Next Article in Journal
Non-Weighted Two-Stage Model Predictive Control Strategy Based on Three-Level NPC Inverter
Next Article in Special Issue
Smart Transmission Expansion Planning Based on the System Requirements: A Comparative Study with Unconventional Lines
Previous Article in Journal
Short-Term Electric Load Forecasting Based on Signal Decomposition and Improved TCN Algorithm
Previous Article in Special Issue
Modeling Time-Evolving Electrical Conductivity in Air Ionization Plasma under DC Voltage: A Finite-Difference Time-Domain Approach for Needle-Plate Setup Based on Laboratory Experiments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Self-Oscillating Converter Based on Phase Tracking Closed Loop for a Dynamic IPT System

1
Xiamen Kehua Digital Energy Tech Co., Ltd., Xiamen 361000, China
2
Department of Instrumental and Electrical Engineering, Xiamen University, Xiamen 361005, China
3
School of Automative and Mechanical Engineering, Xiamen University of Technology, Xiamen 361005, China
*
Authors to whom correspondence should be addressed.
Energies 2024, 17(8), 1814; https://doi.org/10.3390/en17081814
Submission received: 2 March 2024 / Revised: 3 April 2024 / Accepted: 5 April 2024 / Published: 10 April 2024
(This article belongs to the Special Issue Energy, Electrical and Power Engineering 2024)

Abstract

:
The coupling of converters with resonant networks poses significant challenges for frequency tracking and power control in inductive power transfer (IPT) systems. This paper presents an implementation method that addresses these issues by dividing the system’s operation into two distinct states: self-oscillating and power-injecting. Based on these states, a phase-closed loop is constructed. Within this closed loop, the phase tracking unit detects and tracks frequency drift, while the power regulating unit incorporates an integrator and adopts a control variable to adjust the output power by modifying the duration of the power injecting state. Meanwhile, the oscillating unit operates in the self-oscillating state. Operating in this manner, the system achieves self-oscillation and demonstrates the capability to effectively track and compensate for system variations within a single cycle. A verification prototype has been constructed, and it demonstrates that the converter within it completely decoupled from the resonant network. Experimental results validate that altering the control variable solely affects the duration of the power-injecting state, allowing for independent control of the output power. When the control variable changes from 2.0 V to 3.5 V, the output power changes from 178 W to 519 W while the self-oscillating state remains unchanged. Furthermore, the system accurately tracks frequency changes, even under significant variations in the coupling coefficient or load, without compromising the power injection state. When the air gap changes from 3 cm to 12 cm, the duration of the self-oscillating state changes from 22.1 μs to 26.3 μs, while the power injecting state remains unchanged. This approach exhibits a robust performance, particularly suitable for dynamic IPT systems sensitive to parameter variations.

1. Introduction

Inductive power transfer (IPT) is a technology that enables power transmission through a magnetic flux without physical contact. It has found widespread applications in various fields, including electric vehicles, mobile phones, and medical transplantation [1,2,3]. However, IPT systems are classified as loosely coupled systems compared to tightly coupled systems. In an IPT system, the primary and secondary coils form a loosely coupled transformer, resulting in poor magnetic coupling between the two sides. This leads to power reduction, decreased efficiency, and increased VA capacity of the converter.
To address this issue, extensive research has been conducted. Mainstream solutions involve incorporating a resonant network (compensation network) into a series between the converter and primary circuit, as well as the secondary circuits and load circuit [3,4,5]. The resonant network primarily consists of reactive elements, with capacitors being the main component. Resonant networks are categorized into nine basic types based on the series or parallel connection of capacitors [6], with four (S-S, S-P, P-S and P-P) being extensively studied and considered to be mainstream topologies [7,8,9].
However, the introduction of a resonant network in an IPT system introduces coupling and mutual interference between the converter and the resonant network during operation. To ensure the stability of the IPT system, it is necessary for the converter frequency to align with the resonant network frequency, thus ensuring a zero phase angle impedance (ZPA) in the primary circuit [10]. Additionally, IPT systems experience the problem of frequency bifurcation [11]. Due to the possibility of misalignment, an uncertainty between the loosely coupled transformer and resonant network occurs, which leads to frequency shifts and bifurcation [12,13], therefore, frequency tracking is necessary.
One common approach is to insert a controlled impedance network on the primary and secondary sides to match any changes that may occur [14,15,16,17]. These controllable impedance networks typically employ self-switching reactive elements, such as self-switching capacitors [1,14]. However, this tracking method has a disadvantage in that it cannot cover the entire frequency range due to the limited number of switches available. Another approach is to utilize a phase-locked loop (PLL) to construct a frequency loop for accurate frequency tracking [18,19,20]. However, the drawback of PLL is that it introduces circuit complexity and may have tracking delay.
The self-oscillating technique is widely employed in various applications, including induction heating, and has also been used in IPT system. This technique utilizes a sensor and sampling circuit to track the zero crossing of the primary current. The switches are then triggered at the zero crossing point to generate self-excited oscillation. By employing this method, the converter and resonant network frequencies can be synchronized, ensuring accurate frequency tracking [21,22,23,24,25]. Nevertheless, in these self-oscillation applications, it is crucial for the converter frequency to align with the resonant network frequency. Consequently, some coupling between the converter and the resonant network still exists, requiring the operation of switches at the zero crossing point. This, in turn, amplifies the complexity of the control strategy. Furthermore, due to the inherent delay introduced by the sensor, sampling and shaping circuits, the actual switching point may deviate from the zero crossing, resulting in the loss of soft switching opportunities.
In practice applications, it is necessary to regulate the voltage and power to achieve the desired output. Conventional solutions involve using a DC/DC converter in either the primary or secondary sides to control the bus voltage and load impedance [26,27] or regulating phase shift to control power flow [28,29]. However, inserting a DC/DC converter into the system introduces additional losses, which can result in decreased overall efficiency. On the other hand, the phase-shifting technology may cause the switching point to deviate from the zero crossing, leading to the loss of the soft switching condition.
In the literature [30,31], a quasi-resonant control strategy is presented that divides an energy transfer period into two states: the resonant state and the energy-independent injection state. This control strategy provides an effective scheme for power control. During the resonant state, the converter is isolated from the resonant network, allowing the system to resonate freely. This resonant state lasts for less than one cycle, enabling frequency tracking during this period. In the energy injection state, the resonant network is isolated from the converter, and the power source directly injects energy into the primary inductor. This decouples the converter from the resonant network, allowing for the independent regulation of power by adjusting the duration of the energy injection state. The limitation of the literature [30,31] lies in the complexity of the system control strategy and the absence of a closed-loop scheme.
In this paper, the objective is to find a way to eliminate the coupling between the converter and the resonant network, allowing them to operate independently, and reducing the problems caused by coupling. To accomplish this, this paper proposes a variable structure approach that divides a converter into three fundamental operating states: switch-on state, blocking state, and diode-on state, based on their functions within an IPT system.
In these three states, the blocking state serves to achieve quasi-resonance within the system and is employed for tracking changes in system frequency. The switch-on state is utilized to inject power and adjust the output power. The diode state maintains soft switching conditions for the converter when transitioning between the blocked and switched states.
Furthermore, these three states can be combined into two distinct states: quasi-resonant and power injection states, which are independent of each other.
Building upon this foundation, this paper introduces a phase-closed loop control topology based on the quasi-resonant and power injection modes. In this closed-loop system, the phase of the quasi-resonant state acts as a feedback signal, triggering the self-oscillation of the system. This means that the system can achieve self-sustained oscillation without the need for external control. Additionally, within this closed-loop, a control variable is employed to regulate the duration of the power injection mode for power adjustment. To facilitate clear understanding, we refer to the quasi-resonant state as the self-oscillating state, while the power injection state retains its original designation.
For a better understanding, Table 1 briefly compares the coupling between existing converters and resonant networks, as well as the power regulation techniques.
Adopting a closed-loop control strategy in prior studies has essentially solved the problem of frequency tracking. However, the coupling between the converter and the resonant network persists, requiring switching points at current zero crossings, which complicates system control. Additionally, power regulation in prior studies still relies on techniques like phase shifting or adding DC/DC converters, whose drawbacks have been discussed extensively in much of the literature.
In document [31], a strategy of time-sharing energy injection and free resonance is adopted based on the quasi-resonant mode to address the coupling and power regulation issues. However, it falls short in establishing closed-loop control and providing a comprehensive solution.
To address these shortcomings, this paper proposes dividing the converter’s operation into three modes: blocking, diode conduction, and switch conduction. These modes allow for dividing the energy transmission period into two states: free resonance and power injection. This approach achieves decoupling between the converter and the resonant network and enables power adjustment by controlling the energy injection time. With closed-loop control, it exhibits favorable dynamic characteristics and power adjustment capabilities.

2. Variable Structure Analysis of an IPT System

Figure 1a depicts an IPT charging system with a converter in the primary side and a resonant network (compensation network) on both the primary and secondary sides. In Figure 1a, the UDC represents the DC source that supplies power to the system, and RL stands for load. Cp and Lp constitute the primary resonant network, while Cs and Ls constitute the secondary resonant network. Mutual inductance, M, establishes a magnetic link between the primary and the secondary to transmit power. Furthermore, Lp and Ls together form a loosely coupled transformer. It is evident that this IPT system belongs to P–P compensation topology.
It is widely recognized that the converter can be constructed using various topologies such as full-bridge, half-bridge, or single switch configurations. These topologies employ semiconductor switches, with each switch being equipped with a body diode. Consequently, during the operation of an IPT system, the converter can be categorized into three states based on its working structure: the switch-on state, the blocking state, and the diode-on state.
To visually illustrate these three states, Figure 1b–d employ a single switch converter as an example. In this context, S represents a semiconductor device, such as an IGBT, while D signifies the body diode associated with S.
Figure 1b illustrates the behavior of the switch-on state of a converter. During this state, the switch S is activated to deliver UDC to Lp, generating a primary inductor current ip with the slope of KE, injecting power into the system, and the value of KE can be determined as follows:
K E = U DC L p
In the switch-on state, due to the connection of UDC to Lp, the voltage across Cp is clamped at UDC, and there is current flowing into the secondary coil.
Figure 1c depicts the behavior of the blocking state, where both switches and diodes are deactivated, leading to the isolation of UDC from the resonant network. During this state, the resonant network initiates oscillation with an angular frequency ω. This angular frequency is solely determined by the system parameters, such as Cp, Lp and secondary reflected impedance, as the resonant network is no longer influenced by the converter.
It is important to note that Cp is connected in parallel with the output of the converter, ensuring that the voltage across Cp does not exceed UDC. As a result, the oscillation of the system in the blocking state is incomplete.
Figure 1d illustrates the behavior of the diode-on state. In this state, the diode D is activated, allowing UDC to be reconnected to Lp. These activated diodes provide a pathway for the current ip to flow back to UDC. Similar to the switch-on state, the voltage across Cp remains clamped at UDC during the diode-on state.
While we employ a single-switch converter in the analysis of these three states, in fact, it is worth noting that these three states are also applicable to the half-bridge and full-bridge converter mentioned earlier.
If the loosely coupled transform in Figure 1 is replaced by the T model, the resonant network can be represented by the equivalent circuit shown in Figure 2. In this circuit, Lpk and Lsk denote the leakage inductances of the primary and secondary, respectively, while LM represents the mutual inductance. The relationship between these parameters can be expressed as follows:
{ L M = M L pk = L p M L sk = L s M k = M L p L s
In Equation (2), k represents the coupling coefficient, which indicates the efficiency of energy transferring from the primary to secondary side. For an IPT system, the coupling coefficient k is typically weak. As a result, only a portion of the power can be transmitted to the secondary side, and a significant amount of energy remains stored in leakage inductance Lpk. This energy needs to be returned to UDC during a power transfer period.
Based on the characteristics of the resonant network in Figure 2, the operational reality can be described as follows: during the switching on state, UDC injects power into the primary inductor Lp; during the blocking state, the injected power excites the resonant network, causing it to oscillate; and during the diode-on state, the residual energy stored in leakage inductance Lpk is returned to UDC.
In Figure 2, capacitor Cp, inductors Lpk, and LM form a compensation topology. In this compensation network, the power injected into the mutual inductance LM is transmitted to the secondary side. Meanwhile, the power remaining in leakage inductance Lpk is utilized to sustain the diode-on state and subsequently returned to the power source UDC.
A characteristic of the compensation network is its ability to harness the energy stored in the leakage inductance to sustain the diode-on state. This bestows a substantial soft-switching margin upon the converter. Consequently, the diode-on state can be employed to establish an isolated transition zone between the blocking state and the switch-on state. Moreover, by designing a repeating converter operational cycle as follows: blocking state → diode-on state → switch-on state → blocking state, and by using some hardware circuits, we can construct a self-oscillating closed loop for an IPT system, as shown in Figure 3. With the diode-on state acting as a buffer, the system designed by this method has a wide soft switching margin.
In this self-oscillating loop, the blocking state facilitates free oscillation, the switch-on state implements power injection, and the diode-on state contributes to the soft switch behavior. By employing this approach, the decoupling between the converter and the resonant network can be achieved while also realizing self-oscillation of the system.
Figure 4 provides a vector trajectory diagram and curves of this self-oscillating loop, demonstrating the dynamic behavior of ip, up. To illustrate the relationship between the switch S and ip, up, the control signal of S, vg, is also depicted in Figure 4.
In Figure 4, the left side displays the vector trajectory diagram. The dotted circle represents the oscillating vector trajectory, while the solid black line represents the actual running vector trajectory of ip, up. Assuming that the system has achieved stability in the nth period, it can be seen that the vector trajectory forms a closed curve divided into three segments: τ1, τ2, and τ3, respectively. tn0 denotes the starting point of the switch-on state, tn1 is the starting point of blocking state, and tn2 is the starting point of the diode-on state.
Referring to Figure 1 and Figure 4, during the τ1 segment, [tn0, tn1], the vector trajectory is in the switch-on state, with up clamped at UDC. As a result, the vector trajectory in this segment is a straight line parallel to the ip axis, where up = UDC. It starts from ip = 0 and extends to the oscillating circle at tn1. The phase angle occupied by τ1 is represented by βτ1.
During the τ2 segment, [tn1, tn2], the system enters the blocking state, and the oscillation begins. However, due to attenuation, the vector trajectory deviates from the oscillating circle and follows a decaying circular curve. The phase angle occupied by τ2 is represented by βτ2.
During the τ3 segment, [tn2, t(n+1)0], the system is in the diode-on state, similar to τ1, with up clamped at UDC. Consequently, the vector trajectory in this segment appears as a parallel line parallel to axis ip, where up = UDC. The phase angle occupied by τ3 is represented by βτ3.
In Figure 4, the right side presents the curves of ip, up, and the relationship between vector trajectory and these curves is indicated by dot-dash lines. Referring to Figure 1, the following is an analysis of ip, up curves in different states:
[tn0, tn1]: This interval represents the switch-on state, τ1. The control signal vg has been at a high level, indicating that switch S has been turned on and UDC supplied to Lp. The current ip increases linearly from zero with slope KE, while up remains clamped at UDC. At tn1, when ip reaches ip (tn1), vg is pulled to a low level, turning off switch S, and the system exits the switch-on state. As the voltage of Cp is equal to UDC, switch S is turned off with the zero voltage switch (ZVS) condition.
[tn1, tn2]: This interval represents the blocking state. During this interval, the system experiences oscillation at the angular frequency ω, causing in-sinusoidal variations in both up and ip. Since the phase angle occupied by oscillation segment, βτ2, is greater than π but less than 2π, the ip exhibits two amplitudes, Ipm1, Ipm2, within this interval. Between these two amplitudes, ip passes through a zero crossing.
After tn1, the capacitor voltage up starts to decrease from its maximum value, UDC. It then reaches its lowest point at the zero crossing of the current, ip. Subsequently, as up begins to rise again, it reaches UDC once more at tn2. When up reaches UDC, diode D is turned on, clamping up to UDC again. Consequently, the system loses the oscillation condition and exit the blocking state at tn2. At tn2, the current ip reaches value of ip (tn2).
[tn2, t(n+1)0]: This interval represents the diode-on state. After tn2, the current ip increase linearly from ip (tn2) with a slope of kE, indicating that the energy stored in the leakage inductor Lpk flows back to power source during this interval. Assuming that the switch S has been turned on before ip reaches its zero crossing, as shown in Figure 4, the system will automatically transition from the diode-on state to the switch-on state at the zero crossing of ip, denoted as point t(n+1)0. Therefore, after t(n+1)0, the system will exit from the diode-on state and enter the switch-on state again, initiating a new power injection process.
Although the converter mentioned above is divided into three states, Figure 4 shows that the vector trajectories of the switch-on state and the diode-on state are the same line segment parallel to ip axis. Additionally, as depicted in Figure 1, in both states, UDC is directly connected to Lp, and the voltage of Cp is clamped at UDC. That is to say, the function of both states is to maintain the power flow in or out of Lp. Therefore, these two states can be collectively referred to as power injecting states, and the durations τ1 and τ3 can be combined into τinj = τ1 + τ3. Similarly, the blocking state can also be referred to as a self-oscillating state, and the duration τ2 can be denoted as τosc.
The value of τosc depends on the phase angle βτ2 and the angular frequency ω, which are determined by real-time parameters. Therefore, frequency tracking can be achieved by monitoring τosc and implementing appropriate compensation. On the other hand, the power injected from UDC to Lp is determined by the duration of power injection state, τinj. That is to say, τinj can be used as a control variable to regulate the system power.
Figure 5 is utilized to illustrate the impact of parameter changes on τosc and τinj during the operation of the proposed IPT system. When parameters are modified, the system follows a new vector trajectory, resulting in corresponding alterations in the curves of ip, up. In Figure 5, the trajectories and curves with unchanged parameters are represented in green, while those after the changes are depicted in red. By comparing the green and red voltage curve, it can be observed that changes in system parameters result in a shift in the start time of the oscillation state from tn1 to tn1’, a shift in the end time from tn2 to tn2’, a change in the angular frequency from ω to ω’ and a change in the phase angle from βτ2 to βτ2’. These changes cause a shift in the duration of the oscillating state from τosc to τosc’.

3. Mathematical Model

Figure 6 illustrates a schematic diagram of the power injecting model.
On the primary side, the UDC directly supplies to Lp. Taking into account the copper resistance of the coil, Rps and Rss denote the loss resistances of the primary and secondary inductors, respectively. Lp and Ls act as power transfer channels, creating a loosely coupled transformer between the primary and secondary sides. The symbol M represents the mutual inductance of the transformer. Additionally, the controlled voltage M × dip/dt signifies the impact of the secondary current on the primary side, while M × dis/dt represents the effect of the primary current on the secondary side.
Figure 7 depicts the schematic diagram of the oscillating state model. In this model, the power source UDC illustrated in Figure 6 is replaced with the capacitor Cp, while the remaining components remain unchanged as depicted in Figure 6.
If a power electronic system exhibits multiple structures periodically during each stabilization period, these structures are determined by the actual state of the circuit. Meanwhile, these structures possess their own linear dynamics, and the system can be modeled using a general state space representation [32]:
{ x = A i x + B i u i = osc , inj y = C x  
where, A is the characteristic matrix, B is the input matrix, C is the output matrix and i represents the serial number of each structures. Respectively, x represents the state variable, with x = [up, ip, is, u0]T; u denotes the input variable, with u = [UDC]; and y represents the output variable.
Based on Figure 6 and Figure 7, the characteristic matrix and input matrix of the power-injecting and self-oscillating states can be written as Equations (4) and (5).
A inj = [ 1 0 0 0 0 L s R ps Δ M R ss Δ M Δ 0 M R ps Δ L p R ss Δ L p Δ 0 0 1 C s 1 C s R L ] B inj = [ 1 L s Δ M Δ 0 ] T
A osc = [ 0 1 C p 0 0 L s Δ L s R ps Δ M R ss Δ M Δ M Δ M R ps Δ L p R ss Δ L p Δ 0 0 1 C s 1 C s R L ] B osc = [ 0 0 0 0 ] T
In the time domain, when the variable t is used as the time variable for each state running process, then, the solution of Equation (3) represents the time function that describes the system behavior [33]:
x ( t ) = Φ i ( t t 0 ) x 0 + A i 1 ( Φ i ( t t 0 ) I ) B i u i = osc , inj
where, x0 represents the initial value at the beginning of each state at time t0, with x0 = x(t0). I denotes the identity matrix. Φ(t) is defined as Φ(t) = exp{Ait}, i = inj, osc.
Referring back to Figure 5 to discuss the contents of the green line, in the proposed IPT system, as described earlier, there exist two states durations, namely τosc and τinj, respectively. Therefore, as depicted in Figure 5, a power transfer period T can be defined as follows:
T = τ osc + τ inj
Let us begin by analyzing the interval [tn1, tn2], which represents the oscillating state duration. According to Equation (5), in this interval, the input matrix B is a zero matrix. Therefore, Equation (6) can be rewritten as follows:
x ( t ) = Φ osc ( t t n 1 ) x n 1 t n 1 t < t n 2
Then, let us analyze the interval [tn2, t(n+1)1], which represents the power injecting state duration. According to Equation (4), Equation (6) can be rewritten as follows:
x ( t ) = Φ inj ( t t n 2 ) x n 2 + A inj 1 ( Φ inj ( t t n 2 ) I ) B inj u t n 2 t ( n + 1 ) 1
Referring to Figure 5, by substituting τosc and τinj into Equations (8) and (9), the state variables xn1 and xn2 at the end of the power injecting state and the oscillating state can be obtained, respectively. xn1 and xn2 can be derived as follows:
{ x n 1 = Φ inj ( τ inj ) x n 2 + A inj 1 ( Φ inj ( τ inj ) I ) B inj u t n 2 t ( n + 1 ) 1 x n 2 = Φ osc ( τ osc ) x n 1 t n 1 t < t n 2
By combining the two equations in Equation (10), xn1 can be expressed as follows:
x n 1 = ( I Φ inj ( τ inj ) Φ osc ( τ osc ) ) 1 A inj 1 ( Φ inj ( τ inj ) I ) B inj u
Referring to the vector trajectory shown in Figure 4, if we assume that x has already stabilized at tn1 with a value of xn1, then, after a duration of one period T, x will repeatedly return to xn1 at t(n+1)1. In other words, xn1 will be equal to x(n+1)1, indicating that xn1 is a fixed point. By applying the fixed point mapping calculation, and considering the end-time point of power injection state tn1, the capacitor voltage of up is clamped at UDC. Therefore, the end value of the power injecting stat can be obtained as C × xn1 = UDC, where the output matrix C = [1,0,0,0]. As a result, Equation (11) can be modified as follows:
{ U DC = C ( I Φ inj ( τ inj ) Φ osc ( τ osc ) ) 1 A inj 1 ( Φ inj ( τ inj ) I ) B inj U DC C = [ 1 0 0 0 ]
Taking a power transfer period T as a known variable and Substituting τinj = Tτosc into Equation (12), Equation (12) can be modified as:
{ U DC = C ( I Φ inj ( T τ osc ) Φ osc ( τ osc ) ) 1 A inj 1 ( Φ inj ( T τ osc ) I ) B inj U DC C = [ 1 0 0 0 ]
After determining the value of T, the oscillating and injecting durations, τosc and τinj, can be obtained by solving Equation (13). The expressions for τosc and τinj as functions of T can be written as follows:
{ τ osc = f ( T ) τ inj = T f ( T )
By utilizing Equation (14), the duration of τosc and τinj can be calculated. By substituting these calculated results into Equations (10) and (11), the fixed points, xn1 and xn2, can be expressed as functions in the following manner:
{ x n 1 = F inj ( τ osc , τ inj ) x n 2 = F osc ( τ osc , x n 1 )
By substituting Equation (15) into Equation (8), the time domain solution of the state variable of the oscillating state can be obtained:
x ( t ) = Φ osc ( t t n 1 ) F inj ( τ osc , τ inj ) t n 1 t < t n 2
By substituting Equation (15) into Equation (9), the time domain solution of the state variable of the power injecting state can be obtained:
x ( t ) = Φ i ( t t n 2 ) F osc ( τ osc , x n 1 ) + A inj 1 ( Φ inj ( t t n 2 ) I ) B inj u t n 2 t ( n + 1 ) 1
According to Equation (3), if we define Cup = [1, 0, 0, 0] and Cip = [0, 1, 0, 0], the time domain expression of the voltage up and current ip can be obtained as follows:
{ u p ( t ) = C u p Φ osc ( t t n 1 ) F inj ( τ osc , τ inj ) i p ( t ) = C i p Φ osc ( t t n 1 ) F inj ( τ osc , τ inj ) t n 1 t < t n 2
If we define Cuo = [0,0,0,1], the output power can be obtained as follows:
P o = 1 T 0 T ( C u o Φ osc ( t t n 1 ) F inj ( τ osc , τ inj ) ) 2 R L d t
To verify whether the models in Figure 6 and Figure 7 and the associated theoretical analysis are consistent with Figure 4, as well as to confirm the accuracy of the proposed self-oscillating IPT system, Equations (3)–(18) were simulated and tested using MATLAB R2016b. The parameters used for the validation process are listed in Table 2.
Figure 8 illustrates the calculation results, where the voltage curve up and current curve ip align with Figure 4 and Figure 5, respectively. This confirms the accuracy of the proposed theoretical model. The calculation results indicate that the system enters the oscillating state at 20 μs, and remains in that state until 40 μs, resulting in a duration of 20 μs.
Subsequently, the system transitions to the power injecting state at 40 μs and exits this state at 60 μs, with a duration of 20 μs. During the onset of the oscillating state, the current value ip is 15.1 A. Following a 2 μs delay, ip reaches its maximum value of 15.7 A (Ipm1). Similarly, prior to the system exiting the oscillating state, ip reaches its negative maximum of −10.3 A (Ipm2). After a 4 μs delay, when the system exits the oscillating state at 40 μs, ip measures −8.5 A.
Figure 9 illustrates the impact of the coupling coefficient k on the curves of ip and up with a fixed value of T = 40 μs. As depicted in the figure, with k varying from 0.5 to 0.1, the starting time of the oscillating state advances by approximately 5 μs, while the amplitude of ip decreases by approximately 6A. For a more detailed analysis of the effect of k on the oscillating state duration, τosc, and the amplitude, Ipm1, please refer to Figure 10.
For a more detailed analysis of the effect of coupling coefficient k on the duration of the oscillating state, τosc, and the amplitude of the primary current ip, Ipm1, Figure 10 illustrates the relationship between τosc and Ipm1 as a function of k. The figure demonstrates that the τosc decreases with increasing values of k. This is attributed to the fact that a higher coupling coefficient results in a smaller leakage inductance Lpk for the loosely coupled transformer, consequently leading to a shorter oscillation period. Furthermore, it should be noted that an increase in the coupling coefficient also corresponds to an amplified amplitude of the primary current ip. This can be attributed to the enhanced power transmission to the secondary side as k increases. Therefore, on the primary side, a greater amount of power needs to be injected during the power injecting state duration.
Clearly, the calculated results of Figure 9 and Figure 10 precisely correspond to those of Figure 4 and Figure 5, providing further evidence of the accuracy of the proposed theoretical models.

4. Control Method

As analyzed in Section 2, achieving oscillation phase tracing and power injection duration control are the key objectives of the proposed IPT system. Therefore, this chapter focuses on the methods employed to fulfill these requirements. Figure 11 illustrates the circuit diagram of the dynamic IPT system studied in this paper.
In this IPT system, a closed-loop phase tracking circuit is utilized to track the phase and trigger self-oscillation. This self-oscillating circuit consists of three components: phase tracking unit, power regulating unit, and oscillating unit. With this closed-loop control, the system achieves self-oscillation.

4.1. The Oscillating Unit

The oscillating unit adopts a topology that includes a single switch converter and an oscillating network. The converter employs an IGBT as switch S, while the oscillating network consists of the primary capacitor Cp, and inductor Lp. As depicted in Figure 1, the oscillation unit can alternate between the oscillating state and the power injecting state by controlling the switch S. To acquire the phase information of the system oscillation, a voltage divider comprising Rs1 and Rs2 is utilized as a sensor to monitor the voltage across switch S, denoted as uds. The expression for uds is:
u ds = ( U DC u p )
The output of the sensor, vfb, can be calculated using the following formula:
v fb = R s 2 R s 1 + R s 2 u ds

4.2. Phase Tracking Unit

The phase tracking unit comprises Rd, Dd and comparator A1, which is employed for tracking phase changes. Rd and Dd are responsible for generating a comparison threshold voltage vd of 0.7 V, which is applied to the negative terminal of A1. The feedback signal vfb is then fed to the positive terminal of A1 and compared with vd. The output of A1, vps, can be expressed as follows:
v ps = { 0 v fb < v d U C v fb > v d
Based on Figure 4, and Equation (21), the signal vfb is a sinusoidal pulse sequence in which pulse width signifies the duration of the oscillating state, τosc (phase angle βτ2). Consequently, the duration of the high level output of vps, aligns with the duration of the oscillation state, τosc. Notably, according to Equation (22), the high-level duration of vps can be automatically adjusted to track τosc (phase angle βτ2), thereby facilitating the phase tracking within the system.
A phase tracking process can be demonstrated using Figure 12. In the figure, the curves before the phase change are highlighted in green, while the curves after the phase change are marked in red. Suppose a phase shift occurs due to some reason, this results in a change in the width of vfb from τosc to τosc’ as the phase transitions from βτ2 to βτ2’. By comparing it with the threshold vd, the width of vps also tracks this change, shifting from τosc to τosc’. As shown in Figure 12, this method enables the system to promptly track the phase shift.

4.3. Power Injecting Unit

The power regulating unit comprises a comparator A2 and an integrator. The integrator is composed of a resistor Rint and capacitor Cint. The control signal, vps, is connected to the input of the integrator, as depicted in Figure 13. This configuration demonstrates how the integrator operates under the control of output of comparator A1, vps. To simplify the analysis, the output of A1 is considered equivalent to a transistor push–pull structure.
Referring to Figure 13, the control process of vps for the integrator can be described as follows:
(1)
When vps is at a high level, the upper switch of the push–pull structure is turned on, as depicted in Figure 13b. In this case, both ends of the capacitor Cint are connected to Uc, resulting in zero voltage across Cint. Consequently, the integrator ceases its operation, and the integrator output vint = Uc.
(2)
As vps is lowered, the lower switch of the push–pull structure is activated, as shown in Figure 13c. In this scenario, the input of the integrator is connected to the ground, initiating the charging of capacitor Cint through Rint by Uc. As a result, the output of the integrator, vint, begins to increase from zero. The expression for vint can be represented as follows:
v int = 1 C int t 0 t U C R int d t
In Equation (23), t0 represents the time at which the integrator begins its operation. Assuming that the moment vps is pulled to low level corresponds to the time when the integrator starts working, then t0 can be set to 0. In this case, vint can be expressed as follows:
v int = 1 C int 0 t U C R int d t = U C R int C int t
The output of the integrator, vint, is then compared with a threshold vr in comparator A2, and the comparison result vg can be expressed as follows:
v g = { U C v int < v r 0 v int > v r
The control signal vg is sent to the oscillating unit to control the switch S, turning it on or off. It is evident that by controlling the duration of vg at the high level, the output power of the system can be regulated. Since vint is a signal that starts at zero and rises linearly, according to Equation (25), the duration of vg is determined by the threshold vr. Hence, vr can be used as a variable to control the system power.
Figure 14 illustrates the principle of utilizing vr to control the system power. In the figure, the green and red lines represent the waveforms before and after the change in vr, respectively. Figure 14 demonstrates that when the power control variable is increased from vr to vr’, the end time of power injecting state extends from t(n+1)1 to t(n+1)1’, and the duration of power injecting extends from τinj to τinj’.
As depicted in Figure 14, with the increase in τinj, the amplitudes of ip also increase from Ipm1 and Ipm2 to Ipm1’ and Ipm2’, respectively, indicating an increase in the power injected into the system. Conversely, if vr is decreased, the power injected into the system will be reduced accordingly.
Hereto, the challenges to tracking the oscillation phase and controlling the duration of power injection have been successfully resolved. To provide a clearer representation, Figure 11 is depicted as a block diagram in Figure 15, illustrating the operational process of the proposed IPT system.
Initially, the signal vfb, which carries the oscillating network information, is directed to the phase tracking unit. This unit detects the duration of the oscillating state, τosc, and generates the output signal vps.
Subsequently, vps serves as a trigger signal to initiate the power regulating unit. By utilizing the control variable vr, the power regulating unit determines the duration of the power injecting state, τinj, and generates the control signal vg.
Ultimately, the control signal vg is transmitted to the oscillating unit, which seamlessly transitions between the self-oscillating state and the power injecting state, thereby completing the self-oscillating closed loop of the system.
Figure 16 presents the curves of crucial signals during each operation of the proposed IPT system, elucidating the temporal relationship between these signals. This visualization aims to enhance comprehension of the system’s working principle.

5. Validate Prototypes and Experiment

A verification platform was built to verify the theoretical analysis of the proposed dynamic IPT system, as shown in Figure 17, which includes a prototype made according to Figure 11. In the experimental platform, an oscilloscope (ROHDE&SCHWARZ RTB2004, Munich, Germany) and a power analyzer (YOKOGAWA WT500, Tokyo, Japan) are hired to monitor the working waveform and analyze the system power, respectively. A resistance group consisting of four resistors and switches is used to provide different load resistor. In the experiment, the main parameters of the experimental platform are listed in Table 3, and Lp, Ls, Cp, Cs can be obtained from Table 2.
The primary and secondary coils used in the experimental platform are made of a 3 mm diameter Leeds wire wound 18 turns, and a cross-shaped ferrite core was installed behind the coil. Both the coils are tested by a digital electric bridge (VICTOR 4091C LCR, Shenzhen, China). According to the test results, the relationship of the coupling coefficient k between the coils and the air gap d can be obtained, as shown in Figure 18.

5.1. Self-Oscillation Characteristic Verification

The curves of ip, icp, iDC, and vds were obtained from the experiment, as depicted in Figure 19. Here, vds represents the voltage across switch S, given by Equation (20), where it is used to represent up. The following phenomena can be observed in this experiment:
(1)
The system operates in a closed loop, following the sequence of the blocking (self-oscillating) state, τ2, the diode-on state τ3, and the switch-on state τ1, as described in Figure 4. This demonstrates that the system is capable of self-oscillation, with the oscillation process repeating in the oscillating state, τosc, and the power injecting state, τinj.
(2)
During the oscillating state, iDC is equal to zero, indicating that the oscillating network is isolated from UDC. The capacitor Cp and inductor Lp form a resonant tank and initiate oscillation. The capacitor voltage, vds (up), capacitor current, icp, and the inductor current, ip, exhibit sinusoidal changes, with icp and ip being equal. Clearly, the oscillation is sustained for less than one cycle. The oscillation state duration τosc being 22 μs, slightly larger than the simulation results shown in Figure 8. This deviation may be caused by the deviation of component parameters.
(3)
In the diode-on state, iDC is negative, indicating that the current ip flows back to UDC through the diode. In the switch-on state, iDC is positive, indicating that UDC injects a current into the inductor Lp. Combining the diode-on state, τ3, and the switch-on state, τ1, into the power injecting state τinj, it can be observed that during the power injecting state, ip linearly increases from negative to positive with a rise slope of 1 × 106 A/s, consistent with the calculated result from Equation (1). Additionally, the capacitor current, icp, is zero during the power injecting state, implying that the capacitor voltage is clamped to UDC during these two states.
This experimental results are consistent with the theoretical analysis presented in Figure 1 and Figure 4. Additionally, the experimental results closely correspond to the simulation result depicted in Figure 8, thus validating the accuracy of the model established in Section 3.

5.2. Power Regulating Verification

Figure 20 exhibits the experimental curves of uds, vg, vint and ip, serving as a means to assess the performance of the IPT system outlined in Figure 11, as well as the impact of the control variable vr on the system’s output power regulation. Specifically, in subfigure (a), with vr set to 2.5 V, and in subfigure (b), with vr set to 3.5 V. The experimental results yield the following observations:
(1)
The experimental curves are exactly the same as the principle curves shown in Figure 16, which indicates that the system can realize self-oscillation according to the order of the self-oscillating state and power injecting state.
(2)
The power injecting state duration can only be controlled by the control variable vr. The comparison between (a) and (b) shows that when vr increases from 2.5 V to 3.5 V, the power injection duration τinj increases from 24 μs to 36 μs, and the amplitude Ipm1 of the current ip increases from 14 A to 20 A, indicating that the power can be adjusted by controlling vr.
It is noted that when the control variable vr adjusts the power injecting state duration, the self-oscillating state duration τosc is not affected and is maintained at 22 μs. This shows that the power injection process is decoupled from the self-oscillation process.
Figure 21 illustrates the relationship between output power, efficiency, and the control variable vr. This experimental result demonstrates a monotone positive correlation between output power and vr. The reason for this correlation is that the proposed IPT system’s power injection process is decoupled from the self-oscillation process. The power is solely regulated by the control variable vr. This characteristic simplifies the design of output power regulation and provides a convenient technical implementation for voltage, current, or power closed-loop control within the system.
There is a discrepancy between the calculated output power value and the experimental value shown in Figure 21. This deviation occurs because the phase tracking unit employs the forward voltage vd of diode D, 0.7 V, as a threshold, as illustrated in Figure 16. Consequently, the detected value of the self-oscillation state duration is smaller than the actual value, resulting in the experimental output power value being lower than the value calculated using Equation (19). Nevertheless, as the output power increases, the influence of the self-oscillating state duration diminishes, bringing the experimental output power value closer to the calculated value.
It can be observed that in Figure 21, the efficiency at vr = 2.5 V and 3.0 V reaches an optimal value of 89.2%, while it decreases to 83.5% and 87.6% at vr = 2 V and 3.5 V, respectively. This phenomenon occurs because the power is regulated by adjusting the transmission period, which can lead the system to deviate from the optimal equivalent impedance.

5.3. Phase Tracking and Soft Switch Condition Verification

Figure 22 illustrates the experimental curves of ip, uds and the control signal vg of switch S, providing evidence of the phase tracking characteristic of the proposed IPT system. The results reveal that when d = 3 cm, the self-oscillating state lasts for τosc = 22.1 μs, and the power injecting state persists for τinj = 17.8 μs. With an increase in d = 12 cm, the duration of the self-oscillating state extends to 26.3 μs, while the power injecting state remains unchanged at τinj = 17.8 μs.
This experiment clearly demonstrates the system’s excellent good phase (frequency) tracking capability and confirms that the self-oscillation method proposed in this paper does not interfere with each other between the self-oscillating state and the power injecting state. Significantly, it can be observed that, despite the variation in the self-oscillating state duration, the switch S effectively tracks the change and remains at the ZVS soft switching point.

5.4. Dynamic Characteristics Verification

To test the dynamic characteristics and robustness of the system, an experiment was conducted involving the shifting of the coil’s position. The procedure for the experiment is as follows: initially, the secondary coil was positioned more than 20 cm away from the primary coil. Subsequently, the secondary coil was rapidly moved to a position just 3 cm away from the primary coil. The experimental results are represented by the curves of vg and ip, which were captured by the oscilloscope and depicted in Figure 23.
Figure 23a illustrates the curves obtained at a scale of 500 ms/div, which is used to depict the overall dynamic process in the experiment. On the other hand, (b,c) display the curves unfolded at a scale of 20 μs/div, captured at different time points from (a), aiming to present detailed information in dynamic experiments. The experimental results demonstrate that the system can accurately track the phase change caused by the alteration of coupling coefficient. Throughout the experiment, as the coil distance varies, the duration of the oscillating state automatically adjusts from 24.2 μs to 21.4 μs.
Furthermore, an experiment on sudden load tolerance was conducted. In this experiment, the load steps from 10 Ω to 30 Ω, and then returned to 20 Ω. Figure 24a depicts the experimental results obtained at a scale of 500 ms/div, demonstrating that all the curves effectively track the changes in load.
Figure 24b provides an expanded of Figure 24a at time point ts1, offering a detailed examination of the transition from 10 Ω to 30 Ω. The revealed details indicate that the curve transition remarkably smooth, taking only 10 cycles to complete.
Moving on to Figure 24c,d, it further amplifies the data from Figure 24b, this time at a scale of 20 μs/div. Specifically, Figure 24c focuses on the RL = 10 Ω scenario, while Figure 24d looks at RL = 30 Ω. Notably, the power injection time remaining at the self-oscillation time increases from 21.4 μs to 23.6 μs. This outcome suggests that the load change solely impacts the resonant angular frequency of the self-oscillation, highlighting the decoupling of the power injection process from the self-oscillation process.
These experimental findings substantiate the effectiveness of the proposed closed loop self-oscillating structure, which exhibits the ability to track parameter changes in one period while showcasing remarkable robustness of the system.

6. Conclusions

To address the challenges associated with frequency tracking and power control in conventional IPT system, which arise due to the coupling between the converter and the resonant network, this study proposes a method for implementing an IPT system based on a phase-closed loop. This method divides a power transfer process into two distinct states: the self-oscillating state and power injecting state, and ensures their independence from each other. The effectiveness of this method was verified through simulation and experiments, which led to the following conclusions:
(1)
The working process of a converter can be divided into three distinct states: the switch-on state, blocking state, and diode-on state. This division allows for decoupling of the resonant network from the converter. By combining the switch-on state and diode-on state into power injecting state, and considering the blocking state as oscillating state, the working process of an IPT system consists of two state sequences: power injecting and oscillating states. By controlling the duration of the power injecting state, the output power can be adjusted independently. Furthermore, by detecting the duration of the self-oscillating state, changes in frequency can be accurately tracked and compensated for. The diode-on state’s soft switching characteristics enable a smooth transition between the power injecting and oscillating states.
(2)
The proposed phase-closed loop, comprising the phase tracking unit, power regulating unit, and oscillating unit, proves to be effective. This closed loop enables the system to switch between the power injecting state and oscillating state under soft switching condition.
(3)
Experimental results demonstrate that the phase tracking unit accurately tracks frequency drift caused by system parameters, indicating excellent frequency tracking characteristics. Furthermore, the experiments show that the system exhibits robust frequency tracking, even under conditions of a large coupling coefficient and maximum speed change.
(4)
The proposed power control method utilizing an integrator is proved to be effective. With the power injecting process decoupling from the oscillating process, power regulation only requires manipulation of the control variable vr. The experimental results indicate that the output power monotonically increases with the control variables vr. This power regulation characteristic simplifies and enhances the reliability of designing an IPT system control strategy.

Author Contributions

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

Funding

This research was funded by Natural Science Foundation of Fujian Province, grant number 2021J05262 and 2021J011202, Fujian Provincial Science and Technology Plan Project, grant number 2022T3061, and National Natural Science Foundation of China, grant number 52177222.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

IPTInductive Power Transfer
ZPAZero Phase Angle
PLLPhase-Locked Loop
ZVSZero Voltage Switch
UDCBus-bar voltage
LpPrimary inductance
LsSecondary inductance
LpkPrimary leakage inductance
LskSecondary leakage inductance
MMutual inductance
LMmutual inductance
kCoupling coefficient
KEPrimary current slop
CpPrimary capacitance
CsSecondary capacitance
CintIntegrating capacitance of power injecting unit
RLEquivalent load
RpsPrimary loss resistance
RssSecondary loss resistance
RintIntegrating resistance of power injecting unit
A1Phase tracking unit comparator
A2Power regulating unit comparator
SPower switch
ipPrimary inductor current
isSecondary inductor current
upPrimary capacitor voltage
uoOutput voltage
udsVoltage across switch S
vgControl signal of switch
vfbFeedback signals from the oscillating network
vdComparison threshold of Phase tracking unit
vpsOutput voltage of Phase tracking unit
vintIntegrator output voltage
vrSystem power control voltage
τ1Duration of switch-on state
τ2Duration of blocking state
τ3Duration of diode-on state
τinjDuration of power injecting state
τoscDuration of self-oscillating state
ωSelf-oscillating angular frequency
βτ1Phase angle occupied by switch-on state
βτ2Phase angle occupied by blocking state
βτ3Phase angle occupied by diode-on state
P0Output power
Ipm1/IPm2Maximum value of primary current, ip
TEnergy transfer period
dThe air gap between the two coils
IdcBus current

References

  1. Liu, H.; Shao, Q.; Fang, X. Modeling and Optimization of Class-E Amplifier at Subnominal Condition in a Wireless Power Transfer System for Biomedical Implants. IEEE Trans. Biomed. Circuits Syst. 2016, 11, 35–43. [Google Scholar] [CrossRef]
  2. Ruffo, R.; Cirimele, V.; Diana, M.; Khalilian, M.; La Ganga, A.; Guglielmi, P. Sensorless control of the charging process of a dynamic inductive power transfer system with interleaved nine-phase boost converter. IEEE Trans. Ind. Electron. 2018, 65, 7630–7639. [Google Scholar] [CrossRef]
  3. Zhao, L.; Thrimawithana, D.J.; Madawala, U.K.; Hu, A.P. A Push–Pull Parallel Resonant Converter-Based Bidirectional IPT System. IEEE Trans. Power Electron. 2019, 35, 2659–2667. [Google Scholar] [CrossRef]
  4. Wang, C.S.; Covic, G.A.; Stielau, O.H. Investigating an LCL load resonant inverter for inductive power transfer applications. IEEE Trans. Power Electron. 2004, 19, 995–1002. [Google Scholar] [CrossRef]
  5. Zhang, W.; Wong, S.-C.; Tse, C.K.; Chen, Q. Design for Efficiency Optimization and Voltage Controllability of Series–Series Compensated Inductive Power Transfer Systems. IEEE Trans. Power Electron. 2013, 29, 191–200. [Google Scholar] [CrossRef]
  6. Stielau, O.H.; Covic, G.A. Design of loosely coupled inductive power transfer systems. In Proceedings of the 2000 International Conference on Power System Technology, Perth, WA, Australia, 4–7 December 2000. [Google Scholar] [CrossRef]
  7. Mohamed, A.A.S.; Berzoy, A.; de Almeida, F.G.N.; Mohammed, O. Modeling and Assessment Analysis of Various Compensation Topologies in Bidirectional IWPT System for EV Applications. IEEE Trans. Ind. Appl. 2017, 53, 4973–4984. [Google Scholar] [CrossRef]
  8. Shevchenko, V.; Husev, O.; Strzelecki, R.; Pakhaliuk, B.; Poliakov, N.; Strzelecka, N. Compensation Topologies in IPT Systems: Standards, Requirements, Classification, Analysis, Comparison and Application. IEEE Access 2021, 7, 120559–120580. [Google Scholar] [CrossRef]
  9. Houran, M.A.; Yang, X.; Chen, W. Magnetically Coupled Resonance WPT: Review of Compensation Topologies, Resonator Structures with Misalignment, and EMI Diagnostics. Electronics 2018, 7, 296. [Google Scholar] [CrossRef]
  10. Mohamed, A.A.S.; Shaier, A.A.; Metwally, H.; Selem, S.I. An Overview of Dynamic Inductive Charging for Electric Vehicles. Energies 2022, 15, 5613. [Google Scholar] [CrossRef]
  11. Wang, C.S.; Covic, G.A.; Stielau, O.H. Power transfer capability and bifurcation phenomena of loosely coupled inductive power transfer systems. IEEE Trans. Ind. Electron. 2004, 51, 148–157. [Google Scholar] [CrossRef]
  12. Villa, J.L.; Sallan, J.; Osorio, J.F.S.; Llombart, A. High-Misalignment Tolerant Compensation Topology for ICPT Systems. IEEE Trans. Ind. Electron. 2011, 59, 945–951. [Google Scholar] [CrossRef]
  13. Fotopoulou, K.; Flynn, B.W. Wireless Power Transfer in Loosely Coupled Links: Coil Misalignment Model. IEEE Trans. Magn. 2011, 47, 416–430. [Google Scholar] [CrossRef]
  14. Su, Y.-G.; Zhang, H.-Y.; Wang, Z.-H.; Hu, A.P.; Chen, L.; Sun, Y. Steady-State Load Identification Method of Inductive Power Transfer System Based on Switching Capacitors. IEEE Trans. Power Electron. 2015, 30, 6349–6355. [Google Scholar] [CrossRef]
  15. Li, Y.; Dong, W.; Yang, Q.; Jiang, S.; Ni, X.; Liu, J. Automatic Impedance Matching Method with Adaptive Network Based Fuzzy Inference System for WPT. IEEE Trans. Ind. Inform. 2020, 16, 1076–1085. [Google Scholar] [CrossRef]
  16. Luo, Y.; Yang, Y.; Chen, S.; Wen, X. A Frequency-Tracking and Impedance-Matching Combined System for Robust Wireless Power Transfer. Int. J. Antennas Propag. 2017, 2017, 5719835. [Google Scholar] [CrossRef]
  17. Zahiribarsari, V.; Thrimawithana, D.J.; Covic, G.A. An Inductive Coupler Array for In-Motion Wireless Charging of Electric Vehicles. IEEE Trans. Power Electron. 2021, 36, 9854–9863. [Google Scholar]
  18. Gati, E.; Kampitsis, G.; Manias, S. Variable Frequency Controller for Inductive Power Transfer in Dynamic Conditions. IEEE Trans. Power Electron. 2016, 32, 1684–1696. [Google Scholar] [CrossRef]
  19. Matysik, J.T. The Current and Voltage Phase Shift Regulation in Resonant Converters with Integration Control. IEEE Trans. Ind. Electron. 2007, 54, 1240–1242. [Google Scholar] [CrossRef]
  20. Kim, D.H.; Kim, M.S.; Kim, H.J. Frequency-Tracking Algorithm Based on SOGI-FLL for Wireless Power Transfer System to Operate ZPA Region. Electronics 2020, 9, 1303. [Google Scholar] [CrossRef]
  21. Hong, J.; Guan, M.; Lin, Z.; Fang, Q.; Wu, W.; Chen, W. Series-Series/Series Compensated Inductive Power Transmission System with Symmetrical Half-Bridge Resonant Converter: Design, Analysis, and Experimental Assessment. Energies 2019, 12, 2268. [Google Scholar] [CrossRef]
  22. Namadmalan, A. Self-Oscillating Tuning Loops for Series Resonant Inductive Power Transfer Systems. IEEE Trans. Power Electron. 2016, 31, 7320–7327. [Google Scholar] [CrossRef]
  23. Xu, L.; Chen, Q.; Ren, X.; Wong, S.; Tse, C. Self-Oscillating Resonant Converter with Contactless Power Transfer and Integrated Current Sensing Transformer. IEEE Trans. Power Electron. 2016, 32, 4839–4851. [Google Scholar] [CrossRef]
  24. Namadmalan, A. Self-Oscillating Pulse Width Modulation for Inductive Power Transfer Systems. IEEE J. Emerg. Sel. Top. Power Electron. 2019, 8, 1813–1820. [Google Scholar] [CrossRef]
  25. Wei, Z.; Zhang, B.; Lin, S.; Wang, C. A Self-Oscillation WPT System with High Misalignment Tolerance. IEEE Trans. Power Electron. 2023, 39, 1870–1887. [Google Scholar] [CrossRef]
  26. Li, H.; Li, J.; Wang, K.; Chen, W.; Yang, X. A Maximum Efficiency Point Tracking Control Scheme for Wireless Power Transfer Systems Using Magnetic Resonant Coupling. IEEE Trans. Power Electron. 2015, 30, 3998–4008. [Google Scholar] [CrossRef]
  27. Miller, J.M.; Onar, O.C.; Chinthavali, M. Primary-Side Power Flow Control of Wireless Power Transfer for Electric Vehicle Charging. IEEE J. Emerg. Sel. Top. Power Electron. 2014, 3, 147–162. [Google Scholar] [CrossRef]
  28. Hu, H.; Cai, T.; Duan, S.; Zhang, X.; Niu, J.; Feng, H. An Optimal Variable Frequency Phase Shift Control Strategy for ZVS Operation within Wide Power Range in IPT Systems. IEEE Trans. Power Electron. 2019, 35, 5517–5530. [Google Scholar] [CrossRef]
  29. Berger, A.; Agostinelli, M.; Vesti, S.; Oliver, J.A.; Cobos, J.A.; Huemer, M. A Wireless Charging System Applying Phase-Shift and Amplitude Control to Maximize Efficiency and Extractable Power. IEEE Trans. Power Electron. 2015, 30, 6338–6348. [Google Scholar] [CrossRef]
  30. Wu, W.; Luo, D.; Hong, J.; Tang, Z.; Chen, W. A Six-Switch Mode Decoupled Wireless Power Transfer System with Dynamic Parameter Self-Adaption. Electronics 2023, 12, 2314. [Google Scholar] [CrossRef]
  31. Chen, L.; Hong, J.; Lin, Z.; Luo, D.; Guan, M.; Chen, W. A Converter with Automatic Stage Transition Control for Inductive Power Transfer. Energies 2020, 13, 5268. [Google Scholar] [CrossRef]
  32. Dranga, O.; Buti, B.; Nagy, I.; Funato, H. Stability analysis of nonlinear power electronic systems utilizing periodicity and introducing auxiliary state vector. IEEE Trans. Circuits Syst. I Regul. Pap. 2005, 52, 168–178. [Google Scholar] [CrossRef]
  33. Buti, B.; Nagy, I.; Ohsaki, H.; Masada, E. Novel Approach in Stability Analysis Presented in Controlled Boost Converter. In Proceedings of the Power Conversion Conference, Nagoya, Japan, 2–5 April 2007; IEEE: Piscataway, NJ, USA, 2007; pp. 581–587. [Google Scholar]
Figure 1. (a) An IPT system using P–P compensation network. (b) The IPT system that the converter is in the switch-on state. (c) The IPT system that the converter is in the blocking state. (d) The IPT system that the converter is in the diode-on state.
Figure 1. (a) An IPT system using P–P compensation network. (b) The IPT system that the converter is in the switch-on state. (c) The IPT system that the converter is in the blocking state. (d) The IPT system that the converter is in the diode-on state.
Energies 17 01814 g001
Figure 2. T equivalent model of loosely coupled transformer.
Figure 2. T equivalent model of loosely coupled transformer.
Energies 17 01814 g002
Figure 3. The states repeating order in a power transfer period.
Figure 3. The states repeating order in a power transfer period.
Energies 17 01814 g003
Figure 4. The vector trajectory diagram and curves of ip, up.
Figure 4. The vector trajectory diagram and curves of ip, up.
Energies 17 01814 g004
Figure 5. Effect of the system parameters on τosc and τinj.
Figure 5. Effect of the system parameters on τosc and τinj.
Energies 17 01814 g005
Figure 6. Power injecting model. Here, ip, is represent the primary and secondary current, and up signifies the primary capacitor voltage.
Figure 6. Power injecting model. Here, ip, is represent the primary and secondary current, and up signifies the primary capacitor voltage.
Energies 17 01814 g006
Figure 7. Oscillating model. Here, ip, is represent the primary and secondary current, and up signifies the primary capacitor voltage.
Figure 7. Oscillating model. Here, ip, is represent the primary and secondary current, and up signifies the primary capacitor voltage.
Energies 17 01814 g007
Figure 8. Theoretical calculation results of ip, up.
Figure 8. Theoretical calculation results of ip, up.
Energies 17 01814 g008
Figure 9. The influence of parameters on self-oscillation.
Figure 9. The influence of parameters on self-oscillation.
Energies 17 01814 g009
Figure 10. The relationship between coupling coefficient k, self-oscillating state duration τosc, and amplitude of ip.
Figure 10. The relationship between coupling coefficient k, self-oscillating state duration τosc, and amplitude of ip.
Energies 17 01814 g010
Figure 11. The circuit diagram of the proposed dynamic IPT system.
Figure 11. The circuit diagram of the proposed dynamic IPT system.
Energies 17 01814 g011
Figure 12. Phase tracking schematic diagram.
Figure 12. Phase tracking schematic diagram.
Energies 17 01814 g012
Figure 13. Schematic diagram of integrator operation. (a) Integrator stop state. (b) Integrator push–up state. (c) Integrator pull–down state.
Figure 13. Schematic diagram of integrator operation. (a) Integrator stop state. (b) Integrator push–up state. (c) Integrator pull–down state.
Energies 17 01814 g013
Figure 14. The principle of power regulation.
Figure 14. The principle of power regulation.
Energies 17 01814 g014
Figure 15. Self-oscillating closed loop block diagram based on Phase Tracking.
Figure 15. Self-oscillating closed loop block diagram based on Phase Tracking.
Energies 17 01814 g015
Figure 16. The curves of the proposed IPT system.
Figure 16. The curves of the proposed IPT system.
Energies 17 01814 g016
Figure 17. The verification prototype of the propose dynamic IPT system.
Figure 17. The verification prototype of the propose dynamic IPT system.
Energies 17 01814 g017
Figure 18. The relationship between the coil coupling coefficient and air gap.
Figure 18. The relationship between the coil coupling coefficient and air gap.
Energies 17 01814 g018
Figure 19. The curves of ip, icp, iDC, and vds, under UDC = 100 V, d = 5 cm, RL = 10 Ω.
Figure 19. The curves of ip, icp, iDC, and vds, under UDC = 100 V, d = 5 cm, RL = 10 Ω.
Energies 17 01814 g019
Figure 20. The performance experimental, under UDC = 100 V, d = 5 cm, RL = 10 Ω. (a) With vr set to 2.5 V. (b) With vr set to 3.5 V.
Figure 20. The performance experimental, under UDC = 100 V, d = 5 cm, RL = 10 Ω. (a) With vr set to 2.5 V. (b) With vr set to 3.5 V.
Energies 17 01814 g020
Figure 21. Power and efficiency vs. the control variable vr, under UDC = 200 V, d = 5 cm, RL = 10 Ω.
Figure 21. Power and efficiency vs. the control variable vr, under UDC = 200 V, d = 5 cm, RL = 10 Ω.
Energies 17 01814 g021
Figure 22. The experimental of phase tracking characteristic, under UDC = 100 V, RL = 10 Ω, (a) d = 3 cm, (b) d = 12 cm.
Figure 22. The experimental of phase tracking characteristic, under UDC = 100 V, RL = 10 Ω, (a) d = 3 cm, (b) d = 12 cm.
Energies 17 01814 g022
Figure 23. The experiment for dynamic characteristics and robustness of the system. (a) Experimental curves of air gap change from 20 cm to 3 cm at a scale of 500 ms/div. (b) The curve unfolded at a scale of 20 μs/div at point A. (c) The curve unfolded at a scale of 20 μs/div at point B.
Figure 23. The experiment for dynamic characteristics and robustness of the system. (a) Experimental curves of air gap change from 20 cm to 3 cm at a scale of 500 ms/div. (b) The curve unfolded at a scale of 20 μs/div at point A. (c) The curve unfolded at a scale of 20 μs/div at point B.
Energies 17 01814 g023
Figure 24. Experiment of dynamic characteristics under step change of load. (a) The curve that depicts the experimental results obtained at a scale of 500 ms/div. (b) The curve that provides an expanded of Figure 24a at time point ts1, offering a detailed examination of the transition from 10 Ω to 30 Ω. (c,d) The curve that further amplifies the data from (b), this time at a scale of 20 μs/div. Specifically, (c) focuses on the RL = 10 Ω scenario, while (d) looks at RL = 30 Ω.
Figure 24. Experiment of dynamic characteristics under step change of load. (a) The curve that depicts the experimental results obtained at a scale of 500 ms/div. (b) The curve that provides an expanded of Figure 24a at time point ts1, offering a detailed examination of the transition from 10 Ω to 30 Ω. (c,d) The curve that further amplifies the data from (b), this time at a scale of 20 μs/div. Specifically, (c) focuses on the RL = 10 Ω scenario, while (d) looks at RL = 30 Ω.
Energies 17 01814 g024
Table 1. The comparison of the decoupling ability and power regulation effect between this work and the literature.
Table 1. The comparison of the decoupling ability and power regulation effect between this work and the literature.
Ref.Compensation
Network
Closed-LoopFeedback SignalResonant ModeCoupling
between
Converter and Network
Power Control
Strategy
[20]S-SYesPrimary side CurrentPerfect resonanceStrong
[23]S-SYesPrimary side CurrentPerfect resonanceStrongRegulating phase
[25]S-CCYesCurrent from current transformerPerfect resonanceStrongRegulating phase
[26]S-SYesLoad voltagePerfect resonanceStrongAdditional converter
[28]LCC-SYesLoad voltage and currentPerfect resonanceStrongRegulating phase
[31]P-SNoquasi-resonantNoRegulating the duration of energy injection state
This workP-PYesVoltage from primary sidequasi-resonantNoRegulating the duration of energy injection state
Table 2. The main simulation parameters.
Table 2. The main simulation parameters.
ParameterDesign ValueParameterDesign Value
UDC100 VLp, Ls100 μH, 100 μH
M50 μHCp, Cs0.3 μF, 0.3 μF
Rps, Rss0.07 Ω, 0.07 ΩT40 μs
Table 3. The main parameters of the verify prototype.
Table 3. The main parameters of the verify prototype.
ParameterValueParameterValue
UDC, Uc100–200 V, 5 VComparatorLM339
Rint, Cint50 k, 20 nF Rd, Dd1 k, 1N4148
IGBTH30R1602RL10–50 Ω
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

Chen, L.; Luo, D.; Hong, J.; Guan, M.; Chen, W. Self-Oscillating Converter Based on Phase Tracking Closed Loop for a Dynamic IPT System. Energies 2024, 17, 1814. https://doi.org/10.3390/en17081814

AMA Style

Chen L, Luo D, Hong J, Guan M, Chen W. Self-Oscillating Converter Based on Phase Tracking Closed Loop for a Dynamic IPT System. Energies. 2024; 17(8):1814. https://doi.org/10.3390/en17081814

Chicago/Turabian Style

Chen, Lin, Daqing Luo, Jianfeng Hong, Mingjie Guan, and Wenxiang Chen. 2024. "Self-Oscillating Converter Based on Phase Tracking Closed Loop for a Dynamic IPT System" Energies 17, no. 8: 1814. https://doi.org/10.3390/en17081814

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop