity by 1D-SHC and CFD.
Wevj 14 00130 g015
Figure 16. Prototype photographs and test platform.
Figure 16. Prototype photographs and test platform.
Wevj 14 00130 g016
Figure 17. FEA-predicted and experimental torque versus phase currents.
Figure 17. FEA-predicted and experimental torque versus phase currents.
Wevj 14 00130 g017
Figure 18. The temperature rises of the FSPM machine under forced water-cooling conditions.
Figure 18. The temperature rises of the FSPM machine under forced water-cooling conditions.
Wevj 14 00130 g018
Figure 19. Steady-state temperature distribution of the FSPM machine under forced water-cooling conditions.
Figure 19. Steady-state temperature distribution of the FSPM machine under forced water-cooling conditions.
Wevj 14 00130 g019
Figure 20. The equivalent stator core temperatures in the 1D-SHC model, CFD model, and experiment.
Figure 20. The equivalent stator core temperatures in the 1D-SHC model, CFD model, and experiment.
Wevj 14 00130 g020
Table 1. Design specifications of the FSPM machine.
Table 1. Design specifications of the FSPM machine.
ParameterSymbolValueUnit
DC-link voltageUDC144V
Phase numberm3-
Stator slotsNs12-
Rotor pole pairsNr10-
PM pole pairsNPM6-
Rated powerPN10kW
Rated speednN1000r/min
Rated torqueTN95.5Nm
Stator outer diameterDso260mm
Rotor inner diameterDri50mm
Air-gap lengthg00.9mm
Stack lengthLa55mm
Table 2. Thermal conductivities of materials.
Table 2. Thermal conductivities of materials.
MaterialsThermal Conductivity (W/m/°C)Specific Heat Capacity (J/kg/°C)Density (kg/m3)
Steel silicon234607650
Copper3803858978
PM95047500
Aluminum2378332688
Air0.0262410051.205
Table 3. Component thermal resistances of the LPTN model.
Table 3. Component thermal resistances of the LPTN model.
Thermal ResistancesValue (°C/W)Thermal ResistancesValue (°C/W)
Rsh1, Rsh2, Rsh30.0004307Rst1, Rst20.1479
Rsh40.1094Rst30.01449
Rsh50.07664Rcoil10.6297
Rsh6, Rsh7, Rsh80.0004469Rcoil24.895
Rsy1, Rsy20.00359Rcoil30.7244
Rsy30.2107Rpm10.01197
Rsy40.1128Rpm20.04699
Rsy50.003729Rpm30.05441
Rsy60.01632Rpm40.2054
Rair1, Rair2, Rair31.096Rpm50.04829
Rair4, Rair5596Rrt10.02435
Rair6, Rair727.84Rry10.03246
Rair830.76Rshaft10.2613
Rair956.14--
Table 4. Thermal capacitances of the LPTN model.
Table 4. Thermal capacitances of the LPTN model.
Thermal CapacitanceValue (J/°C)Thermal CapacitanceValue (J/°C)
Csh130.82Cst171.37
Csh221.86Ccoil10.0036
Csh310.75Cpm110.67
Csy131.63Cpm233.14
Csy222.98Cr497.2
Table 5. Comparison of time consumed by temperature prediction methods.
Table 5. Comparison of time consumed by temperature prediction methods.
Temperature Prediction MethodsSteady StateTransient
LPTN method2 s7 s
1D-SHC method0.6 s1.4 s
CFD method11 min100 min
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

Yu, W.; Wu, Z.; Hua, W. Thermal Analysis of a Flux-Switching Permanent Magnet Machine for Hybrid Electric Vehicles. World Electr. Veh. J. 2023, 14, 130. https://doi.org/10.3390/wevj14050130

AMA Style

Yu W, Wu Z, Hua W. Thermal Analysis of a Flux-Switching Permanent Magnet Machine for Hybrid Electric Vehicles. World Electric Vehicle Journal. 2023; 14(5):130. https://doi.org/10.3390/wevj14050130

Chicago/Turabian Style

Yu, Wenfei, Zhongze Wu, and Wei Hua. 2023. "Thermal Analysis of a Flux-Switching Permanent Magnet Machine for Hybrid Electric Vehicles" World Electric Vehicle Journal 14, no. 5: 130. https://doi.org/10.3390/wevj14050130

Article Metrics

Back to TopTop