Next Article in Journal
Intelligent Terrain Mapping with a Quadruped Spider Robot: A Bluetooth-Enabled Mobile Platform for Environmental Reconnaissance
Previous Article in Journal
MPC-Based Sensor Fault-Tolerant Control: Application to a Heat Exchanger with Measurement Noise
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modern Control Techniques and Operational Challenges in Permanent Magnet Synchronous Motors: A Comprehensive Review

by
Mahmoud M. Elkholy
1,2,*,
Mohamed M. Algendy
2 and
Enas A. El-Hay
2
1
Electrical Engineering and Computer Science Department, College of Engineering, A’Sharqiyah University (ASU), Ibra 400, Oman
2
Electrical Power and Machines Engineering Department, Faculty of Engineering, Zagazig University, Zagazig 44519, Egypt
*
Author to whom correspondence should be addressed.
Automation 2025, 6(4), 49; https://doi.org/10.3390/automation6040049
Submission received: 15 July 2025 / Revised: 6 September 2025 / Accepted: 18 September 2025 / Published: 23 September 2025

Abstract

This paper presents a comprehensive overview of permanent magnet synchronous motors (PMSMs), including their classifications, applications, and vector control strategies. It explores various control techniques, including maximum torque per ampere (MTPA), maximum current (MC), field weakening (FW), maximum torque per voltage (MTPV), sensorless control, and parameter identification, as discussed in this paper. These methods address key challenges in PMSM control, such as improving motor efficiency and accurately estimating rotor position and speed. Additionally, this paper presents the PMSM parameters due to many factors such as electric current, phase angle, saturation, and temperature. The survey findings provide a deeper understanding of PMSMs’ control strategies, aiding in the more efficient and reliable motor studies.

1. Introduction

Permanent magnet synchronous motors (PMSMs) are commonly utilized in industrial applications like servo drives, high-speed trains, electric vehicles (EVs), actuators, hard disk drives, rolling mills, and household appliances due to their high efficiency, power density, and broad speed range. Using of permanent magnets (PMs) in PMSMs has many advantages including, avoiding the required energy for field excitation, improved dynamic performance and ease of construction and maintenance [1]. In particular, as it is well known that the use of EVs nowadays is very necessary to replace traditional internal combustion engine vehicles (ICEVs), PMSMs can help cut down on greenhouse gas emissions, lessen dependency on fossil fuels, and benefit the electric grid and consumers. The greenhouse emissions produced for the required electricity for EVs are 50% [1] lower than that produced from traditional ICEVs, and this amount of reduction will increase by increasing the production of electricity from renewable energy resources. Furthermore, many countries such as Germany, Britain, France, and China have plans for gradually getting rid of the manufacturing of cars powered by gasoline and diesel in a few decades. By 2040, it is anticipated that 54% of new automobile sales would be electric. For instance, Ford has invested USD 11 billion on 40 new electric vehicles over the coming years [1]. For the foregoing reasons and notes and the wide applications for PMSMs, the control of PMSMs is required to be robust, accurate, fast dynamic, and efficient. The control of PMSMs faces some problems, resulting in reducing the performance of the controllers. The primary issue with PMSMs is the variation of parameters like resistance and d-q axis inductance. Specifically, the d-q-axis inductances fluctuate based on the motor’s shape, mechanical power, and operating characteristics; hence, inaccurate parameter estimate results in high temperatures and deteriorated control system performance. High temperatures can lead to an increase in stator resistance, while simultaneously reducing both the rotor flux linkage and the d-q axis inductances, potentially resulting in torque ripples [2,3]. Another major challenge in PMSM control is the estimating of the rotor speed and position using traditional methods that rely on built-in sensors, like optical and rotary encoders. These conventional sensors take up space, making motor installation difficult, and may cause the motor to malfunction due to sensor failures in harsh operating environments [4,5]. Therefor this paper addresses the main PMSM control topics in the literature to achieve its merits and reduce its problems. The remainder of the paper is organized as follows: Section 2 presents the types of PMSMs, comparing their different classifications and applications. Section 3 discusses the mathematical model of PMSMs and explains the four operating regions: maximum torque per ampere (MTPA), maximum current (MC), field weakening (FW), and maximum torque per voltage (MTPV). Section 4 covers MTPA control schemes aimed at enhancing motor efficiency and performance and explores artificial-intelligence-assisted methods for MTPA control. Section 5 presents sensorless control techniques, which eliminate traditional built-in sensors to improve system reliability and reduce costs. Additionally, Section 6 discusses parameter identification control methods to address nonlinear parameter variations in PMSM control systems, including those related to MTPA and sensorless approaches. Finally, Section 7 provides the findings and conclusions.

2. PMSM Types and Applications with Recent Technologies in Constructions

2.1. PMSM Types and Applications

Synchronous motors can be classified mainly into three types, namely, wound rotor synchronous machines (WRSMs), PMSMs, and reluctance synchronous motors (RSMs). WRSMs are appropriate for situations requiring high power. PMSMs and RSMs are favored in the low- to medium-power range, where induction motors are gradually being replaced [1,6]. PMSMs are better than RSMs and IMs in their power density and efficiency, but their cost is more than them, as shown in Figure 1. PMSMs are the best choice for multiple industrial uses because of their many benefits. In EV application, PMSMs are the best choice for the manufacturers of EVs based on a comparison between IMs, SRMs, and PMSMs. According to [1], in air conditioning compressors, ref. [6] presents a comparison between PMSMs and SYNRMs. In pump load applications, ref. [7] shows the different Ch/s between the PMSM and the other motors such as induction motors, squirrel cage induction motors, synchronous reluctance motors, and switched reluctance motors. In the aerospace industry, PMSMs are also a better option, particularly for the newest electric starter/generator system that operates in a high-speed range for future aircraft [8].
Permanent magnet synchronous machines (PMSMs) are widely used in wind electric generation systems because they are efficient, reliable, and can work well at variable speeds. In wind turbines, PMSMs help capture the most energy from changing wind speeds by using advanced control methods like model predictive control and maximum power point tracking, which improve power output and system efficiency [9,10,11]. PMSMs are also valued for their ability to keep working even if a fault occurs, especially in open-winding designs, which allow the wind turbine to continue operating—though at lower power—until repairs can be made [12,13]. New design and control strategies, such as fuzzy logic, sliding mode control, and optimization algorithms, further improve the stability, power quality, and fault tolerance of PMSM-based wind systems [14,15,16]. These features make PMSMs a strong choice for both large wind farms and small wind turbines, helping to ensure reliable and efficient renewable energy generation.
PMSMs could be classified into two types, namely, interior permanent magnets (IPMSM)s and surface permanent magnets (SPMSMs). Both IPMSMs and SPMSMs could take different designs according to the position of permanent magnets in the rotor, as shown in Figure 2. Table 1 presents a comparison between IPMSMs and SPMSMs [1,2,6,7,17,18,19,20,21,22,23,24,25,26,27,28].

2.2. PMSM Recent Technology Construction Design

Recent studies, however, show many new trends. One important direction is the use of hybrid permanent magnets and new rotor shapes to increase efficiency and improve resistance to demagnetization [29]. Axial-flux PMSMs with coreless designs are also receiving attention because they provide very high torque and power density, although they bring challenges in cooling and mechanical stability [30]. New winding approaches, such as hairpin windings combined with direct oil-cooling, improve thermal performance and are increasingly used in high-power traction motors [31]. Other research highlights the impact of rotor losses in rare-earth PMSMs, especially during field weakening at high speeds, which is important for aerospace and EV drives [32]. Cost modeling studies show that winding type, cooling, and magnet material all have strong influence on manufacturing cost, which is critical for large-scale production of electric vehicle motors [33]. In addition, surface treatments and protective coatings for permanent magnets are being developed to reduce the risk of demagnetization and corrosion in harsh operating environments [34]. Reliability is another key trend. Nine-phase and five-phase PMSMs with advanced fault-tolerant control strategies have been proposed to ensure continuous operation during open-circuit faults [35,36]. Dual-winding PMSM structures are also under investigation to enable smooth operation under faults and reduce harmonic effects [37]. These developments confirm that PMSM construction is still an active research field, with innovations in materials, winding and cooling methods, cost reduction, and fault tolerance.
Multiphase permanent magnet synchronous motors (PMSMs), such as five-phase and six-phase machines, are increasingly vital in modern applications due to their enhanced reliability and operational flexibility. Recent research has focused on advanced torque optimization and fault-tolerant control (FTC) strategies to fully exploit the additional degrees of freedom these motors offer. Optimization-based FTC schemes, including those utilizing particle swarm optimization and interior point methods, have demonstrated the ability to maximize torque output while minimizing copper losses, even under open-phase or open-switch fault conditions [38,39,40]. These approaches often outperform conventional methods by achieving higher torque capacity—up to 73.3% of the healthy state and reducing losses by up to 20% during faults [38,39]. Furthermore, innovative control strategies, such as model predictive control and current reference optimization, enable smooth transitions between healthy and faulty operations, suppress torque ripple, and maintain system stability without complex hardware modifications [41,42,43]. The integration of radial force compensation and consideration of phase current limits further enhances the robustness and efficiency of multiphase PMSMs, making them highly suitable for critical and demanding applications [44,45,46]. Discussing these advancements in torque optimization and FTC will significantly strengthen the technical depth and practical relevance of any paper on multiphase PMSM drives.

3. PMSM Operation and Control

The PMSM mathematical model and its four operational zones are shown in this section, named as MTPA, MC, FW, and MTPV.

3.1. Mathematical Model of PMSMs

In the rotor reference frame, the PMSM’s voltage equations may be expressed as follows [25]:
v d = R s i d + d ψ d d t ω e ψ q
v q = R s i q + d ψ q d t + ω e ψ d
ψ d = L d i d + λ f
ψ q = L q i q
And for the steady-state operation, the equations will be
v d = R s i d ω r L q i q
v q = R s i q + ω e L d i d + λ f
The electromagnetic torque is expressed as
T e = 0.75 p λ i q s + 0.75 p L d L q i d i q
where v d and v q are the d - q axis stator voltages; i d and i q are the d - q axis current. The d - q axis inductance is denoted by L d and L q . The stator resistance, the permanent-magnet flux linkage, and the pole pairs are denoted by R s , λ f , and p , respectively. The electrical angular frequency is denoted by ω e .

3.2. PMSM Operation Regions

Motors for PMSM applications, particularly mild-hybrid and fully electric cars, need to have a large speed range at constant power and good torque at low speeds. Consequently, based on the speed ranges reported in the literature, the operating zones may be divided into four sections [47,48,49,50,51,52,53]. These four regions are maximum torque per ampere (MTPA), maximum current (MC), flux weakening (FW), and maximum torque per voltage (MTPV). The current plane characteristics of IPMSMs and SPMSMs are shown Figure 3. Figure 3 illustrates how the saliency of IPMSMs causes the voltage limit to take the form of an ellipse in IPMSMs and a circle in SPMSMs. Therefore, the MTPA and MTPV trajectories in IPMSMs are hyperbolic curves, as opposed to SPMSMs, where they are straight lines. For the four operation regions, if the characteristics curves of IPMSM are used in representation, consider that the red star point serves as the operating point, as shown in Figure 4. Firstly, the motor in MTPA runs at the base speed n b on the OM curve, which is analytically constructed in (8) and displayed in Figure 4a. Point (O) represents the origin point in the current plane, and point (M) represents the intersection of the MTPA trajectory with the current limit circle. The voltage limit ellipse crossing the MTPA trajectory at the current limit circle defines n b . The point where the MTPA and load torque curves connect will serve as the operational point. It must be noted that in the MTPA region, the optimal values of i d and i q depend on the load torque, regardless of the required speed. Secondly, in the MC region, the speed is more than the base speed and under the boundary speed n b o is defined by the voltage constraint ellipse passing through the origin and intersects with the MC circle at F, as displayed in Figure 4b, if it is assumed that the required operating speed is n r e q . n r e q is represented by the voltage limit intersecting with the current limit circle and OM curve at R and D, respectively. For operating in the MC region at n r e q , an additional negative d-axis current is wanted for moving from M to R. The operating point is point R at the load torque T 3 . If the load torque is smaller than T 3 as T 2 , the motor will move from R on the voltage limit to operating point X. If the torque is lighter as T 1 , the motor will continue moving on the voltage limit to intersect with the MTPA curve at R. The motor cannot go across the MTPA curve to intersect with T 3 at Z but it will move along MTPA curve to intersect with T 3 at N.
i d = λ f 2 L q L d λ f 2 4 L q L d 2 + i q 2
Thirdly, the speed in the flux weakening region is greater than n b o and lower than n v . n v is the deep flux-weakening speed shown in Figure 4c by the voltage constraint ellipse going through the intersection of the MTPV locus and current circle at point V . If the required speed n r e q is higher than n b o , the motor will operate at point Y if the required torque is T 2 , as shown in Figure 4c. For a smaller torque T 1 , the motor will operate within the voltage limit of n r e q to operate at point k , as shown in the blue path in the Figure 4c.
Fourthly, for a speed more than n v the motor operates in the deep-flux-weakening region MTPV. The trajectory of MTPV can be derived analytically, as in [53]. So, for n r e q in Figure 4d, it is defined by the voltage limit that intersects with the MC limit at point Q . The motor will operate at the MTPV trajectory if the required torque intersects with the n r e q voltage limit at the MTPV trajectory at point Q for torque T 4 . If the torque is smaller than T 4 as T 1 , the motor will move along the n r e q voltage limit to operate at the intersection point P .

3.3. General Control Methods for PMSMs in Solving PMSM Issues

3.3.1. Vector Control

There are three primary topologies for the vector control approach, namely, field-oriented control (FOC), direct torque control techniques (DTCs), and voltage vector control (VVC) [54,55]. In FOC, the current i d s = 0 for max torque at low speeds and for increasing speed i d s will be negative. FOC can be used with or without a position sensor, but when using a sensor, the FOC has small tolerance and is expensive and not reliable [56,57,58,59]. DTC circuits are simpler in structure than the FOC circuit. Since precise rotor position information is not needed in DTC systems (apart from the synchronous motor’s startup scenario), the inverter with PWMs is not necessary. All computations are carried out in the stator reference frame. In DTCs, the controllers’ computing capabilities are comparatively minimally required. DTCs respond rapidly to changes in load, offer excellent dynamic properties, and are less susceptible to disturbances and variations in motor parameters. However, high stator current, flux linkage, and torque ripple levels, particularly at low speeds, are characteristics of steady-state operation that significantly restrict their usage for high-precision drives in automotive applications [56,60,61,62,63,64]. The VVC method is relatively simple, and the information regarding the rotor’s position is not required. One of the advantages of VVC is that it has low sensitivity to motor parameter changes; however, it includes higher ripple than in the case of using FOC and high torque overshoot amplitude, which makes using the VVC in high-precision driving systems challenging [54,65,66,67].
Recently, in non-linear control theory, passivity-based control (PBC) has drawn a lot of interest because of its dependability, effectiveness, and rather straightforward construction. The EL formalism yields the PMSM model in terms of energy quantities. Additionally, it can be observed that the PMSM defines specific passive maps, after which its dynamics are handled as two distinct EL passive subsystems: mechanical and electrical. The electrical subsystem’s mechanical dynamics are regarded as passive disruptions. In order to strengthen the passivity quality, dumping is also introduced into the electrical subsystem. The internal closed loop energy is reshaped to provide current tracking. In the PBC method, the closed loop system is subjected to a passivity property in order to achieve the control objective. This is accomplished by allocating the desired dissipation and storage functions to the closed loop dynamics [60,68]. This method eliminates the rotor location from the coefficients of the differential equations and reduces the number of variables. This simplification produces stability analysis, controller synthesis, and the determination of steady state conditions—all of which are crucial for evaluating the machine’s performance [60,68].

3.3.2. Scalar Control

Open loop constant V/f inverters are frequently utilized in industrial variable speed applications, especially in induction motor (IM) drives due to their advantages of simple implementation, low cost, and general purpose. However, the drawbacks include limited dynamic operation, low energy efficiency, and a failure to account for the motor’s coupling effects. For a PMSM drive, the concept of the open loop constant V/f-controlled method is usually not directly applied because the open loop system is generally unstable. By adding a frequency damper to the control to increase the system damping constant, the stability issue can be prevented. By including two feedforward channels for the slip and stator voltage magnitude in the control scheme, the scalar controller’s dynamic response can be improved. Its ability to boost performance is dependent on how accurate these feedforward routes are. This is why applications with fast dynamic requirements, such as traction applications, do not use the v/f approach. Nonetheless, certain multilayer traction applications with slower dynamics, such as electric ships, use it [1,69,70,71,72,73,74].

3.3.3. Flux-Weakening Control Methods

(a)
Feed forward method
In feed forward techniques, control variables that have been studied analytically or empirically are calculated to maximize motor performance within the presumptive feeding voltage and current limits. Complex nonlinear equations must be solved online for the intersections (solutions) involving the torque hyperbola, voltage ellipse, and current circle. The accuracy of the current set-points declines because of the frequent ignoring of magnetic saturation and resistive voltage loss to speed up the solving procedure. Although feed-forward techniques are known for their stability and transient responses, they have several disadvantages, including temperature volatility, material qualities that change over time, and saturation [75,76,77,78,79,80,81,82].
(b)
Feedback technique
The feedback systems track the voltage limit as the speed increases to adjust the negative d-axis current. Once the voltage or speed has been measured, the demagnetizing current vector can be modified by monitoring the voltage or speed error. These methods are robust since they are not affected by parameters variations and have higher efficiency than previous methods. However, inadequate performance during transients is the price paid for this [83,84,85,86,87,88,89,90].
(c)
Hybrid technique
In order to benefit from the advantages of both feed-forward and feedback methods, hybrid strategies have been created. While the torque command and d-axis current feedback are used to construct the q-axis current command, the optimization objectives are used to modify the pre-computed d-axis current command for MTPA. The hybrid control techniques are highly resilient to model uncertainty and parameter variations because they typically use look-up tables to determine the excitation current commands for MTPA while accounting for magnetic saturation and varying parameters. The feed-back technique is then used to determine the current commands for flux-weakening operation. However, because of the lengthy experiments needed to obtain the look-up tables, the transient response could be delayed [82,91,92].

3.3.4. Control and Converter Considerations

In permanent magnet synchronous machines (PMSMs), the power converter plays a significant role in motor control. The supply’s harmonic distortion, switching frequency, and available voltage vectors are all defined by the converter topology [93,94]. These elements have a direct effect on the greatest control bandwidth that may be attained, torque ripple, and current ripple [93,95]. Different control possibilities are made possible by different converter structures. Two-level inverters are simple and reliable but give limited voltage vectors. Multilevel inverters, such as NPC or T-type, provide more voltage levels, reduce harmonics, and improve torque control [94,96]. However, these converters need complex balancing techniques, careful thermal design, and more sophisticated control [96,97]. The converter and control algorithms need to be compatible. The additional voltage vectors of multilayer inverters can be used by model predictive control and direct torque control to reduce torque ripple and produce faster dynamic response [95,98]. They rely on precise measurements and raise computational needs in exchange [98,99]. Imperfections in the converter also come into play. Errors in voltage generation are caused by dead-time, switching delay, and finite DC-link voltage. If these effects are not compensated in the control, performance degrades under transient or high-speed operation [97]. For this reason, PMSM and converter design should be considered together. Co-design methods that optimize the motor, the inverter, and the control algorithm at the same time lead to higher efficiency, greater power density, and better fault tolerance [99]. This strategy is particularly crucial for high-speed industrial machinery, renewable energy systems, and electric vehicle drives [99].

3.3.5. Single-Loop Control Structure for PMSM Speed and Current Regulation

The single-loop control structure, which integrates the speed and current loops into a merged control loop, has lately concerned significant consideration in PMSM drive research. Compared to old cascade control, this approach enables more direct control action, greater control bandwidth, and improved speed dynamic performance. Recent studies have demonstrated that single-loop controllers—often based on advanced sliding mode control, disturbance observers, or hybrid reaching laws—can achieve fast transient response, strong robustness against disturbances, and effective overcurrent protection, all while simplifying the control system. For example, single-loop strategies using current-constrained sliding mode control and time-varying nonlinear disturbance observers have shown improved speed and current tracking, reduced overshoot, and enhanced disturbance rejection compared to conventional cascade structures [100,101,102,103,104]. These advancements confirm the promise of single-loop control paradigms for PMSM drives and highlight the need for further in-depth investigation into their practical applications and optimization [101,105,106,107,108].

4. Maximum Torque per Ampere Control Methods

The MTPA techniques can be categorized as shown in Figure 5.

4.1. Machine-Parameter-Dependent MTPA Techniques

The values of machine parameters, such as PM flux linkages and inductances, are required from offline testing, an approximated fitting function, or an estimating strategy during machine operation in machine parameter dependent MTPA procedures.

4.1.1. Premeasurements Assisted Methods

The premeasurements assisted MTPA methods that depend on look up table (LuT). In particular, the pre-testing data is used to adjust for changes in machine parameters over the speed and torque ranges. The drive system, where measurements were gathered through numerous offline tests, has made extensive use of lookup tables (LuT). LuTs have been used extensively in drive systems where measurements were gathered through numerous offline tests. According to the torque reference, some techniques choose the reference flux linkage that corresponds to the MTPA profile directly [109,110,111,112,113]. In other methods, LuT provides the i d and i q references, which show the optimal currents as a proportional to torque [91,114,115,116]. Based on load circumstances, the armature current in [117] is calculated on a LuT. If the inverter nonlinearities are taken into account, the speed range of LuT is increased [116]. Apart from the straightforward approaches above, Ref. [118] uses the Gauss–Newton algorithm to generate reference currents from the online calculation block. The prefabricated LuT is used to determine the associated machine parameters. These approaches have the disadvantages of requiring a lot of computer storage space, requiring an accurate interpolation technique, and requiring a lot of testing time.

4.1.2. Approaches Based on Parameter Estimates and Approximated Functions

These techniques use a polynomial curve/surface fitting to trend how machine parameters vary under various operating situations. Although there are fewer pre-tests, the basic idea behind these approaches is the same as that described in the previous section. To avoid using the LuT, for example, a two-dimensional curve fitting function is developed in [119,120,121,122] to provide the best fit to a set of measurements. For considering the problem of magnetic saturation and cross-coupling drawbacks, surface fitting techniques are presented in [123,124] to estimate the flux linkage profiles. The more intricate in modelling process, the more pre-test measurements are required. To balance the precision of MTPA operation with a fair testing duration, ref. [125] suggests using just two pretest points on the approximated MTPA curve to calculate the parameters of a simplified MTPA model on the flux linkage synchronous frame. Many of these methods employ iterative algorithms like gradient descent with Lagrange multipliers and Gauss–Newton. In [53], the Newton–Raphson algorithm is used to improve the control precision of the current set-points in MTPA, MC, flux weakening, and MTPV regions. In the meantime, the effects of several parameters, such as resistive voltage drop and magnetic saturation, are thoroughly examined. As was previously established, the real current trajectory deviates from the ideal one when the resistive voltage drop and magnetic saturation are ignored. Compared to the LuT-based approach, this method has the benefit of requiring less pre-testing time.

4.1.3. Parameter-Adaption-Based Methods

These methods depend on online parameter identification strategies such as recursive least squares, which is a popular technique for identifying parameters [126,127,128,129]. Recursive least squares are used to track the variation of dq–axis inductances and permanent magnet flux coupling, which are involved in the computation of MTPA points, in order to match the estimations to the actual one. The DTC scheme could be used to replace FOC in that it eliminates the q-axis inductance effect and computes the stator flux linkage that satisfies the MTPA requirement directly [130]. Also, the parameters could be estimated by the knowledge of harmonic components. So, the series of filters could be implemented to identify machine parameters. This method is comparable to the method based on virtual signal injection that will be introduced in the parameter identification section [131,132].
This approach seeks to address the issues of low parameter adaptivity, the need for a lot of data storage, and nonlinearity; however, it increases the operational burden and real-time processing cost, which causes a slow reaction.

4.2. Machine-Parameter-Independent MTPA Techniques

The MTPA trajectory can be tracked using the parameter-independent method without requiring knowledge of the machine parameters. Extremum-searching-based, power-measurements-based, and artificial-intelligence-assisted techniques are the three categories into which the parameter-independent approaches are separated, as illustrated in Figure 5.

4.2.1. Extremum Searching Based

There are two kinds of machine-parameter-insensitive signal-injection-based schemes: real signal injection and virtual signal injection. The extreme searching technique relies on the extraction of derivatives to determine the MTPA point and does not require prior data [109,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152]. It can be done by injecting disturbance signals such as high-frequency (HF) current [136,153] and HF current angle [137,138,139,140,141,142,143] into the system. The magnitude of injected perturbation is relatively small [38] so that the operational conditions are not significantly affected. The resultant torque, which includes the high-frequency component, has often been expanded using the Taylor series. Since the techniques described in this section necessitate additional time-consuming components like integrators and filters, one of their main disadvantages is that they limit dynamic performance.

4.2.2. Power-Measurement-Based Methods

The MTPA point is equal to the maximum efficiency point in the region of constant torque, which is necessary for power-measurement-based methods to work. Therefore, the additional dc-link current and voltage sensors can help this method. This approach uses the gradient descent algorithm to find the maximum efficiency point after calculating mechanical power and requiring dc power measurements. An offline map of the phase advance angle at various speed conditions is created by varying the dc-link power. To find MTPA spots while the machine is operating, the gradient descent approach is used. Due to the absence of three-phase current measurements, power observation is advantageous for certain alternative control systems, such as sensorless and V/f.

4.2.3. Artificial-Intelligence-Assisted Methods

The description of the nonlinear relationship between torque and currents heavily relies on artificial intelligence. Artificial intelligence techniques, such as fuzzy, neural network, and optimization algorithms, are employed and trained to provide the best current directives during MTPA operation. For instance, to determine the ideal values of i d and i q for reaching MTPA, a genetic algorithm is utilized [154]. The foundation of this technique is the discovery of an unknown function that delivers the maximum torque per current as a relation between the currents. A controller-based genetic algorithm was used to find the i_d, and a PI controller was used to calculate the i_q. Fast dynamic reaction and cheap real-time computing cost are two advantages of artificial-intelligence-assisted approaches, which are accurate, simple, and quick to configure in real time. They could require more pretests and training procedures, though.

5. Sensorless Control Methods

In order to circumvent the limitations of employing sensors like encoders and resolvers to estimate the rotor position and speed, sensorless control schemes are recommended in PMSM drives. The system becomes less reliable and more expensive when sensors are used. The literature provides a variety of methods and algorithms for figuring out the position and speed for sensorless control. Some techniques rely on feedback to lower the estimation errors, while others are direct and do not require a feedback loop; however, this leads to high error numbers. Thus, open loop, close loop, and signal-injection-based sensorless control techniques can be divided into three categories.

5.1. Open Loop Methods

After converting the three-phase variable quantities into a two-phase (d-q) reference frame, the rotor angle and speed values may be calculated immediately. The primary benefits are ease of use and quick dynamic reaction. However, the problem of parameter fluctuation has a significant impact on this ethos [155,156,157,158]. Additionally, as the inductance is dependent on the rotor position, the rotor angle and speed might be calculated using the stator inductance. This technique works well with high-saliency motors, such as switching reluctance motors and interior motors. It works well at low speeds, even at zero speed, where the EMF is zero. However, because it is an open loop method, the accuracy of the estimation cannot be guaranteed [159,160,161,162,163,164]. The reverse EMF is necessary for another open loop sensorless. The back EMF can effectively assess the rotor speed and angle because it is widely known that the back EMF generated in a motor is proportional to the rotor’s speed or angle. Determining back EMF without relying on voltage probes could lower the method’s cost and boost its reliability. This method is straightforward, so it does not need a complex observer, which makes it simple, accurate at higher frequencies, cost effective, and robust, but it is not accurate at low speed, especially at starting and the difficulty in calculating the back EMF [156,158,165,166,167,168,169].

5.2. Closed Loop Methods

This is an adaptive approach that adjusts to various circumstances and acts accordingly. Comparing the results of systems that are mathematically modeled with those that are real is how this tactic is carried out. Reducing the error between the two systems is the primary goal of these techniques. These adaptive sensorless controllers might make use of an extended Kalman filter. It is important to note that extended Kalman filters (EKF) are utilized for nonlinear systems, while Kalman filters are utilized for linear systems. These filters determine the optimal state of a nonlinear system’s dynamics via least square variance estimation. The solution of the Riccati differential equation, which calls for two covariance matrices, is the foundation of this approach. Measurement noises is the name given to one matrix, and parameter uncertainties to the other. Among the advantages of this approach are its ability to reduce computation time and its insensitivity to measurement noise and parameter error. However, the primary drawback is the subpar performance at slower rates (<5 Hz) [170,171,172,173,174,175].
Sensorless control uses the model-reference-based adaptive system (MRAS), which compares a reference model value with an adjustable model value to minimize the error to zero. It is based on Popov’s stability criterion, which is a very desirable scheme used in speed and rotor resistance estimation, and its greatest improvement is its high speed of adaptation [176,177,178,179,180,181]. Another adaptive sensorless method is the SMO [182,183]. It gets its name from its underlying theory or strategy, which is when certain functions or control laws are chosen so that they go to a shared boundary and then slide on that boundary or surface known as a sliding surface. The SMO is based on defining a sliding surface firstly, and secondly defining control laws or functions to have convergence of the systems towards the predefined sliding surface as these functions make the system slide on that surface as shown. A sliding hyperplane is chosen in SMOs based on the error vector ∆I_s. The system then slides across this hyper-plane until it reaches the smallest error. This method’s robustness ensures that there is no inaccuracy in the rotor value estimations. An SMO can solve the significant issue of parameter fluctuations that other schemes face because it is simple to implement and performs well with nonlinear systems due to its adaptability. However, because the back EMF value is so low at low speeds, it performs poorly at standstill or low speeds. The method’s conflict between a quick rate of convergence and a good steady state result is another issue. The gain coefficient should be selected to guarantee accurate speed and rotor angle values, as well as good speed estimation outcomes with improved convergence [182,183,184,185,186,187,188].

5.3. Saliency and Signal Injection (Non-Ideal Property)-Based Method

Since the estimation of rotor speed and position depends on the back-EMF, which is dependent on rotor speed, it is evident that there is a significant issue with sensorless control schemes at low speeds. At low speeds, the back-EMF is low, making it challenging to estimate the PMSM’s speed and position. The following section discusses several approaches and strategies for sensorless PMSM drive control in the low-speed region in order to address this issue [189]. First, the high-frequency injection method, which is dependent on the saliency of the motors, is appropriate for prominent motors. Its foundation is the use of a fundamental frequency signal to excite the motor at a high frequency. The additional losses brought about by this signal are minimized by maintaining that the stator voltage and current magnitudes are low relative to the basic values. Even at moderate speeds or even at a standstill, the change in inductance becomes notable due to the huge current derivatives caused by applying an HF signal. Saturation issues and poor dynamic performance are issues with this approach. But coordinate transformation is avoided in this method, and it is not influenced by rotor position error [190,191,192,193,194,195]. Also, low-frequency signal injection methods could be used. The saliency of the motor is not necessary in the low-frequency (LF) signal injection method as it is firstly implemented for an SPMSM to calculate the position and speed of its rotor in low-speed ranges. Its foundation is the injection of a low-frequency signal into the reference d axis, which causes back-EMF ripples that are used to determine the position and speed of the rotor. The drawbacks of this estimator are slow dynamic response and depending on the motor inertia [196,197,198,199]. Since the on-line reactance measurement (INFORM) technique does not rely on the back-EMF, it is utilized for indirect flux detection in low-speed ranges and at standstill. Its basic idea is to inject space vector voltages in various directions and then measure the current response. The current response determines the inductance fluctuations, which in turn determines the rotor position. The main benefit of this approach, in terms of not relying on the back EMF, is its insensitivity to motor parameters. However, as the motor flux is thought to be sinusoidally distributed, any distortion in motor flux will result in inaccurate speed and rotor position estimates. Then, a major issue with this strategy is the current ripple [200,201,202,203]. The H sensorless control technique is an effective nonlinear control approach for handling multivariable system problems by combining control criteria such robust stability, asymptotic tracking, and disturbance attenuation into a single control problem. The foundation of the H approach is field-oriented control. Avoiding the arctangent method allows the H methodology to overcome the restriction at zero speed and low-speed range. The causal dynamic output feedback compensator and the necessary speed reference provide the rotor position information [204,205,206,207].

6. Parameter Identification Control Methods

It is commonly recognized that local self-saturation, cross-saturation mechanical power, motor form, and operational characteristics cause parameter variance in PMSM control approaches. Therefore, to eliminate issues like low-efficient operation, output power decrease, and out-of-synchronization caused by these parameters’ change, it is essential to estimate the precise values of the d- and q-axis inductances for appropriate controller design. Because of the estimation time in system operation, the parameter variation methods can be divided into offline and online schemes [2].

6.1. Offline Parameter Estimation Techniques

Firstly, the DC current decay test is one offline technique for estimating parameters. It depends on measuring the current decay applied to the motor’s stator, increasing its initial value from a big value to zero while maintaining the rotor’s alignment with the d or q axis. The DC current decay test is marked by its simplicity in the required equipment and current measuring. However, because the test is conducted at standstill and the iron losses that arise from rotating are not considered, there are mistakes in the inductance calculation figures. To improve the performance of the DC current decay test strategy, certain improvements have been made to this method based on DC-Chopper, providing the armature at standstill, pseudo random binary sequence (PRBS) voltages, PWM voltages, and DC decay. Additionally, there are suggested techniques that include measure saturation and cross-saturation effects in measuring [208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225].
The second method is the AC standstill test. Calculating the d-q axis inductances from the stator winding’s self and mutual inductances is the base of the ac standstill test. That is, one phase is subjected to an AC voltage, and the induced phase voltage is measured on a different phase. Park’s transformation is used to calculate the d- and q-axis inductances at each rotor position. Since the test is still conducted at a standstill and takes a fair amount of time, the measured values are not the true value, even if the AC standstill test is easier to use and more accurate than the DC test for inductance estimation. For these reasons, some contributions are suggested, such as applying a three-phase AC voltage source so that the vector control drive is not required and using a multi-sine AC standstill test to shorten the time needed to evaluate the effect of saturation, cross saturation, and frequency on the d- and q-axis parameters with a VSI for signal generation. Because the d- and q-axis inductances are measured while also accounting for the saturation and cross-magnetizing effects, it is hence ideal for standard laboratory investigations. Other sources suggested including iron losses into the AC standstill test. Another idea is to use two single-phase currents flowing through the suggested circuit to provide d- and q-axis currents that are comparable to those under operating conditions, bringing the measurements close to the actual values [225,226,227,228,229,230,231,232,233,234,235]. The vector control approach could also be used to determine parameters. Knowing the values of v d and v q from measuring the fundamental voltage and current amplitudes and their phases with respect to the rotor position allows one to determine the d- and q-axis flux linkages. In this system, the cross-coupling inductances are considered to be zero and the magnet flux to be constant. Even with space harmonics, the motor’s inductance could be determined using the vector control method. However, the vector-controlled approach requires extra tools such a position sensor, oscilloscope, and dynamometer [225,232,236,237,238,239,240].
The finite element method is among the most effective offline techniques for determining the motor parameters while accounting for saturation and cross-coupling effects as well as magnetic nonlinearity based on equivalent magnetic circuits [241,242,243,244,245,246]. Techniques like [247] can be used to determine the PMSM parameters without measuring torque, allowing it to account for iron losses and eliminate inaccuracies caused by the copper resistance value. High-frequency injection signal techniques to identify the inductances could decrease the detailed parameter estimation inaccuracy caused by the inverter nonlinearity influence at various rotor positions, as in [248].

6.2. Online Parameter Estimation Techniques

One of the online parameter estimation techniques is the recursive least squares (RLC). Fitting a mathematical model derived from PMSM voltage equations to a series of observed data is the foundation of recursive least squares (RLC). By reducing the sum of squares of the discrepancy between the computed and observed data, these data include the measured stator current and the indirect reference stator voltage. RLC estimates unknown parameters like R s , L d , L q , λ f , and VSI nonlinearity using known parameters like voltages and currents. Notably, RLC is utilized for sensorless control, and their control strategy relies on figuring out three values. For example, R s , L d , and L q and others calculate four values. For example, R s , L d , L q , and λ f . Controllers for techniques for calculating four values incur a significant burden and, because of the inadequate quality of the data supplied, were unable to reach a consensus on the solution. Without parameter modification, the RLC estimation’s correctness cannot be guaranteed, and because RLC uses a fixed gain, the disturbance observer may produce an unstable transient response characteristic. Real-time parameter fluctuations and inaccuracies lead to deteriorating system performance, given that the fixed gain is used by the RLS parameter estimator [127,132,249,250,251,252,253,254,255,256,257,258,259,260,261]. In terms of the networks of artificial neurons, when it comes to real-time PMSM control, ANN is a viable option for addressing issues with parameter fluctuations including inductance, armature resistance, and reverse EMF constant for motor drives. Neural network controllers are robust and computationally simple, allowing them to respond quickly to field weakening control with minimal operating range loss. ANNs are very effective in vector control for PMSMs. ANN outperforms other algorithms like MRAS and extended Kalman filter (EKF) in terms of convergence time [262,263,264,265,266,267,268,269]. MRAS that has been discussed above in the sensorless control methods section could be considered moreover as a parameter identification strategy. Some benefits of MRAS include its ability to fix inverter or converter voltage imbalance and its accuracy in estimating under a variety of scenarios. However, it has drawbacks, such as the challenge of synchronizing the PI gain at different operating points and constructing the adaptive mechanism. Efficiency will drop as a result of the stator current command being impacted by the d- and q-axis inductance inaccuracy. Therefore, in order to correct for the voltage loss component, stator resistance must be calculated [270,271,272,273,274,275,276,277,278,279,280,281,282]. The extended Kalman filter is an additional online parameter estimation technique. It was also covered in the section on sensorless control techniques. For nonlinear systems, EKF is the best recursive parameter estimator. This is because it can take into account the impact of noise disturbances. However, the EKF may not converge on the correct state if the system model is erroneous, because processing input data with noise continually necessitates a significant computational load [264,274,279,283,284,285,286]. At a high rate of convergence, the stator inductances on the d- and q-axes are evaluated. However, the motor torque, stator resistance, and magnetic flux linkage are evaluated at a sluggish rate of convergence. This approach allows the adaptive controllers’ control gains to be updated continually using the motor torque and evaluated parameters [287,288].

7. Conclusions

This paper reviewed various types of permanent magnet synchronous motors (PMSMs), highlighting their applications and key control strategies, including maximum torque per ampere (MTPA), sensorless control techniques, and parameter identification approaches. The analysis shows all these areas face common nonlinearity challenges, such as the cross-coupling effect, magnetic saturation, and temperature variations during operation. These factors cause parameter variations, particularly in inductance, which is a critical parameter in designing control systems to achieve high torque performance. Moreover, recent studies indicate that certain strategies, such as the model reference adaptive system (MRAS) and extended Kalman filter (EKF), are commonly applied for both sensorless control and parameter identification. Based on the reviewed techniques, it is recommended that an advanced sensorless control system be developed that integrates MTPA while accounting for parameter variations, thereby enhancing both the steady-state and dynamic performance of PMSMs.

Author Contributions

Conceptualization, M.M.E., M.M.A. and E.A.E.-H.; Methodology, M.M.A., M.M.E. and E.A.E.-H.; Software, M.M.A., M.M.E. and E.A.E.-H.; Validation, M.M.E., M.M.A. and E.A.E.-H.; Formal analysis, M.M.A., M.M.E. and E.A.E.-H.; Investigation, M.M.A., M.M.E. and E.A.E.-H.; Data curation, M.M.E., M.M.A. and E.A.E.-H.; Writing—original draft, M.M.A.; Writing—review & editing, M.M.E., M.M.A. and E.A.E.-H.; Visualization, M.M.A., M.M.E. and E.A.E.-H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Poorfakhraei, A.; Narimani, M.; Emadi, A. A review of modulation and control techniques for multilevel inverters in traction applications. IEEE Access 2021, 9, 24187–24204. [Google Scholar] [CrossRef]
  2. Ahn, H.; Park, H.; Kim, C.; Lee, H. A Review of State-of-the-art Techniques for PMSM Parameter Identification. J. Electr. Eng. Technol. 2020, 15, 1177–1187. [Google Scholar] [CrossRef]
  3. Rafaq, M.S.; Jung, J.-W. A comprehensive review of state-of-the-art parameter estimation techniques for permanent magnet synchronous motors in wide speed range. IEEE Trans. Ind. Inform. 2019, 16, 4747–4758. [Google Scholar] [CrossRef]
  4. Li, Y.; Hu, H.; Shi, P. A Review of Position Sensorless Compound Control for PMSM Drives. World Electr. Veh. J. 2023, 14, 34. [Google Scholar] [CrossRef]
  5. Singh, S.; Tiwari, A. Various techniques of sensorless speed control of PMSM: A review. In Proceedings of the 2017 Second International Conference on Electrical, Computer and Communication Technologies (ICECCT), Coimbatore, India, 22–24 February 2017. [Google Scholar]
  6. Murakami, H.; Honda, Y.; Kiriyama, H.; Morimoto, S.; Takeda, Y. The performance comparison of SPMSM, IPMSM and SynRM in use as air-conditioning compressor. In Proceedings of the Conference Record of the 1999 IEEE Industry Applications Conference. Thirty-Forth IAS Annual Meeting (Cat. No. 99CH36370), Phoenix, AZ, USA, 3–7 October 1999. [Google Scholar]
  7. Shankar, V.K.A.; Umashankar, S.; Paramasivam, S.; Hanigovszki, N. A comprehensive review on energy efficiency enhancement initiatives in centrifugal pumping system. Appl. Energy 2016, 181, 495–513. [Google Scholar] [CrossRef]
  8. Diab, A.M.; Yeoh, S.S.; Bozhko, S.; Guo, F.; Gerada, C.; Galea, M. Discrete flux weakening controller design for future aircraft electric starter/generator system operating in high-speed range. In Proceedings of the 2021 IEEE Workshop on Electrical Machines Design, Control and Diagnosis (WEMDCD), Modena, Italy, 8–9 April 2021. [Google Scholar]
  9. Belhadj, J.; Roboam, X. Investigation of different methods to control a small variable-speed wind turbine with PMSM drives. J. Energy Resour. Technol. 2007, 129, 200–213. [Google Scholar] [CrossRef]
  10. Soyed, A.; Aouiti, A.; Bacha, F. Study and Analysis of the Indirect Matrix Converter for Wind Energy System based on FOC-SVM and MPC Control Techniques. In Proceedings of the 2023 9th International Conference on Control, Decision and Information Technologies (CoDIT), Rome, Italy, 3–6 July 2023. [Google Scholar]
  11. Nouraldin, N.A.; Chebaani, M.; Számel, L.; Abdelwahab, S.A.M.; Abdellatif, W.S. Experimental investigation of predictive control for PMSM-based wind turbine generation system. Comput. Electr. Eng. 2024, 119, 109554. [Google Scholar] [CrossRef]
  12. Zhan, H.; Zhu, Z.-Q.; Odavic, M.; Li, Y. A novel zero-sequence model-based sensorless method for open-winding PMSM with common DC bus. IEEE Trans. Ind. Electron. 2016, 63, 6777–6789. [Google Scholar] [CrossRef]
  13. Rahman, A.; Dutta, R.; Chu, G.; Xiao, D.; Thippiripati, V.K.; Rahman, M.F. Open-winding permanent magnet synchronous generator for renewable energy—A review. Energies 2023, 16, 5268. [Google Scholar] [CrossRef]
  14. Shneen, S.W.; Mao, C.; Wang, D. Advanced optimal PSO, Fuzzy and PI controller with PMSM and WTGS at 5Hz side of generation and 50 Hz Side of Grid. Int. J. Power Electron. Drive Syst. (IJPEDS) 2016, 7, 173. [Google Scholar] [CrossRef]
  15. Mani, P.; Rajan, R.; Shanmugam, L.; Joo, Y.H. Adaptive fractional fuzzy integral sliding mode control for PMSM model. IEEE Trans. Fuzzy Syst. 2018, 27, 1674–1686. [Google Scholar] [CrossRef]
  16. Azom, M.A.; Khan, M.Y. Recent developments in control and simulation of permanent magnet synchronous motor systems. Control Syst. Optim. Lett. 2025, 3, 84–91. [Google Scholar] [CrossRef]
  17. Vagati, A.; Pellegrino, G.; Guglielmi, P. Comparison between SPM and IPM motor drives for EV application. In Proceedings of the XIX International Conference on Electrical Machines—ICEM 2010, Rome, Italy, 6–8 September 2010. [Google Scholar]
  18. Pellegrino, G.; Vagati, A.; Guglielmi, P.; Boazzo, B. Performance comparison between surface-mounted and interior PM motor drives for electric vehicle application. IEEE Trans. Ind. Electron. 2011, 59, 803–811. [Google Scholar] [CrossRef]
  19. Reddy, P.B.; El-Refaie, A.M.; Huh, K.-K.; Tangudu, J.K.; Jahns, T.M. Comparison of interior and surface PM machines equipped with fractional-slot concentrated windings for hybrid traction applications. IEEE Trans. Energy Convers. 2012, 27, 593–602. [Google Scholar] [CrossRef]
  20. Wang, J.; Yuan, X.; Atallah, K. Design optimization of a surface-mounted permanent-magnet motor with concentrated windings for electric vehicle applications. IEEE Trans. Veh. Technol. 2012, 62, 1053–1064. [Google Scholar] [CrossRef]
  21. You-Young, C.; Ki-Doek, L.; Ik-Sang, J.; Mi-Jung, K.; Sang-Hwan, H.; Ju, L.; Kwang-Cheol, K. Research on the Output Characteristics of IPMSM according to the Pole-Slot Combinations. Energy Procedia 2012, 14, 1187–1192. [Google Scholar] [CrossRef]
  22. Fasolo, A.; Alberti, L.; Bianchi, N. Performance comparison between switching-flux and IPM machines with rare-earth and ferrite PMs. IEEE Trans. Ind. Appl. 2014, 50, 3708–3716. [Google Scholar] [CrossRef]
  23. Dong, J.; Huang, Y.; Jin, L.; Lin, H. Comparative study of surface-mounted and interior permanent-magnet motors for high-speed applications. IEEE Trans. Appl. Supercond. 2016, 26, 5200304. [Google Scholar] [CrossRef]
  24. Thampi, P.; Kiran, C. A review on controlling techniques for permanent magnet synchronous motor (PMSM) and current state of the art in the research area. Appl. Electron. 2019, 7, 8–17. [Google Scholar] [CrossRef]
  25. Thike, R.; Pillay, P. Mathematical model of an interior PMSM with aligned magnet and reluctance torques. IEEE Trans. Transp. Electrif. 2020, 6, 647–658. [Google Scholar] [CrossRef]
  26. Kovacik, M.; Rafajdus, P.; Kocan, S. Comparison of various PMSM rotor topologies for high-speed drives in automotive applications. Transp. Res. Procedia 2021, 55, 995–1002. [Google Scholar] [CrossRef]
  27. Madichetty, S.; Mishra, S.; Basu, M. New trends in electric motors and selection for electric vehicle propulsion systems. IET Electr. Syst. Transp. 2021, 11, 186–199. [Google Scholar] [CrossRef]
  28. Bianchi, N.; Carlet, P.G.; Cinti, L.; Ortombina, L. A Review about Flux-Weakening Operating Limits and Control Techniques for Synchronous Motor Drives. Energies 2022, 15, 1930. [Google Scholar] [CrossRef]
  29. Yang, K.; Zhang, L.; Wang, M.; Du, C. Performance comparison and optimization of a PMSM based on hybrid-type permanent magnet with two kinds of rotor topology. Energies 2024, 17, 557. [Google Scholar] [CrossRef]
  30. Shuaibu, I.; Wei, E.H.T.; Kannan, R.; Samaila, Y.A. Advancements in axial flux permanent magnet machines utilizing coreless technology: A systematic review. Ain Shams Eng. J. 2024, 15, 103091. [Google Scholar] [CrossRef]
  31. Yang, C.; Cai, W.; Shao, B. Study on flow field of oil-cooling permanent magnet synchronous motor with hairpin winding using porous medium model. IET Electr. Power Appl. 2024, 18, 1083–1093. [Google Scholar] [CrossRef]
  32. Yang, G.; Wu, H.; Wang, D.; Zhang, P.; Jiang, S.; Zhang, Y. Research on Rotor Loss and Rotor Structure Optimization of Rare Earth Permanent Magnet Synchronous Motors in the Weak Field Area. Energies 2023, 16, 7423. [Google Scholar] [CrossRef]
  33. Hemsen, J.; Nowak, N.; Eckstein, L. Production cost modeling for permanent magnet synchronous machines for electric vehicles. Automot. Engine Technol. 2023, 8, 109–126. [Google Scholar] [CrossRef]
  34. Tomše, T.; Kubelka, P.; Moreno López, R.; Fleissner, P.; Grau, L.; Zaplotnik, M.; Burkhardt, C. Magnetic Performance and Anticorrosion Coating Stability of Thermally Demagnetized Nd-Fe-B Permanent Magnets for Reuse Applications. Materials 2024, 17, 5927. [Google Scholar] [CrossRef] [PubMed]
  35. Sui, Y.; Zheng, P.; Yin, Z.; Wang, M.; Wang, C. Open-circuit fault-tolerant control of five-phase PM machine based on reconfiguring maximum round magnetomotive force. IEEE Trans. Ind. Electron. 2018, 66, 48–59. [Google Scholar] [CrossRef]
  36. Wu, L.; Yi, J.; Lyu, Z.; Hu, S. Fault-tolerant operation for five-phase permanent magnet synchronous machine drives with five-phase six-leg inverter. IET Electr. Power Appl. 2024, 18, 489–502. [Google Scholar] [CrossRef]
  37. Gao, C.; Zheng, H.; Kaku, C. Design of Fault-Tolerant Control Strategy for Dual-Winding Permanent Magnet Synchronous Motor in Vehicles; SAE Technical Paper; SAE International: Warrendale, PA, USA, 2025. [Google Scholar]
  38. Mao, Y.; Wang, X.; Shi, P.; Lu, L.; Wang, Z. Optimization-based maximum-torque fault-tolerant control of dual three-phase PMSM drives under open-phase fault. IEEE Trans. Power Electron. 2022, 38, 3653–3663. [Google Scholar] [CrossRef]
  39. Lu, L.; Wang, X.; Jin, L.; Mao, Y.; Bao, Q.; Wang, Z. Fault-tolerant optimization control of open-winding dual three-phase PMSM drives in full operation range. IEEE Trans. Power Electron. 2024, 40, 3385–3394. [Google Scholar] [CrossRef]
  40. Mohammad, M.S.; Lee, D.-C. Optimized Maximum Torque and Minimum Loss Fault-Tolerant Control Schemes for Dual Three-Phase PMSM. In Proceedings of the 2025 IEEE Applied Power Electronics Conference and Exposition (APEC), Atlanta, GA, USA, 16–20 March 2025. [Google Scholar]
  41. Cruz, S.M.; Mendes, A.M.S. Fault-tolerant predictive current control of six-phase PMSMs with minimal reconfiguration requirements. IEEE J. Emerg. Sel. Top. Power Electron. 2022, 11, 2084–2093. [Google Scholar] [CrossRef]
  42. Huang, W.; Zhu, X.; Zhang, H.; Hua, W. Generalized fault-tolerant model predictive control of five-phase PMSM drives under single/two open-switch faults. IEEE Trans. Ind. Electron. 2023, 70, 7569–7579. [Google Scholar] [CrossRef]
  43. Chen, L.; Chen, M.; Li, B.; Sun, X.; Jiang, F. Optimized fault-tolerant control of dual three-phase PMSM under open-switch faults. Energies 2024, 17, 5198. [Google Scholar] [CrossRef]
  44. Lenz, R.; Kugi, A.; Kemmetmüller, W. Optimal torque control with radial force compensation for multiphase PMSMs under an open-circuit fault. IFAC-PapersOnLine 2023, 56, 4412–4417. [Google Scholar] [CrossRef]
  45. Lenz, R.; Deutschmann-Olek, A.; Kugi, A.; Kemmetmüller, W. Optimal fault-tolerant control with radial force compensation for multiple open-circuit faults in multiphase PMSMs-A comparison of n-phase and multiple three-phase systems. Control Eng. Pract. 2024, 147, 105924. [Google Scholar] [CrossRef]
  46. Ding, B.; Lu, Y.; Lai, C.; Feng, G. Single Open-Phase Fault Tolerant Control of Salient Dual Three-Phase PMSMs With Maximized Torque to Total Loss Ratio Considering Peak Phase Current Limit. IEEE Trans. Ind. Electron. 2025, 72, 6852–6864. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Lin, F.; Zhang, Z.; Xu, L. Direct voltage vector control for field weakening operation of PM machines. In Proceedings of the 2010 Power and Energy Conference at Illinois (PECI), Urbana, IL, USA, 12–13 February 2010. [Google Scholar]
  48. Hu, D.; Zhu, L.; Xu, L. Maximum Torque per Volt operation and stability improvement of PMSM in deep flux-weakening Region. In Proceedings of the 2012 IEEE Energy Conversion Congress and Exposition (ECCE), Raleigh, NC, USA, 15–20 September 2012. [Google Scholar]
  49. Li, H.; Qian, Y.; Asgarpoor, S.; Sharif, H. Simulation study on on-line MTPA/MTPV trajectory tracking in PMSMs with power management. Electr. Power Components Syst. 2020, 48, 241–255. [Google Scholar] [CrossRef]
  50. Miguel-Espinar, C.; Heredero-Peris, D.; Gross, G.; Llonch-Masachs, M.; Montesinos-Miracle, D. Maximum torque per voltage flux-weakening strategy with speed limiter for PMSM drives. IEEE Trans. Ind. Electron. 2020, 68, 9254–9264. [Google Scholar] [CrossRef]
  51. Xia, Z.; Nalakath, S.; Tarvirdilu-Asl, R.; Sun, Y.; Wiseman, J.; Emadi, A. Online optimal tracking method for interior permanent magnet machines with improved MTPA and MTPV in whole speed and torque ranges. IEEE Trans. Power Electron. 2020, 35, 9753–9769. [Google Scholar] [CrossRef]
  52. Bianchini, C.; Bisceglie, G.; Torreggiani, A.; Davoli, M.; Macrelli, E.; Bellini, A.; Frigieri, M. Effects of the Magnetic Model of Interior Permanent Magnet Machine on MTPA, Flux Weakening and MTPV Evaluation. Machines 2023, 11, 77. [Google Scholar] [CrossRef]
  53. Wang, S.; Kang, J.; Degano, M.; Galassini, A.; Gerada, C. An accurate wide-speed range control method of IPMSM considering resistive voltage drop and magnetic saturation. IEEE Trans. Ind. Electron. 2019, 67, 2630–2641. [Google Scholar] [CrossRef]
  54. Dwivedi, S.K.; Laursen, M.; Hansen, S. Voltage vector based control for PMSM in industry applications. In Proceedings of the 2010 IEEE International Symposium on Industrial Electronics, Bari, Italy, 4–7 July 2010. [Google Scholar]
  55. Bida, V.M.; Samokhvalov, D.V.; Al-Mahturi, F.S. PMSM vector control techniques—A survey. In Proceedings of the 2018 IEEE Conference of Russian Young Researchers in Electrical and Electronic Engineering (EIConRus), Moscow, Russia, 29 January–1 February 2018. [Google Scholar]
  56. Korkmaz, F.; Topaloğlu, İ.; Cakir, M.F.; Gürbüz, R. Comparative performance evaluation of FOC and DTC controlled PMSM drives. In Proceedings of the 4th International Conference on Power Engineering, Energy and Electrical Drives, Istanbul, Turkey, 13–17 May 2013. [Google Scholar]
  57. Geweth, D.; Zanelli, A.; Frison, G.; Vollmer, U.; Diehl, M. Field oriented economic model predictive control for permanent magnet synchronous motors. IFAC-PapersOnLine 2020, 53, 9093–9099. [Google Scholar] [CrossRef]
  58. Akrami, M.; Jamshidpour, E.; Pierfederici, S.; Frick, V. Comparison of Flatness-based Control and Field-Oriented Control for PMSMs in Automotive Water Pump Application. In Proceedings of the 2023 IEEE International Conference on Environment and Electrical Engineering and 2023 IEEE Industrial and Commercial Power Systems Europe (EEEIC/I&CPS Europe), Madrid, Spain, 6–9 June 2023. [Google Scholar]
  59. Kumar, N.; Goutham, P.; Chandrakala, K.V.; Saravanan, S.; Shankar, A. Sliding Mode-Based Field-Oriented Control Using SVPWM for the Speed Control of Permanent Magnet Synchronous Motor. In Proceedings of the 2024 Second International Conference on Emerging Trends in Information Technology and Engineering (ICETITE), Vellore, India, 22–23 February 2024. [Google Scholar]
  60. Mocanu, R.; Onea, A. Passivity Based Torque Control of PMSM used in electrical vehicles. In Proceedings of the 2015 19th International Conference on System Theory, Control and Computing (ICSTCC), Cheile Gradistei, Romania, 14–16 October 2015. [Google Scholar]
  61. Ouhrouche, M.; Errouissi, R.; Trzynadlowski, A.M.; Tehrani, K.A.; Benzaioua, A. A novel predictive direct torque controller for induction motor drives. IEEE Trans. Ind. Electron. 2016, 63, 5221–5230. [Google Scholar] [CrossRef]
  62. Morales-Caporal, R.; Leal-López, M.E.; de Jesús Rangel-Magdaleno, J.; Sandre-Hernández, O.; Cruz-Vega, I. Direct torque control of a PMSM-drive for electric vehicle applications. In Proceedings of the 2018 International Conference on Electronics, Communications and Computers (CONIELECOMP), Cholula, Mexico, 21–23 February 2018. [Google Scholar]
  63. Ghamri, A.; Boumaaraf, R.; Benchouia, M.T.; Mesloub, H.; Goléa, A.; Goléa, N. Comparative study of ANN DTC and conventional DTC controlled PMSM motor. Math. Comput. Simul. 2020, 167, 219–230. [Google Scholar] [CrossRef]
  64. Karthikeyan, A.; Prabhakaran, K.; Nagamani, C. Four quardrant operation of direct torque controlled PMSM drive using speed loop PDFF controller. In Proceedings of the 2018 International Conference on Power, Instrumentation, Control and Computing (PICC), Thrissur, India, 18–20 January 2018. [Google Scholar]
  65. Ghanayem, H.; Alathamneh, M.; Nelms, R.J. Decoupled Speed and Flux Control of Three-Phase PMSM Based on the Proportional-Resonant Control Method. Energies 2023, 16, 1053. [Google Scholar] [CrossRef]
  66. Zhu, Y.; Gu, W.; Lu, K.; Wu, Z. Vector control of asymmetric dual three-phase PMSM in full modulation range. IEEE Access 2020, 8, 104479–104493. [Google Scholar] [CrossRef]
  67. Wang, H.; Wu, X.; Zheng, X.; Yuan, X. Virtual voltage vector based model predictive control for a nine-phase open-end winding PMSM with a common DC bus. IEEE Trans. Ind. Electron. 2021, 69, 5386–5397. [Google Scholar] [CrossRef]
  68. Belkhier, Y.; Abdelyazid, A.; Oubelaid, A.; Khosravi, N.; Bajaj, M.; Vishnuram, P.; Zaitsev, I. Experimental analysis of passivity-based control theory for permanent magnet synchronous motor drive fed by grid power. IET Control. Theory Appl. 2024, 18, 495–510. [Google Scholar] [CrossRef]
  69. Colby, R.S.; Novotny, D. An efficiency-optimizing permanent-magnet synchronous motor drive. IEEE Trans. Ind. Appl. 1988, 24, 462–469. [Google Scholar] [CrossRef]
  70. Perera, P.C.; Blaabjerg, F.; Pedersen, J.K.; Thogersen, P. A sensorless, stable V/f control method for permanent-magnet synchronous motor drives. IEEE Trans. Ind. Appl. 2003, 39, 783–791. [Google Scholar] [CrossRef]
  71. Sue, S.-M.; Hung, T.-W.; Liaw, J.-H.; Li, Y.-F.; Sun, C.-Y. A new MTPA control strategy for sensorless V/f controlled PMSM drives. In Proceedings of the 2011 6th IEEE Conference on Industrial Electronics and Applications, Beijing, China, 21–23 June 2011. [Google Scholar]
  72. Jafari, S.H.; Corzine, K.A.; Huang, J. Efficiency optimization of a sensorless V/f control method for PMSM. In Proceedings of the 3rd IEEE International Symposium on Sensorless Control for Electrical Drives (SLED 2012), Milwaukee, WI, USA, 21–22 September 2012. [Google Scholar]
  73. Patel, F.; Ahir, J.; Patel, P. Control Theory For Permanent Magnet Synchronous Motor—A Review. Asian J. Converg. Technol. (AJCT) 2019, V. [Google Scholar]
  74. Itoh, J.-I.; Nomura, N.; Ohsawa, H. A comparison between V/f control and position-sensorless vector control for the permanent magnet synchronous motor. In Proceedings of the Power Conversion Conference-Osaka 2002 (Cat. No. 02TH8579), Osaka, Japan, 2–5 April 2002. [Google Scholar]
  75. OS, S.B.; Khalid, M.; Binojkumar, A. Evaluation of Single Current Regulator Based Field Weakening Strategy For PMSM Drives and Experimental Validation using HIL. In Proceedings of the 2022 Second International Conference on Next Generation Intelligent Systems (ICNGIS), Kottayam, India, 29–31 July 2022. [Google Scholar]
  76. Koraqi, L.; Koteich, M. Torque Maximization Control of Permanent Magnet Synchronous Motors Using a Two-Stage Virtual Signal Injection. In Proceedings of the 2022 XIV International Symposium on Industrial Electronics and Applications (INDEL), Banja Luka, Bosnia and Herzegovina, 9–11 November 2022. [Google Scholar]
  77. Tursini, M.; Chiricozzi, E.; Petrella, R. Feedforward flux-weakening control of surface-mounted permanent-magnet synchronous motors accounting for resistive voltage drop. IEEE Trans. Ind. Electron. 2009, 57, 440–448. [Google Scholar] [CrossRef]
  78. Pellegrino, G.; Armando, E.; Guglielmi, P. Direct flux field-oriented control of IPM drives with variable DC link in the field-weakening region. IEEE Trans. Ind. Appl. 2009, 45, 1619–1627. [Google Scholar] [CrossRef]
  79. Jahns, T.M.; Kliman, G.B.; Neumann, T.W. Interior permanent-magnet synchronous motors for adjustable-speed drives. IEEE Trans. Ind. Appl. 1986, 22, 738–747. [Google Scholar] [CrossRef]
  80. Uddin, M.N.; Rahman, M.A. High-speed control of IPMSM drives using improved fuzzy logic algorithms. IEEE Trans. Ind. Electron. 2007, 54, 190–199. [Google Scholar] [CrossRef]
  81. Macminn, S.R.; Jahns, T.M. Control techniques for improved high-speed performance of interior PM synchronous motor drives. In Proceedings of the Conference Record of the 1988 IEEE Industry Applications Society Annual Meeting, Pittsburgh, PA, USA, 2–7 October 1988. [Google Scholar]
  82. Jahns, T.M. Flux-weakening regime operation of an interior permanent magnet synchronous motor drive. In Proceedings of the 1986 Annual Meeting Industry Applications Society, Denver, CO, USA, 28 September–3 October 1986. [Google Scholar]
  83. Song, J.-H.; Kim, J.-M.; Sul, S.-K. A new robust SPMSM control to parameter variations in flux weakening region. In Proceedings of the 1996 IEEE IECON. 22nd International Conference on Industrial Electronics, Control, and Instrumentation, Taipei, Taiwan, 5–10 August 1996. [Google Scholar]
  84. Kim, J.-M.; Sul, S.-K. Speed control of interior permanent magnet synchronous motor drive for the flux weakening operation. IEEE Trans. Ind. Appl. 1997, 33, 43–48. [Google Scholar] [CrossRef]
  85. Bae, B.-H.; Patel, N.; Schulz, S.; Sul, S.-K. New field weakening technique for high saliency interior permanent magnet motor. In Proceedings of the 38th IAS Annual Meeting on Conference Record of the Industry Applications Conference, Salt Lake City, UT, USA, 12–16 October 2003. [Google Scholar]
  86. Lenke, R.U.; De Doncker, R.W.; Kwak, M.-S.; Kwon, T.-S.; Sul, S.-K. Field weakening control of interior permanent magnet machine using improved current interpolation technique. In Proceedings of the 2006 37th IEEE Power Electronics Specialists Conference, Jeju, Republic of Korea, 18 June 2006. [Google Scholar]
  87. Kwon, T.-S.; Sul, S.-K. Novel antiwindup of a current regulator of a surface-mounted permanent-magnet motor for flux-weakening control. IEEE Trans. Ind. Appl. 2006, 42, 1293–1300. [Google Scholar] [CrossRef]
  88. Sue, S.-M.; Pan, C.-T. Voltage-constraint-tracking-based field-weakening control of IPM synchronous motor drives. IEEE Trans. Ind. Electron. 2008, 55, 340–347. [Google Scholar] [CrossRef]
  89. Olarescu, N.; Weinmann, M.; Zeh, S.; Musuroi, S. Novel flux weakening control algorithm for PMSMS. In Proceedings of the 2009 International Conference on Power Engineering, Energy and Electrical Drives, Lisbon, Portugal, 18–20 March 2009. [Google Scholar]
  90. Dolecek, R.; Simanek, J.; Novak, J.; Cerny, O. Dynamics of a feedback optimal-current-vector flux-weakening strategy for traction permanent-magnet synchronous motors. In Proceedings of the 2009 8th International Symposium on Advanced Electromechanical Motion Systems & Electric Drives Joint Symposium, Lille, France, 1–3 July 2009. [Google Scholar]
  91. Cheng, B.; Tesch, T.R. Torque feedforward control technique for permanent-magnet synchronous motors. IEEE Trans. Ind. Electron. 2010, 57, 969–974. [Google Scholar] [CrossRef]
  92. Lu, D.; Kar, N.C. A review of flux-weakening control in permanent magnet synchronous machines. In Proceedings of the 2010 IEEE Vehicle Power and Propulsion Conference, Lille, France, 1–3 September 2010. [Google Scholar]
  93. Azizi, A.; Akhbari, M.; Danyali, S.; Tohidinejad, Z.; Shirkhani, M. A review on topology and control strategies of high-power inverters in large-scale photovoltaic power plants. Heliyon 2025, 11, e42334. [Google Scholar] [CrossRef] [PubMed]
  94. Taha, T.A.; Shalaby, M.; Wahab, N.I.A.; Zaynal, H.I.; Hassan, M.K.; Al-Sowayan, S.; Alawad, M.A. Recent Advancements in Multilevel Inverters: Topologies, Modulation Techniques, and Emerging Applications. Symmetry 2025, 17, 1010. [Google Scholar] [CrossRef]
  95. Nguyen, K.-A.; Hieu, L.; Pham, V.-T.; Ngo, V.-Q. Enhanced model predictive current framework for a PMSM powered by three-level NPC inverters with DC-link voltage self-balancing mechanisms. Results Eng. 2025, 27, 105918. [Google Scholar] [CrossRef]
  96. Dong, H.; Zhang, Y. A low-complexity double vector model predictive current control for permanent magnet synchronous motors. Energies 2023, 17, 147. [Google Scholar] [CrossRef]
  97. Qasim, M.M.; Otten, D.M.; Lang, J.H.; Kirtley, J.L.; Perreault, D.J. Comparison of inverter topologies for high-speed motor drive applications. IEEE Trans. Power Electron. 2024, 39, 7404–7422. [Google Scholar] [CrossRef]
  98. An, Q.; Lu, Y.; Zhao, M. Review of Key Technologies of the High-Speed Permanent Magnet Motor Drive. Energies 2024, 17, 5252. [Google Scholar] [CrossRef]
  99. Wellmann, C.; Khaleel, A.R.; Brinkmann, T.; Wahl, A.; Monissen, C.; Eisenbarth, M.; Andert, J. Electric machine co-optimization for EV drive technology development: Integrating Bayesian optimization and nonlinear model predictive control. eTransportation 2025, 23, 100392. [Google Scholar] [CrossRef]
  100. Liu, X.; Yu, H.; Yu, J.; Zhao, L. Combined speed and current terminal sliding mode control with nonlinear disturbance observer for PMSM drive. IEEE Access 2018, 6, 29594–29601. [Google Scholar] [CrossRef]
  101. Wang, Y.; Yu, H.; Liu, Y. Speed-current single-loop control with overcurrent protection for PMSM based on time-varying nonlinear disturbance observer. IEEE Trans. Ind. Electron. 2021, 69, 179–189. [Google Scholar] [CrossRef]
  102. Zhang, Z.; Liu, X.; Yu, J.; Yu, H. Time-varying disturbance observer based improved sliding mode single-loop control of PMSM drives with a hybrid reaching law. IEEE Trans. Energy Convers. 2023, 38, 2539–2549. [Google Scholar] [CrossRef]
  103. Zhang, Z.; Liu, X. An improved super-twisting sliding mode single-loop control with current-constraint for PMSM based on two-time scale disturbance observer. IEEE Trans. Transp. Electrif. 2023, 10, 5389–5399. [Google Scholar] [CrossRef]
  104. Wang, Z.; Dai, B.; Guo, Z.; Yang, J. Single-Loop Current-Constrained Speed Regulation for PMSM via a Switching Controller. IEEE Trans. Transp. Electrif. 2024, 11, 5974–5982. [Google Scholar] [CrossRef]
  105. Liu, J.; Yang, J.; Li, S.; Wang, X. Single-loop robust model predictive speed regulation of PMSM based on exogenous signal preview. IEEE Trans. Ind. Electron. 2023, 70, 12719–12729. [Google Scholar] [CrossRef]
  106. Zhang, J.; Ren, W.; Sun, X.-M. Extended-state-observer-based nonlinear control for PMSM servo systems with current constraints and voltage saturations. IEEE Trans. Transp. Electrif. 2023, 10, 2713–2726. [Google Scholar] [CrossRef]
  107. Liu, J.; Li, M.; Xie, E. Non-cascade structure equivalent SMC for PMSM driving based on improved ESO. IEEE Trans. Power Electron. 2024, 40, 611–624. [Google Scholar] [CrossRef]
  108. Dai, B.; Wang, Z.; Zhao, J.; Li, S. Critical Current-Constrained Continuous Nonsingular Terminal Sliding Mode Control for PMSM Based on Control Barrier Function. IEEE Trans. Power Electron. 2025, 40, 15093–15103. [Google Scholar] [CrossRef]
  109. Foo, G.; Sayeef, S.; Rahman, M.J. Low-speed and standstill operation of a sensorless direct torque and flux controlled IPM synchronous motor drive. IEEE Trans. Energy Convers. 2010, 25, 25–33. [Google Scholar] [CrossRef]
  110. Inoue, Y.; Morimoto, S.; Sanada, M. A Reference Value Calculation Scheme for Torque and Flux and an Anti-Windup Implementation of Torque Controller for Direct Torque Control of Permanent Magnet Synchronous Motor. IEEJ Trans. Ind. Appl. 2010, 130, 777–784. [Google Scholar] [CrossRef]
  111. Inoue, Y.; Morimoto, S.; Sanada, M. Control method suitable for direct-torque-control-based motor drive system satisfying voltage and current limitations. IEEE Trans. Ind. Appl. 2012, 48, 970–976. [Google Scholar] [CrossRef]
  112. Ahmed, A.; Sozer, Y.; Hamdan, M. Maximum torque per ampere control for buried magnet PMSM based on DC-link power measurement. IEEE Trans. Power Electron. 2016, 32, 1299–1311. [Google Scholar] [CrossRef]
  113. Inoue, T.; Inoue, Y.; Morimoto, S.; Sanada, M. Maximum torque per ampere control of a direct torque-controlled PMSM in a stator flux linkage synchronous frame. IEEE Trans. Ind. Appl. 2016, 52, 2360–2367. [Google Scholar] [CrossRef]
  114. Jin, N.; Wang, X.; Gao, H.; Liu, J. Sliding mode based speed regulating of PMSM MTPA control system for electrical vehicles. In Proceedings of the 2011 International Conference on Electronic & Mechanical Engineering and Information Technology, Harbin, China, 12–14 August 2011. [Google Scholar]
  115. Zhang, Z.; Wang, C.; Zhou, M.; You, X. Parameters compensation of permanent magnet synchronous motor in flux-weakening region for rail transit. IEEE Trans. Power Electron. 2020, 35, 12509–12521. [Google Scholar] [CrossRef]
  116. Ge, H.; Miao, Y.; Bilgin, B.; Nahid-Mobarakeh, B.; Emadi, A. Speed range extended maximum torque per ampere control for PM drives considering inverter and motor nonlinearities. IEEE Trans. Power Electron. 2016, 32, 7151–7159. [Google Scholar] [CrossRef]
  117. Cao, M.; Egashira, J.; Kaneko, K. High efficiency control of IPMSM for electric motorcycles. In Proceedings of the 2009 IEEE 6th International Power Electronics and Motion Control Conference, Wuhan, China, 17–20 May 2009. [Google Scholar]
  118. Kim, H.-S.; Lee, Y.; Sul, S.-K.; Yu, J.; Oh, J. Online MTPA control of IPMSM based on robust numerical optimization technique. IEEE Trans. Ind. Appl. 2019, 55, 3736–3746. [Google Scholar] [CrossRef]
  119. Pan, C.-T.; Sue, S.-M. A linear maximum torque per ampere control for IPMSM drives over full-speed range. IEEE Trans. Energy Convers. 2005, 20, 359–366. [Google Scholar] [CrossRef]
  120. Huang, S.; Chen, Z.; Huang, K.; Gao, J. Maximum torque per ampere and flux-weakening control for PMSM based on curve fitting. In Proceedings of the 2010 IEEE Vehicle Power and Propulsion Conference, Lille, France, 1–3 September 2010. [Google Scholar]
  121. Sun, J.; Luo, X.; Ma, X. Realization of maximum torque per ampere control for IPMSM based on inductance segmentation. IEEE Access 2018, 6, 66088–66094. [Google Scholar] [CrossRef]
  122. Liu, J.; Zhang, Y. A universal maximum torque per ampere method for permanent magnet synchronous machines. In Proceedings of the 2019 22nd International Conference on Electrical Machines and Systems (ICEMS), Harbin, China, 11–14 August 2019. [Google Scholar]
  123. Li, Z.; Li, H. MTPA control of PMSM system considering saturation and cross-coupling. In Proceedings of the 2012 15th International Conference on Electrical Machines and Systems (ICEMS), Sapporo, Japan, 21–24 October 2012. [Google Scholar]
  124. Miao, Y.; Ge, H.; Preindl, M.; Ye, J.; Cheng, B.; Emadi, A. MTPA fitting and torque estimation technique based on a new flux-linkage model for interior-permanent-magnet synchronous machines. IEEE Trans. Ind. Appl. 2017, 53, 5451–5460. [Google Scholar] [CrossRef]
  125. Inoue, T.; Inoue, Y.; Morimoto, S.; Sanada, M. Mathematical model for MTPA control of permanent-magnet synchronous motor in stator flux linkage synchronous frame. IEEE Trans. Ind. Appl. 2015, 51, 3620–3628. [Google Scholar] [CrossRef]
  126. Xia, J.; Xu, J.; Huang, Y.; Jatskevich, J.; Guo, Y.; Zhang, X. Parameter-Estimation-Based Adaptive MTPA Control for Interior Permanent Magnet Synchronous Motors. In Proceedings of the 2019 6th International Conference on Control, Decision and Information Technologies (CoDIT), Paris, France, 23–26 April 2019. [Google Scholar]
  127. Nguyen, Q.K.; Petrich, M.; Roth-Stielow, J. Implementation of the MTPA and MTPV control with online parameter identification for a high speed IPMSM used as traction drive. In Proceedings of the 2014 International Power Electronics Conference (IPEC-Hiroshima 2014-ECCE ASIA), Hiroshima, Japan, 18–21 May 2014. [Google Scholar]
  128. Morimoto, S.; Shimizu, Y.; Inoue, Y.; Sanada, M. Experimental Evaluation of the Online MTPA Angle Search Method Based on Flux-linkage Plane Estimation for IPMSMs. IEEJ J. Ind. Appl. 2024, 13, 23009975. [Google Scholar] [CrossRef]
  129. Cao, M.; Migita, H. MTPA Control of IPMSM with Online Parameter Estimation. In Proceedings of the 2019 21st European Conference on Power Electronics and Applications (EPE’19 ECCE Europe), Genova, Italy, 3–5 September 2019. [Google Scholar]
  130. Shinohara, A.; Inoue, Y.; Morimoto, S.; Sanada, M. Direct calculation method of reference flux linkage for maximum torque per ampere control in DTC-based IPMSM drives. IEEE Trans. Power Electron. 2016, 32, 2114–2122. [Google Scholar] [CrossRef]
  131. Niazi, P.; Toliyat, H.A.; Goodarzi, A. Robust maximum torque per ampere (MTPA) control of PM-assisted SynRM for traction applications. IEEE Trans. Veh. Technol. 2007, 56, 1538–1545. [Google Scholar] [CrossRef]
  132. Phowanna, P.; Boonto, S.; Konghirun, M. Online parameter identification method for IPMSM drive with MTPA. In Proceedings of the 2015 18th International Conference on Electrical Machines and Systems (ICEMS), Pattaya, Thailand, 25–28 October 2015. [Google Scholar]
  133. Khazaee, A.; Yazdani, A.; Makhdoumi, A.; Atashin, S.A.; Wu, B. Loss Minimization Algorithm for Surface-Mounted PMSM Using Ripple-Based Extremum Seeking. IEEE Trans. Ind. Electron. 2024, 71, 15325–15335. [Google Scholar] [CrossRef]
  134. Tian, L.; Liu, Y.; Zhao, J.; Sun, J. A novel MTPA control for IPMSMs based on extremum seeking control. In Proceedings of the 2015 34th Chinese Control Conference (CCC), Hangzhou, China, 28–30 July 2015. [Google Scholar]
  135. Zhou, X.; Zhou, Y.; Wang, H.; Lu, M.; Zeng, F.; Yu, Y. An improved MTPA control based on amplitude-adjustable square wave injection. IEEE Trans. Energy Convers. 2020, 35, 956–965. [Google Scholar] [CrossRef]
  136. Li, K.; Wang, Y. Maximum torque per ampere (MTPA) control for IPMSM drives using signal injection and an MTPA control law. IEEE Trans. Ind. Inform. 2019, 15, 5588–5598. [Google Scholar] [CrossRef]
  137. Kim, S.; Yoon, Y.-D.; Sul, S.-K.; Ide, K. Maximum torque per ampere (MTPA) control of an IPM machine based on signal injection considering inductance saturation. IEEE Trans. Power Electron. 2012, 28, 488–497. [Google Scholar] [CrossRef]
  138. Antonello, R.; Carraro, M.; Zigliotto, M. Maximum-torque-per-ampere operation of anisotropic synchronous permanent-magnet motors based on extremum seeking control. IEEE Trans. Ind. Electron. 2013, 61, 5086–5093. [Google Scholar] [CrossRef]
  139. Kim, S.; Yoon, Y.-D.; Sul, S.-K.; Ide, K.; Tomita, K. Parameter independent maximum torque per ampere (MTPA) control of IPM machine based on signal injection. In Proceedings of the 2010 Twenty-Fifth Annual IEEE Applied Power Electronics Conference and Exposition (APEC), Palm Springs, CA, USA, 21–25 February 2010. [Google Scholar]
  140. Zhang, W.; Xiao, F.; Liu, J.; Mai, Z.; Li, C. Maximum torque per ampere control for IPMSM traction system based on current angle signal injection method. J. Electr. Eng. Technol. 2020, 15, 1681–1691. [Google Scholar] [CrossRef]
  141. Lin, F.-J.; Liu, Y.-T.; Yu, W.-A. Power perturbation based MTPA with an online tuning speed controller for an IPMSM drive system. IEEE Trans. Ind. Electron. 2017, 65, 3677–3687. [Google Scholar] [CrossRef]
  142. Bolognani, S.; Sgarbossa, L.; Zordan, M. Self-tuning of MTPA current vector generation scheme in IPM synchronous motor drives. In Proceedings of the 2007 European Conference on Power Electronics and Applications, Aalborg, Denmark, 2–5 September 2007. [Google Scholar]
  143. Bedetti, N.; Calligaro, S.; Olsen, C.; Petrella, R. Automatic MTPA tracking in IPMSM drives: Loop dynamics, design, and auto-tuning. IEEE Trans. Ind. Appl. 2017, 53, 4547–4558. [Google Scholar] [CrossRef]
  144. Sun, T.; Wang, J.; Chen, X. Maximum torque per ampere (MTPA) control for interior permanent magnet synchronous machine drives based on virtual signal injection. IEEE Trans. Power Electron. 2014, 30, 5036–5045. [Google Scholar] [CrossRef]
  145. Sun, T.; Wang, J. Extension of virtual-signal-injection-based MTPA control for interior permanent-magnet synchronous machine drives into the field-weakening region. IEEE Trans. Ind. Electron. 2015, 62, 6809–6817. [Google Scholar] [CrossRef]
  146. Sun, T.; Wang, J.; Koc, M. On accuracy of virtual signal injection based MTPA operation of interior permanent magnet synchronous machine drives. IEEE Trans. Power Electron. 2016, 32, 7405–7408. [Google Scholar] [CrossRef]
  147. Sun, T.; Wang, J.; Koc, M.; Chen, X. Self-learning MTPA control of interior permanent-magnet synchronous machine drives based on virtual signal injection. IEEE Trans. Ind. Appl. 2016, 52, 3062–3070. [Google Scholar] [CrossRef]
  148. Zhao, Y. Online MTPA control for salient-pole PMSMs using square-wave current injection. In Proceedings of the 2016 IEEE Energy Conversion Congress and Exposition (ECCE), Milwaukee, WI, USA, 18–22 September 2016. [Google Scholar]
  149. Sun, T.; Koç, M.; Wang, J. MTPA control of IPMSM drives based on virtual signal injection considering machine parameter variations. IEEE Trans. Ind. Electron. 2017, 65, 6089–6098. [Google Scholar] [CrossRef]
  150. Han, Z.; Liu, J.; Yang, W.; Pinhal, D.B.; Reiland, N.; Gerling, D. Improved online maximum-torque-per-ampere algorithm for speed controlled interior permanent magnet synchronous machine. IEEE Trans. Ind. Electron. 2019, 67, 3398–3408. [Google Scholar] [CrossRef]
  151. Wang, J.; Huang, X.; Yu, D.; Chen, Y.; Zhang, J.; Niu, F.; Fang, Y.; Cao, W.; Zhang, H. An accurate virtual signal injection control of MTPA for an IPMSM with fast dynamic response. IEEE Trans. Power Electron. 2017, 33, 7916–7926. [Google Scholar] [CrossRef]
  152. Zhao, Y. Virtual square-wave current injection based maximum torque per ampere control for interior permanent-magnet synchronous machines. In Proceedings of the 2016 IEEE Transportation Electrification Conference and Expo (ITEC), Dearborn, MI, USA, 27–29 June 2016. [Google Scholar]
  153. Mahmud, M.H.; Wu, Y.; Zhao, Y. Extremum seeking-based optimum reference flux searching for direct torque control of interior permanent magnet synchronous motors. IEEE Trans. Transp. Electrif. 2019, 6, 41–51. [Google Scholar] [CrossRef]
  154. Tahami, F.; Nademi, H.; Rezaei, M. A high-performance vector-controlled PMSM drive with maximum torque per ampere operation. In Proceedings of the 2008 IEEE 2nd International Power and Energy Conference, Johor Bahru, Malaysia, 1–3 December 2008. [Google Scholar]
  155. Hoque, M.A. Speed and Position Sensorless Permanent Magnet Synchronous Motor Drives; Memorial University of Newfoundland: St. John’s, NL, Canada, 1993. [Google Scholar]
  156. Benjak, O.; Gerling, D. Review of position estimation methods for IPMSM drives without a position sensor part I: Nonadaptive methods. In Proceedings of the XIX International Conference on Electrical Machines—ICEM 2010, Rome, Italy, 6–8 September 2010. [Google Scholar]
  157. Zhao, Y.; Wei, C.; Zhang, Z.; Qiao, W. A review on position/speed sensorless control for permanent-magnet synchronous machine-based wind energy conversion systems. IEEE J. Emerg. Sel. Top. Power Electron. 2013, 1, 203–216. [Google Scholar] [CrossRef]
  158. Singh, S.; Singh, S.; Tiwari, A. PMSM drives and its application: An overview. Recent Adv. Electr. Electron. Eng. (Former. Recent Pat. Electr. Electron. Eng.) 2023, 16, 4–16. [Google Scholar] [CrossRef]
  159. Kulkarni, A.B.; Ehsani, M. A novel position sensor elimination technique for the interior permanent-magnet synchronous motor drive. IEEE Trans. Ind. Appl. 1992, 28, 144–150. [Google Scholar] [CrossRef]
  160. Finch, J.W.; Giaouris, D. Controlled AC electrical drives. IEEE Trans. Ind. Electron. 2008, 55, 481–491. [Google Scholar] [CrossRef]
  161. Yongdong, L.; Hao, Z. Sensorless control of permanent magnet synchronous motor—A survey. In Proceedings of the 2008 IEEE Vehicle Power and Propulsion Conference, Harbin, China, 3–5 September 2008. [Google Scholar]
  162. Yeh, H.-C.; Yang, S.-M. Phase inductance and rotor position estimation for sensorless permanent magnet synchronous machine drives at standstill. IEEE Access 2021, 9, 32897–32907. [Google Scholar] [CrossRef]
  163. Park, G.; Kim, G.; Gu, B.-G. Sensorless PMSM Drive Inductance Estimation Based on a Data-Driven Approach. Electronics 2021, 10, 791. [Google Scholar] [CrossRef]
  164. Ye, S.; Yao, X. A modified flux sliding-mode observer for the sensorless control of PMSMs with online stator resistance and inductance estimation. IEEE Trans. Power Electron. 2020, 35, 8652–8662. [Google Scholar] [CrossRef]
  165. Naidu, M.; Bose, B.K. Rotor position estimation scheme of a permanent magnet synchronous machine for high performance variable speed drive. In Proceedings of the Conference Record of the 1992 IEEE Industry Applications Society Annual Meeting, Houston, TX, USA, 4–9 October 1992. [Google Scholar]
  166. Ge, Y.; Yang, L.; Ma, X. Sensorless control of PMSM using generalized extended state observer and adaptive resistance estimation. IET Electr. Power Appl. 2020, 14, 2062–2073. [Google Scholar] [CrossRef]
  167. Zhang, Z. Sensorless back EMF based control of synchronous PM and reluctance motor drives—A review. IEEE Trans. Power Electron. 2022, 37, 10290–10305. [Google Scholar] [CrossRef]
  168. Zhao, Z.; Hu, C.; Wang, Z.; Wu, S.; Liu, Z.; Zhu, Y. Back EMF-based dynamic position estimation in the whole speed range for precision sensorless control of PMLSM. IEEE Trans. Ind. Inform. 2022, 19, 6525–6536. [Google Scholar] [CrossRef]
  169. Wang, T.; Huang, J.; Ye, M.; Chen, J.; Kong, W.; Kang, M.; Yu, M. An EMF observer for PMSM sensorless drives adaptive to stator resistance and rotor flux linkage. IEEE J. Emerg. Sel. Top. Power Electron. 2018, 7, 1899–1913. [Google Scholar] [CrossRef]
  170. Zhang, Z.; Feng, J. Sensorless control of salient PMSM with EKF of speed and rotor position. In Proceedings of the 2008 International Conference on Electrical Machines and Systems, Wuhan, China, 17–20 October 2008. [Google Scholar]
  171. Vaclavek, P.; Blaha, P. PMSM position estimation algorithm design based on the estimate stability analysis. In Proceedings of the 2009 International Conference on Electrical Machines and Systems, Tokyo, Japan, 15–18 November 2009. [Google Scholar]
  172. Nicola, M.; Nicola, C.-I. Sensorless control of PMSM using fractional order SMC and extended Kalman observer. In Proceedings of the 2021 18th International Multi-Conference on Systems, Signals & Devices (SSD), Monastir, Tunisia, 22–25 March 2021. [Google Scholar]
  173. Tondpoor, K.; Saghaiannezhad, S.M.; Rashidi, A. Sensorless control of pmsm using simplified model based on extended kalman filter. In Proceedings of the 2020 11th Power Electronics, Drive Systems, and Technologies Conference (PEDSTC), Tehran, Iran, 4–6 February 2020. [Google Scholar]
  174. Dilys, J.; Stankevič, V.; Łuksza, K. Implementation of extended Kalman filter with optimized execution time for sensorless control of a PMSM using ARM cortex-M3 microcontroller. Energies 2021, 14, 3491. [Google Scholar] [CrossRef]
  175. Niedermayr, P.; Alberti, L.; Bolognani, S.; Abl, R. Implementation and experimental validation of ultrahigh-speed pmsm sensorless control by means of extended kalman filter. IEEE J. Emerg. Sel. Top. Power Electron. 2020, 10, 3337–3344. [Google Scholar] [CrossRef]
  176. Kang, J.; Zeng, X.; Wu, Y.; Hu, D. Study of position sensorless control of PMSM based on MRAS. In Proceedings of the 2009 IEEE International Conference on Industrial Technology, Churchill, VIC, Australia, 10–13 February 2009. [Google Scholar]
  177. Prabhakaran, K.; Karthikeyan, A. Electromagnetic torque-based model reference adaptive system speed estimator for sensorless surface mount permanent magnet synchronous motor drive. IEEE Trans. Ind. Electron. 2020, 67, 5936–5947. [Google Scholar] [CrossRef]
  178. Said, M.A.A.; Sreejeth, M.; Bhattacharya, A. Sensorless control of PMSM using Direct and Model References Adaptive system estimation techniques and observing the effect of using PI Fuzzy logic controller in MRAS Model. In Proceedings of the 2020 IEEE International Conference for Innovation in Technology (INOCON), Bangaluru, India, 6–8 November 2020. [Google Scholar]
  179. Zhang, S.; Shen, A.; Luo, X.; Tang, Q.; Li, Z. Multi-sliding mode current disturbance suppression scheme based model reference adaptive system for sensorless control of permanent magnet synchronous motor. ISA Trans. 2023, 137, 615–628. [Google Scholar] [CrossRef]
  180. Abo-Khalil, A.G.; Eltamaly, A.M.; Alsaud, M.S.; Sayed, K.; Alghamdi, A.S. Sensorless control for PMSM using model reference adaptive system. Int. Trans. Electr. Energy Syst. 2021, 31, e12733. [Google Scholar] [CrossRef]
  181. Tang, Y.; Xu, W.; Liu, Y.; Dong, D. Dynamic performance enhancement method based on improved model reference adaptive system for SPMSM sensorless drives. IEEE Access 2021, 9, 135012–135023. [Google Scholar] [CrossRef]
  182. Ren, N.; Fan, L.; Zhang, Z. Sensorless PMSM control with sliding mode observer based on sigmoid function. J. Electr. Eng. Technol. 2021, 16, 933–939. [Google Scholar] [CrossRef]
  183. Zuo, Y.; Lai, C.; Iyer, K.L.V. A review of sliding mode observer based sensorless control methods for PMSM drive. IEEE Trans. Power Electron. 2023, 38, 11352–11367. [Google Scholar] [CrossRef]
  184. Li, C.; Elbuluk, M. A sliding mode observer for sensorless control of permanent magnet synchronous motors. In Proceedings of the Conference Record of the 2001 IEEE Industry Applications Conference. 36th IAS Annual Meeting (Cat. No. 01CH37248), Chicago, IL, USA, 30 September–4 October 2001. [Google Scholar]
  185. Salvatore, L.; Cupertino, F.; Cascella, G.L. A new approach to sensorless vector control of SPMSM with adaptive sliding-mode observer. In Proceedings of the 2002 IEEE International Symposium on Industrial Electronics (ISIE), L’Aquila, Italy, 8–11 July 2002. [Google Scholar]
  186. Chretien, L.; Husain, I. Position sensorless control of non-salient pmsm from very low speed to high speed for low cost applications. In Proceedings of the 2007 IEEE Industry Applications Annual Meeting, New Orleans, LA, USA, 23–27 September 2007. [Google Scholar]
  187. Lee, H.; Lee, J. Design of iterative sliding mode observer for sensorless PMSM control. IEEE Trans. Control Syst. Technol. 2012, 21, 1394–1399. [Google Scholar] [CrossRef]
  188. Gong, C.; Hu, Y.; Gao, J.; Wang, Y.; Yan, L. An improved delay-suppressed sliding-mode observer for sensorless vector-controlled PMSM. IEEE Trans. Ind. Electron. 2019, 67, 5913–5923. [Google Scholar] [CrossRef]
  189. Xie, G.; Lu, K.; Dwivedi, S.K.; Rosholm, J.R.; Blaabjerg, F. Minimum-voltage vector injection method for sensorless control of PMSM for low-speed operations. IEEE Trans. Power Electron. 2015, 31, 1785–1794. [Google Scholar] [CrossRef]
  190. Jang, J.-H.; Sul, S.-K.; Ha, J.-I.; Ide, K.; Sawamura, M. Sensorless drive of surface-mounted permanent-magnet motor by high-frequency signal injection based on magnetic saliency. IEEE Trans. Ind. Appl. 2003, 39, 1031–1039. [Google Scholar] [CrossRef]
  191. Jang, J.-H.; Ha, J.-I.; Ohto, M.; Ide, K.; Sul, S.-K. Analysis of permanent-magnet machine for sensorless control based on high-frequency signal injection. IEEE Trans. Ind. Appl. 2004, 40, 1595–1604. [Google Scholar] [CrossRef]
  192. Reigosa, D.; gu Kang, Y.; Martínez, M.; Fernández, D.; Guerrero, J.M.; Briz, F. SPMSMs sensorless torque estimation using high-frequency signal injection. IEEE Trans. Ind. Appl. 2020, 56, 2700–2708. [Google Scholar] [CrossRef]
  193. Zhang, Y.; Liu, J.; Zhou, M.; Gao, H. Position sensorless control system of SPMSM based on high frequency signal injection method with passive controller. Int. J. Electr. Hybrid Veh. 2022, 14, 112–133. [Google Scholar] [CrossRef]
  194. Zhao, X.; Wang, C.; Duan, W.; Jiang, J. Research on sensorless control system of low speed and high power PMSM based on improved high frequency signal injection. Energy Rep. 2021, 7, 499–504. [Google Scholar] [CrossRef]
  195. Bi, G.; Zhang, G.; Wang, Q.; Ding, D.; Li, B.; Wang, G.; Xu, D. High-frequency injection angle self-adjustment based online position error suppression method for sensorless PMSM drives. EEE Trans. Power Electron. 2022, 38, 1412–1417. [Google Scholar] [CrossRef]
  196. Eskola, M.; Tuusa, H. Sensorless control of salient pole PMSM using a low-frequency signal injection. In Proceedings of the 2005 European Conference on Power Electronics and Applications, Dresden, Germany, 11–14 September 2005. [Google Scholar]
  197. Xu, D.; Wang, B.; Zhang, G.; Wang, G.; Yu, Y. A review of sensorless control methods for AC motor drives. CES Trans. Electr. Mach. Syst. 2018, 2, 104–115. [Google Scholar] [CrossRef]
  198. Zhao, X. Research Review on Sensorless Measurement Technology of Permanent Magnet Synchronous Motor. Int. J. Mech. Electr. Eng. 2024, 2, 93–102. [Google Scholar] [CrossRef]
  199. Wang, Q.; Wang, S.; Chen, C. Review of sensorless control techniques for PMSM drives. IEEJ Trans. Electr. Electron. Eng. 2019, 14, 1543–1552. [Google Scholar] [CrossRef]
  200. Schrodl, M.; Lambeck, M. Statistic properties of the INFORM method for highly dynamic sensorless control of PM motors down to standstill. In Proceedings of the IECON’03. 29th Annual Conference of the IEEE Industrial Electronics Society (IEEE Cat. No. 03CH37468), Roanoke, VA, USA, 2–6 November 2003. [Google Scholar]
  201. Wang, G.; Ni, R.; Jiang, T.; Gu, S. Position Sensorless Control of PMSMs Using Improved INFORM Method with a Single Current Sensor at Low Speed. In Proceedings of the 2021 24th International Conference on Electrical Machines and Systems (ICEMS), Gyeongju, Republic of Korea, 31 October–3 November 2021. [Google Scholar]
  202. Bui, X.M. Sensorless Control and Fast on-Line Parameter Estimation of IPMSM Based on Current Derivative Measurements; UNSW Sydney: Sydney, Australia, 2019. [Google Scholar]
  203. Bui, M.X.; Le, K.T.; Nguyen, T.T.; Rahman, F. Sensorless control and on-line estimation of inductances of IPMSMs using current derivatives. In Proceedings of the 2019 IEEE Vehicle Power and Propulsion Conference (VPPC), Hanoi, Vietnam, 14–17 October 2019. [Google Scholar]
  204. Isidori, A.; Astolfi, A. Disturbance attenuation and H/sub infinity/-control via measurement feedback in nonlinear systems. IEEE Trans. Autom. Control 1992, 37, 1283–1293. [Google Scholar] [CrossRef]
  205. Van Der Schaft, A. L 2-gain analysis of nonlinear systems and nonlinear state feedback H∞ control. IEEE Trans. Autom. Control 1992, 37, 770–784. [Google Scholar] [CrossRef]
  206. Ponce, I.U.; Bentsman, J.; Orlov, Y.; Aguilar, L.T. Generic Nonsmooth H∞ Output Synthesis: Application to a Coal-Fired Boiler/Turbine Unit With Actuator Dead Zone. IEEE Trans. Control. Syst. Technol. 2015, 23, 2117–2128. [Google Scholar] [CrossRef]
  207. Ramírez-Villalobos, R.; Aguilar, L.T.; Coria, L.N. Sensorless H∞ speed-tracking synthesis for surface-mount permanent magnet synchronous motor. ISA Trans. 2017, 67, 140–150. [Google Scholar] [CrossRef]
  208. Sellschopp, F.; Arjona, M. DC decay test for estimating d-axis synchronous machine parameters: A two-transfer-function approach. IEE Proc.-Electr. Power Appl. 2006, 153, 123–128. [Google Scholar] [CrossRef]
  209. Perez, J.N.H.; Hernandez, O.S.; Caporal, R.M.; Magdaleno, J.d.J.R.; Barreto, H.P. Parameter identification of a permanent magnet synchronous machine based on current decay test and particle swarm optimization. IEEE Lat. Am. Trans. 2013, 11, 1176–1181. [Google Scholar] [CrossRef]
  210. Horning, S.; Keyhani, A.; Kamwa, I. On-line evaluation of a round rotor synchronous machine parameter set estimated from standstill time-domain data. IEEE Trans. Energy Convers. 1997, 12, 289–296. [Google Scholar] [CrossRef]
  211. Zhang, J.; Radun, A. A new method to measure the switched reluctance motor’s flux. IEEE Trans. Ind. Appl. 2006, 42, 1171–1176. [Google Scholar] [CrossRef]
  212. Kilthau, A.; Pacas, J. Parameter-measurement and control of the synchronous reluctance machine including cross saturation. In Proceedings of the Conference Record of the 2001 IEEE Industry Applications Conference. 36th IAS Annual Meeting (Cat. No. 01CH37248), Chicago, IL, USA, 30 September–4 October 2001. [Google Scholar]
  213. Sellschopp, F.S. A tool for extracting synchronous machines parameters from the dc flux decay test. Comput. Electr. Eng. 2005, 31, 56–68. [Google Scholar] [CrossRef]
  214. Hasni, M.; Touhami, O.; Ibtiouen, R.; Fadel, M.; Caux, S. Synchronous machine parameter identification by various excitation signals. Electr. Eng. 2008, 90, 219–228. [Google Scholar] [CrossRef]
  215. Turner, P.; Reece, A.; Macdonald, D. The DC decay test for determining synchronous machine parameters: Measurement and simulation. IEEE Trans. Energy Convers. 1989, 4, 616–623. [Google Scholar] [CrossRef]
  216. Tumageanian, A.; Keyhani, A. Identification of synchronous machine linear parameters from standstill step voltage input data. IEEE Trans. Energy Convers. 1995, 10, 232–240. [Google Scholar] [CrossRef]
  217. Sharma, V.K.; Murthy, S.S.; Singh, B. An improved method for the determination of saturation characteristics of switched reluctance motors. IEEE Trans. Instrum. Meas. 1999, 48, 995–1000. [Google Scholar] [CrossRef]
  218. Kamwa, I.; Viarouge, P.; Dickinson, E. Identification of generalised models of synchronous machines from time-domain tests. IEE Proc. C (Gener. Transm. Distrib.) 1991, 138, 485–498. [Google Scholar] [CrossRef]
  219. Hasni, M.; Touhami, O.; Ibtiouen, R.; Fadel, M.; Caux, S. Estimation of synchronous machine parameters by standstill tests. Math. Comput. Simul. 2010, 81, 277–289. [Google Scholar] [CrossRef]
  220. Boje, E.; Balda, J.; Harley, R.; Beck, R. Time-domain identification of synchronous machine parameters from simple standstill tests. IEEE Trans. Energy Convers. 1990, 5, 164–175. [Google Scholar] [CrossRef] [PubMed]
  221. Groza, V. Experimental determination of synchronous machine reactances from DC decay at standstill. IEEE Trans. Instrum. Meas. 2003, 52, 158–164. [Google Scholar] [CrossRef]
  222. Vicol, L.; Tu Xuan, M.; Wetter, R.; Simond, J.-J.; Viorel, I. On the Identification of the Synchronous Machine Parameters Using Standstill DC Decay Test. In Proceedings of the 2006 ICEM (International Conference on Electrical Machines), Chania, Crete, Greece, 2–5 September 2006. [Google Scholar]
  223. Sandre-Hernandez, O.; Morales-Caporal, R.; Rangel-Magdaleno, J.; Peregrina-Barreto, H.; Hernandez-Perez, J.N. Parameter identification of PMSMs using experimental measurements and a PSO algorithm. IEEE Trans. Instrum. Meas. 2015, 64, 2146–2154. [Google Scholar] [CrossRef]
  224. Boldea, I. Reluctance Synchronous Machines and Drives; Oxford University Press: Oxford, UK, 1996; Volume 38. [Google Scholar]
  225. Gao, Y.; Qu, R.; Chen, Y.; Li, J.; Xu, W. Review of off-line synchronous inductance measurement method for permanent magnet synchronous machines. In Proceedings of the 2014 IEEE Conference and Expo Transportation Electrification Asia-Pacific (ITEC Asia-Pacific), Beijing, China, 31 August–3 September 2014. [Google Scholar]
  226. Gao, Y.; Qu, R.; Liu, Y. An improved AC standstill method for inductance measurement of interior permanent magnet synchronous motors. In Proceedings of the 2013 International Conference on Electrical Machines and Systems (ICEMS), Busan, Republic of Korea, 26–29 October 2013. [Google Scholar]
  227. Balda, J.; Fairbairn, R.; Harley, R.; Rodgerson, J.; Eitelberg, E. Measurement of synchronous machine parameters by a modified frequency response method-Part II: Measured results. IEEE Power Eng. Rev. 1987, EC-2, 4652–4657. [Google Scholar] [CrossRef]
  228. Touhami, O.; Guesbaoni, H.; Iung, C. Synchronous machine parameter identification by a multitime scale technique. IEEE Trans. Ind. Appl. 1994, 30, 1600. [Google Scholar] [CrossRef]
  229. Cavagnino, A.; Lazzari, M.; Profumo, F.; Tenconi, A. Axial flux interior PM synchronous motor: Parameters identification and steady-state performance measurements. In Proceedings of the Conference Record of the 1999 IEEE Industry Applications Conference. Thirty-Forth IAS Annual Meeting (Cat. No. 99CH36370), Phoenix, AZ, USA, 3–7 October 1999. [Google Scholar]
  230. Dedene, N.; Pintelon, R.; Lataire, P. Estimation of a global synchronous machine model using a multiple-input multiple-output estimator. IEEE Trans. Energy Convers. 2003, 18, 11–16. [Google Scholar] [CrossRef]
  231. Senjyu, T.; Urasakt, N.; Simabukuro, T.; Uezato, K. Modelling and parameter measurement of salient-pole permanent magnet synchronous motors including stator iron loss. Math. Comput. Model. Dyn. Syst. 1998, 4, 219–230. [Google Scholar] [CrossRef]
  232. Dutta, R.; Rahman, M. A comparative analysis of two test methods of measuring d- and q-axes inductances of interior permanent-magnet machine. IEEE Trans. Magn. 2006, 42, 3712–3718. [Google Scholar] [CrossRef]
  233. Vandoorn, T.L.; Frederik, M.; Vyncke, T.J.; Melkebeek, J.A.; Lataire, P. Generation of multisinusoidal test signals for the identification of synchronous-machine parameters by using a voltage-source inverter. IEEE Trans. Ind. Electron. 2009, 57, 430–439. [Google Scholar] [CrossRef]
  234. Sun, T.; Kwon, S.-O.; Lee, J.-J.; Hong, J.-P. An improved AC standstill method for testing inductances of interior PM synchronous motor considering cross-magnetizing effect. In Proceedings of the 2009 IEEE Energy Conversion Congress and Exposition, San Jose, CA, USA, 20–24 September 2009. [Google Scholar]
  235. Chiba, A.; Nakamura, F.; Fukao, T.; Rahman, M.A. Inductances of cageless reluctance-synchronous machines having nonsinusoidal space distributions. IEEE Trans. Ind. Appl. 1991, 27, 44–51. [Google Scholar] [CrossRef]
  236. Rahman, K.M.; Hiti, S. Identification of machine parameters of a synchronous motor. IEEE Trans. Ind. Appl. 2005, 41, 557–565. [Google Scholar] [CrossRef]
  237. Nee, H.-P.; Lefevre, L.; Thelin, P.; Soulard, J. Determination of d and q reactances of permanent-magnet synchronous motors without measurements of the rotor position. IEEE Trans. Ind. Appl. 2000, 36, 1330–1335. [Google Scholar] [CrossRef]
  238. Choi, C.; Lee, W.; Kwon, S.-O.; Hong, J.-P. Experimental estimation of inductance for interior permanent magnet synchronous machine considering temperature distribution. IEEE Trans. Magn. 2013, 49, 2990–2996. [Google Scholar] [CrossRef]
  239. Schaible, U.; Szabados, B. Dynamic motor parameter identification for high speed flux weakening operation of brushless permanent magnet synchronous machines. IEEE Trans. Energy Convers. 1999, 14, 486–492. [Google Scholar] [CrossRef]
  240. Stumberger, B.; Kreca, B.; Hribernik, B. Determination of parameters of synchronous motor with permanent magnets from measurement of load conditions. IEEE Trans. Energy Convers. 1999, 14, 1413–1416. [Google Scholar] [CrossRef]
  241. Stumberger, B.; Stumberger, G.; Dolinar, D.; Hamler, A.; Trlep, M. Evaluation of saturation and cross-magnetization effects in interior permanent-magnet synchronous motor. IEEE Trans. Ind. Appl. 2003, 39, 1264–1271. [Google Scholar] [CrossRef]
  242. Lee, J.-Y.; Lee, S.-H.; Lee, G.-H.; Hong, J.-P.; Hur, J. Determination of parameters considering magnetic nonlinearity in an interior permanent magnet synchronous motor. IEEE Trans. Magn. 2006, 42, 1303–1306. [Google Scholar] [CrossRef]
  243. Meessen, K.J.; Thelin, P.; Soulard, J.; Lomonova, E.A. Inductance calculations of permanent-magnet synchronous machines including flux change and self-and cross-saturations. IEEE Trans. Magn. 2008, 44, 2324–2331. [Google Scholar] [CrossRef]
  244. Vaseghi, B.; Takorabet, N.; Meibody-Tabar, F. Fault analysis and parameter identification of permanent-magnet motors by the finite-element method. IEEE Trans. Magn. 2009, 45, 3290–3295. [Google Scholar] [CrossRef]
  245. Kim, W.-H.; Kim, M.-J.; Lee, K.-D.; Lee, J.-J.; Han, J.-H.; Jeong, T.-C.; Cho, S.-Y.; Lee, J. Inductance calculation in IPMSM considering magnetic saturation. IEEE Trans. Magn. 2013, 50, 4001304. [Google Scholar] [CrossRef]
  246. Chen, Y.; Zhu, Z.; Howe, D. Calculation of d-and q-axis inductances of PM brushless ac machines accounting for skew. IEEE Trans. Magn. 2005, 41, 3940–3942. [Google Scholar] [CrossRef]
  247. Fernandez-Bernal, F.; Garcia-Cerrada, A.; Faure, R. Determination of parameters in interior permanent-magnet synchronous motors with iron losses without torque measurement. IEEE Trans. Ind. Appl. 2001, 37, 1265–1272. [Google Scholar] [CrossRef]
  248. Wang, G.; Qu, L.; Zhan, H.; Xu, J.; Ding, L.; Zhang, G.; Xu, D. Self-commissioning of permanent magnet synchronous machine drives at standstill considering inverter nonlinearities. IEEE Trans. Power Electron. 2014, 29, 6615–6627. [Google Scholar] [CrossRef]
  249. Tadokoro, D.; Morimoto, S.; Inoue, Y.; Sanada, M. Method for auto-tuning of current and speed controller in IPMSM drive system based on parameter identification. In Proceedings of the 2014 International Power Electronics Conference (IPEC-Hiroshima 2014-ECCE ASIA), Hiroshima, Japan, 18–21 May 2014. [Google Scholar]
  250. Morimoto, S.; Sanada, M.; Takeda, Y. Mechanical sensorless drives of IPMSM with online parameter identification. IEEE Trans. Ind. Appl. 2006, 42, 1241–1248. [Google Scholar] [CrossRef]
  251. Yoshimi, M.; Hasegawa, M.; Matsui, K. Parameter identification for IPMSM position sensorless control based on unknown input observer. In Proceedings of the 2010 IEEE Symposium on Industrial Electronics and Applications (ISIEA), Penang, Malaysia, 3–5 October 2010. [Google Scholar]
  252. Ichikawa, S.; Tomita, M.; Doki, S.; Okuma, S. Sensorless control of permanent-magnet synchronous motors using online parameter identification based on system identification theory. IEEE Trans. Ind. Electron. 2006, 53, 363–372. [Google Scholar] [CrossRef]
  253. Inoue, Y.; Kawaguchi, Y.; Morimoto, S.; Sanada, M. Performance improvement of sensorless IPMSM drives in a low-speed region using online parameter identification. IEEE Trans. Ind. Appl. 2010, 47, 798–804. [Google Scholar] [CrossRef]
  254. Inoue, Y.; Yamada, K.; Morimoto, S.; Sanada, M. Effectiveness of voltage error compensation and parameter identification for model-based sensorless control of IPMSM. IEEE Trans. Ind. Appl. 2009, 45, 213–221. [Google Scholar] [CrossRef]
  255. Ichikawa, S.; Tomita, M.; Doki, S.; Okuma, S. Sensorless control of synchronous reluctance motors based on extended EMF models considering magnetic saturation with online parameter identification. IEEE Trans. Ind. Appl. 2006, 42, 1264–1274. [Google Scholar] [CrossRef]
  256. Weijie, L.; Dongliang, L.; Qiuxuan, W.; Lili, C.; Xiaodan, Z. A novel deadbeat-direct torque and flux control of IPMSM with parameter identification. In Proceedings of the 2016 18th European Conference on Power Electronics and Applications (EPE’16 ECCE Europe), Karlsruhe, Germany, 5–9 September 2016. [Google Scholar]
  257. Underwood, S.J.; Husain, I. Online parameter estimation and adaptive control of permanent-magnet synchronous machines. IEEE Trans. Ind. Electron. 2009, 57, 2435–2443. [Google Scholar] [CrossRef]
  258. Kim, S.-J.; Lee, H.-W.; Kim, K.-S.; Bae, J.-N.; Im, J.-B.; Kim, C.-J.; Lee, J. Torque ripple improvement for interior permanent magnet synchronous motor considering parameters with magnetic saturation. IEEE Trans. Magn. 2009, 45, 4720–4723. [Google Scholar] [CrossRef]
  259. Xu, Y.; Parspour, N.; Vollmer, U. Torque ripple minimization using online estimation of the stator resistances with consideration of magnetic saturation. IEEE Trans. Ind. Electron. 2013, 61, 5105–5114. [Google Scholar] [CrossRef]
  260. Le, L.; Wilson, W. Synchronous machine parameter identification: A time domain approach. IEEE Trans. Energy Convers. 1988, 3, 241–248. [Google Scholar] [CrossRef]
  261. Mouni, E.; Tnani, S.; Champenois, G. Synchronous generator modelling and parameters estimation using least squares method. Simul. Model. Pract. Theory 2008, 16, 678–689. [Google Scholar] [CrossRef]
  262. Kumar, R.; Gupta, R.; Bansal, A.K. Identification and control of PMSM using artificial neural network. In Proceedings of the 2007 IEEE International Symposium on Industrial Electronics, Vigo, Spain, 4–7 June 2007. [Google Scholar]
  263. Liu, K.; Zhang, Q.; Chen, J.; Zhu, Z.-Q.; Zhang, J. Online multiparameter estimation of nonsalient-pole PM synchronous machines with temperature variation tracking. IEEE Trans. Ind. Electron. 2010, 58, 1776–1788. [Google Scholar] [CrossRef]
  264. Sim, H.-W.; Lee, J.-S.; Lee, K.-B. On-line parameter estimation of interior permanent magnet synchronous motor using an extended Kalman filter. J. Electr. Eng. Technol. 2014, 9, 600–608. [Google Scholar] [CrossRef]
  265. Liu, K.; Zhu, Z.-Q. Online estimation of the rotor flux linkage and voltage-source inverter nonlinearity in permanent magnet synchronous machine drives. IEEE Trans. Power Electron. 2013, 29, 418–427. [Google Scholar] [CrossRef]
  266. Liu, K.; Zhu, Z.-Q.; Stone, D.A. Parameter estimation for condition monitoring of PMSM stator winding and rotor permanent magnets. IEEE Trans. Ind. Electron. 2013, 60, 5902–5913. [Google Scholar] [CrossRef]
  267. Qin, H.; Liu, K.; Zhang, Q.; Shen, A.; Zhang, J. Online estimating voltage source inverter nonlinearity for PMSM by Adaline neural network. In Proceedings of the 2010 IEEE Fifth International Conference on Bio-Inspired Computing: Theories and Applications (BIC-TA), Changsha, China, 23–26 September 2010. [Google Scholar]
  268. Rahman, M.; Hoque, M. On-line adaptive artificial neural network based vector control of permanent magnet synchronous motors. IEEE Trans. Energy Convers. 1998, 13, 311–318. [Google Scholar] [CrossRef]
  269. Liu, T.; Husain, I.; Elbuluk, M. Torque ripple minimization with on-line parameter estimation using neural networks in permanent magnet synchronous motors. In Proceedings of the Conference Record of 1998 IEEE Industry Applications Conference. Thirty-Third IAS Annual Meeting (Cat. No. 98CH36242), St. Louis, MI, USA, 12–15 October 1998. [Google Scholar]
  270. Gatto, G.; Marongiu, I.; Serpi, A. Discrete-time parameter identification of a surface-mounted permanent magnet synchronous machine. IEEE Trans. Ind. Electron. 2012, 60, 4869–4880. [Google Scholar] [CrossRef]
  271. Piippo, A.; Hinkkanen, M.; Luomi, J. Adaptation of motor parameters in sensorless PMSM drives. IEEE Trans. Ind. Appl. 2009, 45, 203–212. [Google Scholar] [CrossRef]
  272. Boileau, T.; Leboeuf, N.; Nahid-Mobarakeh, B.; Meibody-Tabar, F. Online identification of PMSM parameters: Parameter identifiability and estimator comparative study. IEEE Trans. Ind. Appl. 2011, 47, 1944–1957. [Google Scholar] [CrossRef]
  273. An, Q.; Sun, L. On-line parameter identification for vector controlled PMSM drives using adaptive algorithm. In Proceedings of the 2008 IEEE Vehicle Power and Propulsion Conference, Harbin, China, 3–5 September 2008. [Google Scholar]
  274. Shi, Y.; Sun, K.; Huang, L.; Li, Y. Online identification of permanent magnet flux based on extended Kalman filter for IPMSM drive with position sensorless control. IEEE Trans. Ind. Electron. 2011, 59, 4169–4178. [Google Scholar] [CrossRef]
  275. Hamida, M.A.; De Leon, J.; Glumineau, A.; Boisliveau, R. An adaptive interconnected observer for sensorless control of PM synchronous motors with online parameter identification. IEEE Trans. Ind. Electron. 2012, 60, 739–748. [Google Scholar] [CrossRef]
  276. Mohamed, Y.-R.; Lee, T. Adaptive self-tuning MTPA vector controller for IPMSM drive system. IEEE Trans. Energy Convers. 2006, 21, 636–644. [Google Scholar] [CrossRef]
  277. Liu, K.; Zhu, Z.-Q.; Zhang, Q.; Zhang, J. Influence of nonideal voltage measurement on parameter estimation in permanent-magnet synchronous machines. IEEE Trans. Ind. Electron. 2011, 59, 2438–2447. [Google Scholar] [CrossRef]
  278. Hasegawa, M.; Matsui, K. Position sensorless control for interior permanent magnet synchronous motor using adaptive flux observer with inductance identification. IET Electr. Power Appl. 2009, 3, 209–217. [Google Scholar] [CrossRef]
  279. Boileau, T.; Nahid-Mobarakeh, B.; Meibody-Tabar, F. On-line identification of PMSM parameters: Model-reference vs EKF. In Proceedings of the 2008 IEEE Industry Applications Society Annual Meeting, Edmonton, AB, Canada, 5–9 October 2008. [Google Scholar]
  280. Ohnishi, K.; MatSui, N.; Hori, Y. Estimation, identification, and sensorless control in motion control system. Proc. IEEE 1994, 82, 1253–1265. [Google Scholar] [CrossRef]
  281. Liu, L.; Cartes, D. Synchronisation based adaptive parameter identification for permanent magnet synchronous motors. IET Control. Theory Appl. 2007, 1, 1015–1022. [Google Scholar] [CrossRef]
  282. Keerthi, V.D.; Kumar, J. Model reference adaptive control based parameters estimation of permanent magnet synchronous motor drive. J. Electr. Eng. 2013, 2, 267–276. [Google Scholar] [CrossRef][Green Version]
  283. Song, W.; Gang, Y.; Zhi-Jian, Q.; Shuang-shuang, S.; Chao, C. Identification of PMSM based on EKF and elman neural network. In Proceedings of the 2009 IEEE International Conference on Automation and Logistics, Shenyang, China, 5–7 August 2009. [Google Scholar][Green Version]
  284. Bolognani, S.; Tubiana, L.; Zigliotto, M. Extended Kalman filter tuning in sensorless PMSM drives. IEEE Trans. Ind. Appl. 2003, 39, 1741–1747. [Google Scholar] [CrossRef]
  285. Mwasilu, F.; Jung, J.-W. Enhanced fault-tolerant control of interior PMSMs based on an adaptive EKF for EV traction applications. IEEE Trans. Power Electron. 2015, 31, 5746–5758. [Google Scholar] [CrossRef]
  286. Dhaouadi, R.; Mohan, N.; Norum, L. Design and implementation of an extended Kalman filter for the state estimation of a permanent magnet synchronous motor. IEEE Trans. Power Electron. 1991, 6, 491–497. [Google Scholar] [CrossRef]
  287. Rafaq, M.S.; Mwasilu, F.; Kim, J.; Choi, H.H.; Jung, J.-W. Online parameter identification for model-based sensorless control of interior permanent magnet synchronous machine. IEEE Trans. Power Electron. 2016, 32, 4631–4643. [Google Scholar] [CrossRef]
  288. Dang, D.Q.; Rafaq, M.S.; Choi, H.H.; Jung, J.-W. Online parameter estimation technique for adaptive control applications of interior PM synchronous motor drives. IEEE Trans. Ind. Electron. 2015, 63, 1438–1449. [Google Scholar] [CrossRef]
Figure 1. Comparison between PMSMs, SRMs, and IMs.
Figure 1. Comparison between PMSMs, SRMs, and IMs.
Automation 06 00049 g001
Figure 2. PMSM types.
Figure 2. PMSM types.
Automation 06 00049 g002
Figure 3. Current planes of IPMSMs and SPMSMs.
Figure 3. Current planes of IPMSMs and SPMSMs.
Automation 06 00049 g003
Figure 4. Four operating regions of IPMSMs.
Figure 4. Four operating regions of IPMSMs.
Automation 06 00049 g004
Figure 5. MTPA control method classification of IPMSM.
Figure 5. MTPA control method classification of IPMSM.
Automation 06 00049 g005
Table 1. Comparison between IPMSMs and SPMSMs.
Table 1. Comparison between IPMSMs and SPMSMs.
PMSM TypeIPMSMSPMSMRef.
DesignRequire peripheral sleeveNeed a peripheral sleeve in order to protect PMs from centrifugal forces. Because of this, manufacturing SPM rotors is more difficult than IPMs.[26]
LossesMore core lossesReduced core losses as a result of the rotor steel’s greater separation from the air gap and SPMs’ reduced magnetic flux leaking.
Volume of PM materialMore PM material volume is neededCompared to IPMs, less PM material is used.
CostIs superior in costs High cost.[23]
RobustnessLess robustSPM is more robust.
EfficiencyLess efficiencyHigher overall efficiency and cooler.
CoolingLess in coolingCooler.
Speed rangeIf the machine’s saliency is maximized, IPM has an excellent overload capacity over the whole speed rangeThe SPM motor is unable to surpass the continuous power rating, regardless of the applied current overload.[18]
Joule losses vs. speedRequires a suitable number of stator slots and rotor segments to keep the harmonic losses in control, which can increase the fabrication cost, and has larger Joule losses at low speed because of end connectionsExtra-Joule losses for deexciting the PM flux at high speeds and PM losses that necessitate segmentation in both directions (axial and circumferential) have an impact on the SPM motor.
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

Elkholy, M.M.; Algendy, M.M.; El-Hay, E.A. Modern Control Techniques and Operational Challenges in Permanent Magnet Synchronous Motors: A Comprehensive Review. Automation 2025, 6, 49. https://doi.org/10.3390/automation6040049

AMA Style

Elkholy MM, Algendy MM, El-Hay EA. Modern Control Techniques and Operational Challenges in Permanent Magnet Synchronous Motors: A Comprehensive Review. Automation. 2025; 6(4):49. https://doi.org/10.3390/automation6040049

Chicago/Turabian Style

Elkholy, Mahmoud M., Mohamed M. Algendy, and Enas A. El-Hay. 2025. "Modern Control Techniques and Operational Challenges in Permanent Magnet Synchronous Motors: A Comprehensive Review" Automation 6, no. 4: 49. https://doi.org/10.3390/automation6040049

APA Style

Elkholy, M. M., Algendy, M. M., & El-Hay, E. A. (2025). Modern Control Techniques and Operational Challenges in Permanent Magnet Synchronous Motors: A Comprehensive Review. Automation, 6(4), 49. https://doi.org/10.3390/automation6040049

Article Metrics

Back to TopTop