Next Article in Journal
A Flexible Nickel-Oxide-Based RRAM Device Prepared Using the Solution Combustion Method
Previous Article in Journal
Energy-Efficient RIS-Enabled SISO-OFDMA Communication via Lower Bound Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bidirectional Power Control Strategy for On-Board Charger Based on Single-Stage Three-Phase Converter

School of Electrical Engineering, Northeast Electric Power University, Jilin 132012, China
*
Author to whom correspondence should be addressed.
Electronics 2024, 13(6), 1041; https://doi.org/10.3390/electronics13061041
Submission received: 17 January 2024 / Revised: 4 March 2024 / Accepted: 8 March 2024 / Published: 11 March 2024
(This article belongs to the Section Power Electronics)

Abstract

:
To solve the problems of large switching losses and the need for large-capacity electrolytic capacitances in three-phase DC/AC on-board chargers for vehicle-to-grid (V2G) applications, this paper proposes a single-stage bidirectional high-frequency isolated converter that eliminates the need for large-capacity capacitances. Combined with the proposed modulation scheme, it can theoretically reduce the switching loss by about two-thirds with the three-phase converter compared with the conventional modulation scheme, improving the converter’s operating efficiency and power density. Firstly, based on the characteristics of the proposed topology, a hybrid modulation scheme is proposed, which combines a phase-shift modulation scheme based on double modulation waves and a sawtooth carrier with a 1/3 modulation scheme, and the theoretical feasibility of the hybrid modulation scheme is verified using a mathematical modeling equation. Secondly, this paper provides a detailed analysis of the four operating modes of the two full-bridge circuits and the commutation process of the three-phase converter within 1/6 of the fundamental frequency cycle (P1 modulation interval). Then, the control strategy is given for the constant-current and constant-voltage charging and constant-current discharging for electric vehicle batteries. Finally, simulation results verify the correctness of the proposed topology and modulation scheme in vehicle–grid interaction.

1. Introduction

V2G technology enables the interaction between electric vehicles and the power grid, employing intelligent scheduling to manage the orderly charging and discharging of electric vehicles. This approach achieves peak shaving and valley filling and improves the efficiency of power grid operation [1,2,3]. In recent years, the research and development of electric vehicles have witnessed a prominent trend in the emergence of 800 V electric vehicle charging and discharging platforms that utilize V2G technology [4,5,6].
The 800 V electric vehicle charging and discharging platform based on V2G technology has been a new trend in the research and development of electric vehicles over recent years. This innovative approach significantly enhances the charging and driving performance of electric vehicles, thus attracting widespread attention from researchers. In constructing high-power electric vehicle charging and discharging platforms, the power density and operational efficiency of the converter, as the central power conversion device, assume crucial status [7,8,9]. In recent years, single-phase bidirectional DC/AC converters have found extensive application in both on-board and off-board charger construction [10,11,12]. However, they encounter challenges in meeting the demanding criteria of high power transmission capacity and achieving three-phase balanced charging and discharging within the charging and discharging system. A three-phase bidirectional DC/AC converter plays a crucial role in ensuring a balanced voltage and avoiding burdening the power grid [13,14]. However, traditional dual-stage line-frequency isolated three-phase converters have a large volume, high costs, high levels of noise pollution, and the necessity for electrolytic capacitances between the stages, which is challenging to apply to in-vehicle environments. An alternative dual-stage high-frequency isolation approach utilizing a high-frequency transformer (HFT) in place of a line-frequency transformer offers improved power density and operational efficiency [15,16]. Still, it requires a large-capacity capacitance to maintain the stability of the intermediate DC link, which not only increases the system cost but also affects the converter’s operational efficiency.
In response to these challenges, researchers have explored various topologies to enhance transformation efficiency and power density [17,18,19]. Refs. [20,21] introduce a cycloconverter known for its high power density and bidirectional energy flow capability; however, it suffers from significant voltage spikes that compromise its safe operation. Even with the incorporation of buffer circuits and an enhanced control strategy, this approach still adds complexity to the system and complicates the implementation process [22]. In [23], a single-stage three-phase converter with high-frequency isolation and power correction is investigated. This converter supports bidirectional power flow and finds applications in DC power grids, electric vehicles, and various other fields. Nevertheless, the excessive deployment of switching devices and an intricate control strategy diminish both the operational efficiency and system reliability of the converter.
Furthermore, refs. [24,25,26] enhance operational efficiency through the adoption of soft switching technology and the alteration of modulation schemes. In [24], soft switching technology is employed to mitigate the switching loss in three-phase inverters. However, this approach necessitates the incorporation of extra switching devices, passive components, and auxiliary circuits, leading to increased system costs and complexity. To minimize switching loss without introducing additional components or auxiliary circuits, refs. [25,26] employ an advanced modulation scheme. This scheme replaces the high-frequency mode of the switching transistor with a stable state through pulse width modulation, resulting in a one-third reduction in the average switching frequency and consequently lowering the switching loss. In [27,28], although the hybrid modulation strategy reduces 2/3 of the switching loss on partial topology, it increases the switching devices and structural redundancy. Consequently, this paper presents a single-stage, bidirectional high-frequency isolated three-phase DC/AC converter. This converter boasts single-stage direct power conversion, the absence of an intermediate electrolytic capacitance, compact transformer dimensions, and minimal switching loss.
The rest of this paper is organized as follows: The proposed topology structure and modulation scheme are introduced in Section 2, and the establishment of the relevant mathematical models and analysis of theoretical losses are mentioned. In Section 3, a detailed analysis of the proposed converter is presented, which helps to understand the working principles. In Section 4, the control strategy is analyzed to be applicable to the on-board charger during battery charging and discharging. A simulation is conducted in Section 5 for validation of the proposed power conversion system and control strategy; it conducts a comprehensive analysis of the two control modes for electric vehicle battery charging and discharging within the V2G context. In addition, simulations are carried out in programs that resemble experimentation in PSIM to better verify the bidirectional power flow and effectiveness of the topology. A discussion in Section 6 analyzes and summarizes the results obtained, and provides possible direction for future research. Conclusions are drawn and the limitations of the proposed topology are analyzed in Section 7.

2. Analysis of Topology Structure and Modulation Scheme of the Proposed Converter

2.1. Topology Description

Figure 1 illustrates the proposed single-stage bidirectional high-frequency isolated three-phase DC/AC converter topology in this paper. The primary full bridge (Q1–Q4) is interconnected with the secondary full bridge (S1–S4) via the leakage inductance Lr and HFT, which facilitates the voltage transformation and electrical isolation. The secondary-side three-phase converter (T1–T6) connects to the secondary full bridge on one end and to the filter inductances Lf and filter capacitances Cf on the other end before its final connection to the grid. The DC side is characterized by voltage Vdc and current Idc. V1 and V2, i1 and i2 represent the voltage and current on the primary and secondary sides of HFT, respectively, with HFT’s ratio denoted as n (n = 1). Vlink refers to the high-frequency pulse-width voltage of the secondary full-bridge output, while ia, ib, ic and Va, Vb, Vc represent the three-phase phase currents and phase voltages on the grid side.
Both full-bridge circuits operate in a bidirectional phase-shift mode, generating a high-frequency six-pulse pulsating waveform. The three-phase converter on the secondary side is modulated to transform the positive high-frequency pulsating square wave voltage into balanced three-phase sinusoidal voltages through filtering. Subsequently, the output voltage is linked to the grid, achieving DC/AC sine wave conversion. In contrast to the dual-stage bidirectional isolated DC/AC converter, which utilizes a large capacitance, the proposed single-stage DC/AC topology eliminates the need for a DC link. This enables direct power transfer between DC and three-phase AC, resulting in an enhanced converter lifespan and operational efficiency.

2.2. Hybrid Modulation Scheme

Figure 2 depicts a modulation scheme featuring dual modulation waves and a sawtooth carrier, which builds upon the isolated phase-shifted full-bridge modulation scheme and the cycloconverter modulation scheme. GQ1GQ4 and GS1GS4 correspond to the driving signals for the respective Insulated Gate Bipolar Transistors (IGBTs). dp and dn denote the modulation waves for the leading and lagging bridge arms within the primary full bridge, respectively. V1 and V2 represent the high-frequency pulse-width voltages on the primary and secondary sides of HFT. Vc signifies the sawtooth carrier wave, with the switching cycle Ts for each IGBT being twice the carrier cycle Tc. Lastly, Vref represents the reference modulation wave for the primary full bridge, derived by obtaining the absolute value of the three-phase line voltage at its maximum, and its frequency (300 Hz) is six times that of the sinusoidal AC line voltage.
An on-state signal is generated when Vc exceeds Vref, and conversely, when Vc exceeds Vref again, the driving signal is inverted to produce an off-state signal. Q1 and Q4 being in the on-state result in a positive high-frequency voltage on the primary side of HFT. Conversely, the on-state of Q2 and Q3 leads to a negative high-frequency voltage on the primary side of HFT. When either Q1 and Q2 or Q3 and Q4 are simultaneously on, the voltage on both the primary and secondary sides of HFT becomes zero. Q1(Q3) and Q2(Q4) are sequentially turned on, creating a phase shift angle α between Q1 and Q2. The output three-phase line voltage can be linearly adjusted by varying α, which is determined by Vref. Assuming m represents the amplitude of Vref, dp and dn can be expressed as (1).
d p = d d n = d d = m V ref
When S1 and S4 are turned on, S2 and S3 are turned off; when S2 and S3 are turned on, S1 and S4 are turned off. The secondary side full-bridge circuit’s IGBTs consistently operate at a 50% duty cycle. S1 and S4 are switched on only when Q1 and Q2 are switched on, and off when Q3 and Q4 are switched on. Conversely, S2 and S3 are switched on only when Q3 and Q4 are switched on, and off when Q1 and Q2 are switched on. Therefore, the secondary full bridge can generate unipolar high-frequency pulse-width voltage with the characteristics of a six-pulse waveform.
To demodulate the high-frequency pulse-width voltage waveform into a three-phase sinusoidal voltage waveform, modulation of the three-phase converter on the secondary side is required, and a 1/3 modulation scheme is proposed. Since Vref is a six-pulse waveform, a fundamental frequency cycle can be divided into six equal modulation intervals (P1P6) for analysis. Figure 3 illustrates the corresponding waveforms of the 1/3 modulation scheme. Mod-A, Mod-B, and Mod-C represent the reference modulation waveforms for the A-phase, B-phase, and C-phase bridge arms, respectively. GT1GT6 represent the driving signals for T1–T6.
During each modulation interval, a single bridge arm engages in high-frequency switching while the other two maintain a steady state at either a high or low level. It can be inferred that within a fundamental frequency cycle, each IGBT in a three-phase converter operates in a high-frequency switching state for one-third of the cycle. Table 1 illustrates the switching states of T1–T6 during each modulation interval.
Table 1 reveals that each IGBT in a fundamental frequency cycle has four distinct modulation states. Based on the sequence presented in the table, three reference modulation waves for the IGBTs (T1–T6) can be constructed within one cycle: Mod-A, Mod-B, and Mod-C.
A four-option state selector multiplexer (MUX) is utilized to create these waves. Initially, the four modulation states are encoded as per Table 1. This encoding process generates six signals (signal1–signal6) corresponding to each reference modulation wave. Consequently, these signals ensure that the IGBTs align with their respective modulation states during different intervals, ultimately facilitating the generation of the appropriate driving signals.
As illustrated in Table 2, it considers the generation of driving signals GT1 and GT2. Initially, the four modulation states are encoded. “00” corresponds to the high level, “01” to Vab/Vcb, “10” to Vac/Vbc, and “11” to the low level. Subsequently, using the modulation table, this encoding results in signal1 as “000111” and signal2 as “100011”, forming the A-phase reference modulation wave—Mod-A. After comparing them with a sawtooth carrier, Mod-A facilitates the generation of GT1 and GT2. Similarly, GT3GT6 are generated using these methods.
Figure 4 illustrates a phase-shift modulation scheme using dual modulation waves and a sawtooth carrier, integrated with 1/3 modulation to create a hybrid modulation approach for the converter proposed in this paper. This scheme primarily enables direct power transmission from the DC to the AC side, as depicted in Figure 4. It is important to note that this converter design supports bidirectional power conversion, allowing for power flow from the AC to the DC side.

2.3. Loss Analysis

The analysis indicates that in a three-phase converter on the secondary side, all IGBTs are in a high-frequency switching state for only one-third of the fundamental frequency cycle. This design effectively reduces switching loss, and the loss of converters with different topologies mainly includes on-state loss and switching loss. The on-state loss (Pcon) of IGBTs under different modulation schemes can be calculated using the following equation:
P con = m ( I c V ce + I c 2 r fd )
Here, m represents the adjustment factor, Ic is the collector current, Vce is the on-state voltage drop between the IGBT collector and emitter, and rfd is the on-state resistance. Under identical operating conditions, the on-state loss of the converter remains approximately consistent across different modulation schemes. The equation to calculate the switching loss (Pswi) of IGBTs is given as follows.
P swi = 1 2 π V link ( t on + t off ) 2 T s 0 2 π f i ( θ ) d θ
In this context, Vlink represents the input voltage of the three-phase converter during IGBTs’ opening and closing phases. ton and toff represent the respective opening and closing times of the IGBTs, while Ts is the switching cycle time. Additionally, fi(θ) is a function related to the high-frequency switching current. Under identical working conditions, the frequency of IGBT switching directly impacts the switching loss. The three-phase converter employs a constant DC link square wave control combined with sine pulse width modulation (SPWM). The total switching loss can be expressed as (4).
P sw , SPWM = 2 π V link I dc ( t on + t off ) f s
Psw,DPWMMIN represents the total switching loss of the high-frequency isolated three-phase converter, which uses 180° square wave control and a discretized pulse width modulation scheme based on minimum switching loss (DPWMMIN). Psw,ADPWM represents the total switching loss of the three-phase converter using the hybrid modulation scheme proposed in this paper.
P sw , DPWMMIN = P sw , SPWM ( 1 3 4 ) P sw , ADPWM = P sw , SPWM ( 1 3 2 )
The proposed scheme in this paper minimizes switching loss, effectively reducing both the average switching frequency and switching loss. This aligns with the theoretical predictions.

2.4. Mathematics Modeling

Taking the inversion process as an example, the feasibility of the proposed hybrid modulation scheme is analyzed through mathematical modeling. The rectification process follows the same approach. The reference modulation wave of the primary side full-bridge circuit is a six-pulse Vref(t), where one fundamental cycle is divided into six modulation intervals, as expressed in (6).
V ref t = V cb = sin ( ω o t + π 3 ) ,       P 1 : 0 ω o t < π / 3 V ab = sin ω o t ,         P 2 : π / 3 ω o t < 2 π / 3 V ac = sin ( ω o t π 3 ) ,       P 3 : 2 π / 3 ω o t < π V bc = sin ( ω o t 2 π 3 ) ,   P 4 : π ω o t < 4 π / 3 V ba = sin ( ω o t π ) ,       P 5 : 4 π / 3 ω o t < 5 π / 3 V ca = sin ( ω o t 4 π 3 ) ,   P 6 : 5 π / 3 ω o t < 2 π
Taking the P1 modulation interval as an example, a double fourier analysis is performed on the high-frequency pulse width voltage Vlink, resulting in the extraction of the fundamental component of Vlink as (7).
V link ( P 1 ) = m V dc sin ( ω o t + π 3 )
In (7), m represents the modulation ratio, with a value range of 0 ≤ m ≤ 1. Similarly, for the P2 modulation interval, Vlink is expressed as (8).
V link ( P 2 ) = m V dc sin ω o t
If the switch states of the bridge arms for phase A, phase B, and phase C of the three-phase converter on the secondary side are denoted as Sa, Sb, and Sc, respectively, then
V R = ( 2 S a     S b     S c ) V link 3 V Y = ( 2 S b     S a     S c ) V link 3 V B = ( 2 S c     S a     S b ) V link 3
Let the neutral point voltage be represented as eo, and express Kirchhoff’s Voltage Law equation for the three-phase converter on the secondary side as follows.
L f d i a d t + V a + e o = V R L f d i b d t + V b + e o = V Y L f d i c d t + V c + e o = V B
In (10), pairwise subtraction is applied, resulting in expression for the line voltage on the grid side as (11).
V ab = V a V b = ( S a S b ) V link ( L f d i a d t L f d i b d t ) V bc = V b V c = ( S b S c ) V link ( L f d i b d t L f d i c d t ) V ca = V c V a = ( S c S a ) V link ( L f d i c d t L f d i a d t )
The expression for the IGBT states in the P1 modulation interval is as (12).
S a S b = V ab V cb = sin ω o t sin ( ω o t   +   1 3 π ) S b S c = 1 S c S a = 1 V ab V cb = V cb V ab V cb = sin ( ω o t     4 3 π ) sin ( ω o t   +   1 3 π )
When the harmonic components are not considered, substituting (12) into (11) will result in three-phase line voltages on the grid side as (13).
V ab = m V dc sin ω o t L f d i ab d t V bc = m V dc sin ( ω o t 2 3 π ) L f d i bc d t V ca = m V dc sin ( ω o t + 2 3 π ) L f d i ca d t
From (13), the grid-side line voltages are three-phase sine voltages. By repeating the operation of the above equation, the expression of the three-phase line voltages obtained in the P2P6 modulation intervals is consistent with (13). Therefore, it can be proven that the proposed hybrid modulation scheme is feasible.

3. Operating Process and Mechanism of the Converter

3.1. Operating Principle of Two Full-Bridge Circuits

To simplify the analysis, the following assumptions are made. Assuming that all IGBTs, inductances, and capacitances are ideal, that is, the parasitic capacitance of all IGBTs (Q1–Q4, S1–S4, T1–T6) is ignored, and there is no loss in the inductances and capacitances. Assuming that the current commutation process is very rapid, it can be considered to be completed instantly, and the voltage ratio of HFT is 1. The leakage inductance Lr of HFT is equivalent to the primary side of the ideal transformer.
There are a total of eight operating modes in a complete switching cycle of two full-bridge circuits of the converter. Due to the symmetry between the last four operating modes and the first four, in order to avoid lengthy repetition, this paper only analyzes the first four operating modes. In addition, only analyzing the buck operating mode, the analysis of the boost operating mode is similar. The first four modes are described in detail below Figure 5, and graphic representations of such modes are presented in Figure 6.
Mode 1: [t0t1, Figure 5a] Q1 and Q4 are on simultaneously, while i1 flows through Q1, the primary winding, and Q4 in sequence. On the secondary side of HFT, S1 and S4 are both in the on-state, S2 and S3 are both in the off-state, and i2 only flows through the anti-parallel diodes of S1 and S4. V1 and V2 are equal to Vdc, and the output voltage Vlink of the full-bridge circuit on the secondary side is equal to Vdc.
i 1 ( t ) = V dc L r ( t t 0 )
Mode 2: [t1t2, Figure 5b] At t1, Q1 is turned off. i1 commutates from Q1 to the anti-parallel diode of Q3, V1 and V2 become 0, and the primary side of the HFT is in a freewheeling state, so the zero voltage switching (ZVS) of Q3 can be realized. There is no clamping voltage on the secondary side, and all S1–S4 are turned on. The current on the secondary side is also in a circulating state and linearly decreases.
Mode 3: [t2t3, Figure 5c] At t2, Q4 is turned off. i1 flows through the anti-parallel diodes of Q2 and Q3, and the ZVS of Q2 can be realized. The operating process on the secondary side is similar to Mode 2. In addition, Vdc is applied at both ends of Lr, V1 changes from 0 to −Vdc, but V2 remains at 0, with no power transmitted to the secondary side, resulting in duty cycle loss. In this mode, the energy stored in Lr is transmitted back to the DC side, and i1 and i2 rapidly decrease linearly. The expression for i1 is as follows.
i 1 ( t ) = i 1 ( t 2 ) - V dc L r ( t - t 2 )
Mode 4: [t3t4, Figure 5d] At t3, i1 decreases in the forward direction to 0, and i1 commutates from the anti-parallel diodes of Q2 and Q3 to Q2 and Q3. Due to the discharge of capacitance C, both i1 and i2 begin to rapidly increase in reverse, and the expression for i1 is as (16).
i 1 ( t ) = V dc L r ( t t 3 )
At t4, mode 4 ends and the i2 freewheeling state ends. i2 flows through the anti-parallel diodes of S2 and S3, while the anti-parallel diodes of S1 and S4 naturally are turned off. At this point, V1 and V2 are equal to −Vdc, and the duty cycle loss process ends.

3.2. Operating Principle of Three-Phase Converter on the Secondary Side

The commutation process of the three-phase converter in the P1 modulation interval is shown in Figure 7, Figure 8 and Figure 9. In the P1 modulation interval, T4 and T5 are normally on, T3 and T6 are normally off, and T1 and T2 are in a high-frequency modulation state, with a total of three cases. Given that the switching function of T1 is Sa1, and the switching function of T2 is Sa2, so Sa1 + Sa2 = 1.
As shown in case 1 of Figure 7, T1 is in a high-frequency switching state and T4 is normally on. When T1 is turned on, VRY is equal to Vlink; when T1 is turned off, VRY is equal to 0. VRY can be expressed using (17).
V RY = S a 1 V link
As shown in case 2 of Figure 8, both T4 and T5 are normally on, so VYB can be expressed using (18).
V YB = V link
As shown in case 3 of Figure 9, T2 is in a high-frequency switching state, and T5 is normally on. When T2 is turned on, VBR is equal to Vlink; when T2 is turned off, VBR is equal to 0. VBR can be expressed using (19).
V BR = S a 2 V link
Therefore, the high-frequency pulse width line voltages VRY, VYB, and VBR can be transformed into grid line voltages Vab, Vbc, and Vca after being filtered by LC filters. The analysis of the remaining modulation intervals is similar and will not be repeated here.

4. Control Strategy

4.1. Charging and Discharging Control Strategy

Figure 10 shows the closed-loop control of the converter, consisting of outer-loop control for battery charging and discharging and inner-loop control for the current on the grid side.
Battery charging and discharging control encompasses managing both current and voltage levels. To optimize battery lifespan and efficiency, the charging process employs a two-stage approach, initially using a constant current and transitioning to a constant voltage. Initially, a constant current is applied to charge the battery. Once the specified state of charge (SOC) threshold is reached, the current controller disengages and the voltage controller takes over, maintaining a constant-voltage charging state. This two-stage charging method mitigates the issues of overcurrent during the initial phase and inadequate current during constant-voltage charging, as well as preventing overcharging during the later stages of constant-current charging. Furthermore, the switching module within the charging and discharging system enables flexible transitions between battery charging and discharging states.
The battery discharging process utilizes a constant-current method, ensuring consistent current output from the battery. Simultaneously, it guarantees that the discharge current remains within acceptable limits while supplying energy to the grid rapidly. The control system is implemented in the dq coordinate system and employs a proportional–integral (PI) controller. The reference current, determined by the battery voltage or current controller, serves as the reference for the active current on the grid side. To ensure the unit power factor operation of the converter, the reference value for the reactive current on the grid side is set to 0. The specific control equation is provided below.
i d , ref = ( K vp + K vi s ) ( V dc , ref V dc )
i d , ref = ( K ip + K ii s ) ( I dc , ref I dc )
i q , ref = 0
In (20) and (21), Kvp and Kip represent proportional coefficients, while Kvi and Kii represent integral coefficients. Vdc,ref, Idc,ref, id,ref, iq,ref serve as reference values for Vdc, Idc, id, and iq, respectively. Grid-side current control employs an inner-loop current control system utilizing a feedforward decoupling method to achieve independent control of id and iq. Much like the outer-loop control of battery voltage or current, the inner-loop current control predominantly employs a PI controller, with the specific control equation given below.
V dref = ( k ip + k ii s ) ( i d , ref i d ) ω o L f i q + V d V qref = ( k ip + k ii s ) ( i q , ref i q ) + ω o L f i d + V q
In (23), kip represents the proportional coefficient, and kii represents the integral coefficient. Vd and Vq are the dq components of the grid-side voltages, while id and iq correspond to the dq components of the grid-side currents. Observing Figure 10, it is evident that the control strategy involves synchronizing with the phase of the grid-side voltages. To fulfill the requirements of the six-pulse wave modulation scheme, the reference modulation waveform generated using this scheme must be transformed into a line component.

4.2. Phase Adjustment of the Encoded Signals in the MUX

The reference line voltage employed in the hybrid modulation scheme mentioned earlier is derived from (24) and is specified with φ = 0°, indicating that the initial phase is set to 0 as a fundamental condition. Since the actual grid line voltage may not have an initial phase of zero, the reference modulation wave generated by the control strategy will also have a non-zero initial phase (i.e., φ ≠ 0°), resulting in a phase shift in the modulation waves for each modulation interval compared to the original modulation waves.
V ab , ref = sin ( ω o t + φ ) V bc , ref = sin ( ω o t 2 3 π + φ ) V ca , ref = sin ( ω o t + 2 3 π + φ )
Therefore, it becomes essential to align the phase of the encoded signal in the MUX with the initial phase of the actual grid line voltage. Initially, the control strategy samples the initial phase of the reference modulation wave and compares it with the zero phase to determine the phase difference. Subsequently, this phase difference is transformed into a time offset signal and incorporated into the encoded signals from signal1 to signal6.
This ensures that the modulation states of T1–T6 align with each modulation interval. The flowchart illustrating the phase adjustment control strategy is depicted in Figure 11.

5. Simulation Result

5.1. Matlab Simulation

To validate the accuracy of the proposed converter’s topology and control, a comprehensive model of the converter topology and control system is constructed using MATLAB/Simulink. Detailed simulation parameters are provided in Table 3.
For this paper, the inverter process (discharging) is referred to as forward power transmission, and the rectifier process (charging) as reverse power transmission.
Figure 12 presents the simulation results of charging and discharging. During 0.1 s–1.6 s, there is a two-stage constant-current–constant-voltage charging process. In Figure 13, the alternating current is sinusoidal and symmetrical, with a 180° phase difference between voltage and current, achieving unity power factor operation. During 0.1 s–0.8 s, the battery is in the constant-current charging region. The charging current remains constant at −30 A during this time, resulting in a gradual increase in the battery voltage and SOC. When t = 0.8 s is reached, the battery SOC reaches the preset value of 80%, entering the battery’s constant-voltage charging region. The battery terminal voltage remains constant at 659 V while the charging current gradually decreases, resulting in a further increase in the battery SOC.
At t = 1.6 s, the battery enters the constant-current discharge region. At this point, the grid-side current is sinusoidal and symmetrical, voltage and current are in phase, with a power factor of 0.999, and the battery releases energy with a high-power factor into the grid. The constant-current discharge region is set to maintain a constant current of 30 A. During this period, the battery voltage gradually decreases, and the battery SOC decreases accordingly. During the dynamic process of battery charging and discharging, the charging current and voltage quickly follow the set values, and the switching process occurs smoothly and stably. This indicates that the converter can achieve bidirectional energy transfer with a unity power factor and exhibits a good transient response, demonstrating the feasibility of the proposed battery charging and discharging control strategy.
Harmonic analysis of grid-side AC current under both battery charging and discharging states is conducted, as shown in Figure 14a,b. The distortion rate of grid-side current during charging is 1.03%, while it is 1.69% during discharging. The grid-side current distortion is relatively low, complying with grid connection standards.
Figure 15 illustrates the primary-side voltage V1 and the high-frequency pulse-width voltage Vlink output from the secondary-side full-bridge circuit. The waveform of current i2 is consistent with the analysis of various operating modes presented in Figure 6.
Figure 16 shows the reference modulation wave Vref of the primary-side full-bridge circuit, the reference modulation waves Mod-A, Mod-B, Mod-C of the secondary-side three-phase converter, and their driving signals GT1, GT3, GT5. The simulated waveforms align with the analysis presented in the previous section on hybrid modulation, demonstrating that the IGBTs of the proposed converter operate in a high-frequency modulation state only during 1/3 of the fundamental frequency cycle.
The simulation results validate the feasibility of the converter topology and the proposed hybrid modulation scheme in the context of electric vehicle battery charging and discharging scenarios.

5.2. Psim Simulation

In order to better verify the correctness and effectiveness of the topology, simulations were carried out in programs that resemble experimentation in PSIM. The use of PSIM simulation mainly verifies the bidirectional power flow of the proposed topology and the calculation of related efficiency. Detailed specifications and component parameters are listed in Table 4.
Figure 17a,b verifies the correctness of the converter in the inverter state. The DC-side voltage is 600 V, with a modulation ratio of 0.9, and resistance load value under full load on the AC side of 8.06 Ω, and a rated power set to 18 kW. Figure 17a shows the voltage and current waveforms of phase A. Figure 17b shows the three-phase voltage output by the converter with a resistive load on the AC side. Figure 17c verifies the correctness of the converter in the rectified state. The AC-side phase voltage is 220 V, with a modulation ratio of 0.9. Figure 17c shows the DC-side voltage.
The waveforms representing the port characteristics and modulation characteristics are shown in Figure 18. Furthermore, we simulate the efficiency of the inverter when the AC side of the converter is a resistive load. For example, the light load operation is represented by 20 percent of rated power (18 kW) here, indicating that the proposed converter’s output power is 3.6 kW. The values of the resistance load under light- and full-load condition are 40.33 Ω and 8.06 Ω, respectively. Figure 19 shows the variation of PSIM experimental efficiency with the load for the presented converter. The maximum efficiency achieved in the experimental 18 kW prototype in PSIM is 94.45%.
The above simulation results are consistent with the theoretical analysis, which verifies the correctness of the proposed modulation scheme.

6. Discussion

In the Simulink simulation, the scenario of charging and discharging electric vehicle batteries using the constant-voltage and -current methods is simulated. The dynamic performance of battery voltage and current is good, and the target value could be quickly tracked. The results indicate that the converter has greater potential in V2G power conversion. In addition, regarding the modulation characteristics of the converter, simulation results show that compared to SPWM modulation and other types of modulation, the proposed converter with hybrid modulation scheme reduces the average switching frequency and reduces switching losses.
In a PSIM simulation that resembles the experimental environment, the rectification and inversion of the converter are verified, and the efficiency of the converter under inversion is verified under a rated load of 18 kw on the grid side, fully demonstrating the correctness of the converter.
Looking ahead, our future work will focus on exploring advancements in our approach. In future work, a system with supply voltage asymmetry should be considered. Because the actual power grid is not in an ideal state, the proposed converter should overcome power supply voltage imbalance and be more suitable for V2G or other application scenarios. In addition, in order to improve the power density and reduce the waveform distortion rate, and to be better applicable to more scenarios, a control strategy for related harmonic suppression is also the next step from this paper. Finally, we will explore the use of the proposed converter in more scenarios, such as, three-phase uninterruptible power supply in an islanded or standalone microgrid.

7. Conclusions

This paper offers a comprehensive analysis covering modeling derivation, operating principles, a hybrid modulation scheme, battery charging and discharging control strategies, as well as simulation verification for the proposed single-stage high-frequency isolated three-phase bidirectional DC/AC converter topology. In the context of V2G applications, the proposed converter topology and control strategy exhibit the following benefits:
Firstly, the proposed converter utilizes a single-stage topology structure, eliminating the need for an intermediate DC link and enabling genuine direct power transmission. In contrast to the conventional isolated two-stage charging and discharging topology, it offers superior power density and efficiency, making it suitable for vehicle–grid applications.
Secondly, the linear and straightforward relationship between the output voltage and modulation ratio allows for power conversion through a straightforward hybrid modulation scheme, making it highly practical for engineering applications.
Finally, within a single fundamental frequency cycle, each IGBT in the three-phase converter operates in a high-frequency modulation state for just 1/3 of the cycle, maintaining a stable state for the remaining 2/3. This approach minimizes switching loss and enhances the overall operational efficiency of the converter.
However, the proposed topology also has shortcomings for managing energy flows. The application of the proposed converter in high-power conversion scenarios poses challenges. The Vlink is very high, which will affect the operation of IGBT and may generate voltage spikes. In practice, high-power conversion will bring serious waveform distortion and unstable factors to the power grid. This converter is currently suitable for a system with supply voltage symmetry, and application in asymmetric systems remains a challenge.

Author Contributions

Conceptualization and methodology, L.S., X.W. and C.M.; software, L.S. and X.W.; investigation, L.S.; validation, L.S. and X.W.; writing—original draft preparation, X.W.; writing—review and editing, L.S., X.W. and C.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors are grateful for the help provided by the School of Electrical Engineering at Northeast Electric Power University.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

V2GVehicle-to-grid
HFTHigh-frequency transformer
IGBTSInsulated gate bipolar transistors
MUXMultiplexer
SPWMSine pulse width modulation
DPWMMINDiscretized pulse width modulation scheme based on minimum switching loss
ZVSZero voltage switching
SOCState of charge
PIProportional–integral
Q1, Q2, Q3, Q4, S1, S2, S3, S4, T1, T2, T3, T4, T5, T6IGBTS with the proposed topology structure
GQ1, GQ2, GQ3, GQ4
GS1, GS2, GS3, GS4
GT1, GT2, GT3, GT4
GT5, GT6
IGBTS’ driving signals
dp, dnModulation waves for leading and lagging bridge arms in the primary full bridge
V1, V2Voltages on the primary and secondary sides of high-frequency transformer
i1, i2Currents on the primary and secondary sides of high-frequency transformer
nThe voltage ratio of HFT
VcSawtooth carrier wave
TsSwitching cycle
fsIGBT’s switching frequency
TcCarrier cycle
fcCarrier switching frequency
VrefThe reference modulation wave for the primary full bridge
LfAC-side filter inductance
CfAC-side filter capacitance
RAC-side resistance
LrLeakage inductance
LDC-side filter inductance
CDC-side filter capacitance
VdcDC-side voltage
IdcDC-side current
CrRated capacity of battery
VlinkVoltage of the secondary full-bridge output
ia, ib, icThree-phase phase currents
Va, Vb, VcThree-phase phase voltages
Va,ref, Vb,ref, Vc,refReference for three-phase phase voltages
ΦPhase-voltage initial phase
Vab, Vbc, VcaThree-phase line voltages
Vab,ref, Vbc,ref, Vca,refReference for three-phase line voltages
iab, ibc, icaThree-phase line currents
VR, VY, VBBridge arms midpoint voltages
VRY, VYB, VBRBridge arms midpoint line voltages
P1, P2, P3, P4, P5, P6Modulation intervals
Mod-A, Mod-B,
Mod-C
Reference modulation waveforms for the A-phase, B-phase, and C-phase bridge arms
signal1, signal2,
signal3, signal4,
signal5, signal6
Input selection control signals in MUX
IcCollector current
VceOn-state voltage drop
rfdOn-state resistance
PconOn-state loss
PswiSwitching loss
ton, toffOpening and closing times
Psw,DPWMMINTotal switching loss of discretized pulse width modulation scheme based on minimum switching loss
Psw,ADPWMTotal switching loss of hybrid modulation scheme proposed in this paper
Psw,SPWMTotal switching loss of sine pulse width modulation scheme
mModulation ratio
Sa, Sb, ScSwitch states of the bridge arms for phase A, phase B, and phase C
Sa1, Sa2Switching function of T1 and T2
eoNeutral point voltage
ωoAngular frequency
Kvi, Kii, kiiIntegral coefficients
Kvp, Kip, kipProportional coefficients
Vdc,refReference for DC-side voltage
Idc,refReference for DC-side current
id, iqdq components of the grid-side currents
id,ref, iq,refReference for dq components of the grid-side currents
Vd, Vqdq components of the grid-side voltages
Vdref, VqrefReference for dq components of the grid-side voltages

References

  1. Salvatti, G.A.; Carati, E.G.; Cardoso, R.; da Costa, J.P.; Stein, C.M. Electric Vehicles Energy Management with V2G/G2V Multifactor Optimization of Smart Grids. Energies 2020, 13, 1191. [Google Scholar] [CrossRef]
  2. Shi, R.; Li, S.; Zhang, P.; Lee, K.Y. Integration of renewable energy sources and electric vehicles in V2G network with adjustable robust optimization. Renew. Energy 2020, 153, 1067–1080. [Google Scholar] [CrossRef]
  3. Wei, H.; Zhang, Y.; Wang, Y.; Hua, W.; Jing, R.; Zhou, Y. Planning integrated energy systems coupling V2G as a flexible storage. Energy 2022, 239, 122215. [Google Scholar] [CrossRef]
  4. Vijayan, A.K.; Chakkalakkal, S.; Emadi, A. Efficiency Evaluation of 800V Electric Vehicle Powertrain using Two-Level Voltage Source Inverter with different Modulation Techniques. In Proceedings of the 2023 IEEE Transportation Electrification Conference & Expo (ITEC), Detroit, MI, USA, 21–23 June 2023; pp. 1–6. [Google Scholar]
  5. Tran, H.N.; Le, T.T.; Jeong, H.; Kim, S.; Kieu, H.P.; Choi, S. High Power Density DC-DC Converter for 800V Fuel Cell Electric Vehicles. In Proceedings of the 2021 IEEE 12th Energy Conversion Congress & Exposition-Asia (ECCE-Asia), Singapore, 24–27 May 2021; pp. 2224–2228. [Google Scholar]
  6. Aghabali, I.; Bauman, J.; Kollmeyer, P.J.; Wang, Y.; Bilgin, B.; Emadi, A. 800-V Electric Vehicle Powertrains: Review and Analysis of Benefits, Challenges, and Future Trends. IEEE Trans. Transp. Electrif. 2021, 7, 927–948. [Google Scholar] [CrossRef]
  7. Lipu, M.S.H.; Faisal, M.; Ansari, S.; Hannan, M.A.; Karim, T.F.; Ayob, A.; Hussain, A.; Miah, M.S.; Saad, M.H. Review of Electric Vehicle Converter Configurations, Control Schemes and Optimizations: Challenges and Suggestions. Electronics 2021, 10, 477. [Google Scholar] [CrossRef]
  8. Safayatullah, M.; Elrais, M.T.; Ghosh, S.; Rezaii, R.; Batarseh, I. A Comprehensive Review of Power Converter Topologies and Control Methods for Electric Vehicle Fast Charging Applications. IEEE Access 2022, 10, 40753–40793. [Google Scholar] [CrossRef]
  9. Ta, L.A.D.; Dao, N.D.; Lee, D.C. High-Efficiency Hybrid LLC Resonant Converter for On-Board Chargers of Plug-In Electric Vehicles. IEEE Trans. Power Electron. 2020, 35, 8324–8334. [Google Scholar] [CrossRef]
  10. Verma, A.K.; Singh, B.; Shahani, D.T. Grid to vehicle and vehicle to grid energy transfer using single-phase bidirectional AC-DC converter and bidirectional DC-DC converter. In Proceedings of the 2011 International Conference on Energy, Automation and Signal, Bhubaneswar, India, 28–30 December 2011; pp. 1–5. [Google Scholar]
  11. Lisheng, S.; Meintz, A.; Ferdowsi, M. Single-phase bidirectional AC-DC converters for plug-in hybrid electric vehicle applications. In Proceedings of the 2008 IEEE Vehicle Power and Propulsion Conference, Harbin, China, 3–5 September 2008; pp. 1–5. [Google Scholar]
  12. Prasanna, U.R.; Singh, A.K.; Rajashekara, K. Novel Bidirectional Single-phase Single-Stage Isolated AC–DC Converter With PFC for Charging of Electric Vehicles. IEEE Trans. Transp. Electrif. 2017, 3, 536–544. [Google Scholar] [CrossRef]
  13. Wang, J.; Wu, H.; Yang, T.; Zhang, L.; Xing, Y. Bidirectional Three-Phase DC–AC Converter With Embedded DC–DC Converter and Carrier-Based PWM Strategy for Wide Voltage Range Applications. IEEE Trans. Ind. Electron. 2019, 66, 4144–4155. [Google Scholar] [CrossRef]
  14. Almeida, B.R.d.; Araújo, J.W.M.d.; Praça, P.P.; Oliveira, D.d.S. A Single-Stage Three-Phase Bidirectional AC/DC Converter With High-Frequency Isolation and PFC. IEEE Trans. Power Electron. 2018, 33, 8298–8307. [Google Scholar] [CrossRef]
  15. Pool-Mazun, E.; Sandoval, J.J.; Enjeti, P.; Pitel, I.J. A Direct Switch-Mode Three-Phase AC to DC Rectifier with High-Frequency Isolation for Fast EV Battery Chargers. In Proceedings of the 2019 IEEE Applied Power Electronics Conference and Exposition (APEC), Anaheim, CA, USA, 17–21 March 2019; pp. 573–580. [Google Scholar]
  16. Wijeratne, D.S.; Moschopoulos, G. A Three-Phase Single-Stage AC–DC PWM Buck-Type Full-Bridge Converter: Analysis, Design, and Characteristics. IEEE Trans. Ind. Electron. 2013, 60, 4201–4214. [Google Scholar] [CrossRef]
  17. Anderson, J.A.; Haider, M.; Bortis, D.; Kolar, J.W.; Kasper, M.; Deboy, G. New Synergetic Control of a 20kW Isolated VIENNA Rectifier Front-End EV Battery Charger. In Proceedings of the 2019 20th Workshop on Control and Modeling for Power Electronics (COMPEL), Toronto, ON, Canada, 17–20 June 2019; pp. 1–8. [Google Scholar]
  18. Das, D.; Basu, K. A Soft-switched isolated Single Stage Bidirectional Three phase AC-DC Converter. In Proceedings of the 2019 IEEE Energy Conversion Congress and Exposition (ECCE), Baltimore, MD, USA, 29 September–3 October 2019; pp. 596–601. [Google Scholar]
  19. Song, J.; Ding, W.; Yang, D.; Bai, H.; Zhang, C.; Duan, B. A Novel Three-Phase Single-Stage Isolated AC-DC Converter with Symmetrical Structure for Battery Charger. In Proceedings of the 2019 3rd Conference on Vehicle Control and Intelligence (CVCI), Hefei, China, 21–22 September 2019; pp. 1–5. [Google Scholar]
  20. Yamato, I.; Tokunaga, N. Power loss reduction techniques for three phase high frequency link DC-AC converter. In Proceedings of the IEEE Power Electronics Specialist Conference—PESC’93, Seattle, WA, USA, 20–24 June 1993; pp. 663–668. [Google Scholar]
  21. Siyoung, K.; Sul, S.K.; Lipo, T.A. AC/AC power conversion based on matrix converter topology with unidirectional switches. IEEE Trans. Ind. Appl. 2000, 36, 139–145. [Google Scholar] [CrossRef]
  22. Matsui, M.; Nagai, M.; Mochizuki, M.; Nabae, A. High-frequency link DC/AC converter with suppressed voltage clamp circuits-naturally commutated phase angle control with self turn-off devices. IEEE Trans. Ind. Appl. 1996, 32, 293–300. [Google Scholar] [CrossRef]
  23. Almeida, B.R.d.; Oliveira, D.d.S.; Praça, P.P. A bidirectional single-stage three-phase rectifier with high-frequency isolation and power factor correction. In Proceedings of the 2016 IEEE Applied Power Electronics Conference and Exposition (APEC), Long Beach, CA, USA, 20–24 March 2016; pp. 60–65. [Google Scholar]
  24. Sun, P.; Lai, J.S.; Qian, H.; Yu, W.; Smith, C.; Bates, J. High efficiency three-phase soft-switching inverter for electric vehicle drives. In Proceedings of the 2009 IEEE Vehicle Power and Propulsion Conference, Dearborn, MI, USA, 7–10 September 2009; pp. 761–766. [Google Scholar]
  25. Takeda, K.; Ichinose, M.; Taniguchi, M.; Miyata, H. The alternative pulse reduction algorithm for three-phase voltage-source inverter. In Proceedings of the 2010 International Power Electronics Conference—ECCE ASIA, Sapporo, Japan, 21–24 June 2010; pp. 3087–3092. [Google Scholar]
  26. De, D.; Ramanarayanan, V. A Proportional + Multiresonant Controller for Three-Phase Four-Wire High-Frequency Link Inverter. IEEE Trans. Power Electron. 2010, 25, 899–906. [Google Scholar] [CrossRef]
  27. Mazumder, S.K.; Rathore, A.K. Performance evaluation of a new hybrid-modulation scheme for high-frequency-ac-link inverter: Applications for PV, wind, fuel-cell, and DER/storage applications. In Proceedings of the 2010 IEEE Energy Conversion Congress and Exposition, Atlanta, GA, USA, 12–16 September 2010; pp. 2529–2534. [Google Scholar]
  28. Huang, R.; Mazumder, S.K. A Soft Switching Scheme for Multiphase DC/Pulsating-DC Converter for Three-Phase High-Frequency-Link Pulsewidth Modulation (PWM) Inverter. IEEE Trans. Power Electron. 2010, 25, 1761–1774. [Google Scholar] [CrossRef]
Figure 1. The single-stage bidirectional isolated three-phase DC/AC converter topology.
Figure 1. The single-stage bidirectional isolated three-phase DC/AC converter topology.
Electronics 13 01041 g001
Figure 2. Full-bridge circuit-switch timing diagram.
Figure 2. Full-bridge circuit-switch timing diagram.
Electronics 13 01041 g002
Figure 3. Key waveforms of 1/3 modulation scheme.
Figure 3. Key waveforms of 1/3 modulation scheme.
Electronics 13 01041 g003
Figure 4. Overall hybrid modulation scheme.
Figure 4. Overall hybrid modulation scheme.
Electronics 13 01041 g004
Figure 5. Operating modes of the full-bridge circuit: (a) mode 1; (b) mode 2; (c) mode 3; (d) mode 4.
Figure 5. Operating modes of the full-bridge circuit: (a) mode 1; (b) mode 2; (c) mode 3; (d) mode 4.
Electronics 13 01041 g005aElectronics 13 01041 g005b
Figure 6. The theoretical waveforms of the full-bridge circuit in buck mode.
Figure 6. The theoretical waveforms of the full-bridge circuit in buck mode.
Electronics 13 01041 g006
Figure 7. Case 1.
Figure 7. Case 1.
Electronics 13 01041 g007
Figure 8. Case 2.
Figure 8. Case 2.
Electronics 13 01041 g008
Figure 9. Case 3.
Figure 9. Case 3.
Electronics 13 01041 g009
Figure 10. Charging and discharging control strategy.
Figure 10. Charging and discharging control strategy.
Electronics 13 01041 g010
Figure 11. Phase adjustment control strategy.
Figure 11. Phase adjustment control strategy.
Electronics 13 01041 g011
Figure 12. Charging and discharging modes.
Figure 12. Charging and discharging modes.
Electronics 13 01041 g012
Figure 13. Grid-side currents and voltages.
Figure 13. Grid-side currents and voltages.
Electronics 13 01041 g013
Figure 14. Analysis of grid-side current distortion: (a) FFT of grid-side charging current; (b) FFT of grid-side discharging current.
Figure 14. Analysis of grid-side current distortion: (a) FFT of grid-side charging current; (b) FFT of grid-side discharging current.
Electronics 13 01041 g014
Figure 15. Waveforms of i1, V1, and Vlink.
Figure 15. Waveforms of i1, V1, and Vlink.
Electronics 13 01041 g015
Figure 16. Key waveforms during the modulation process of three-phase converter.
Figure 16. Key waveforms during the modulation process of three-phase converter.
Electronics 13 01041 g016
Figure 17. Analysis of bidirectional power flow: (a) voltage and current waveforms of phase A; (b) three-phase output voltages; (c) DC-side voltage.
Figure 17. Analysis of bidirectional power flow: (a) voltage and current waveforms of phase A; (b) three-phase output voltages; (c) DC-side voltage.
Electronics 13 01041 g017aElectronics 13 01041 g017b
Figure 18. Important verification waveforms.
Figure 18. Important verification waveforms.
Electronics 13 01041 g018
Figure 19. Variation of 18 kW converter efficiency with load.
Figure 19. Variation of 18 kW converter efficiency with load.
Electronics 13 01041 g019
Table 1. Modulation states of T1–T6.
Table 1. Modulation states of T1–T6.
VrefIntervalT1  T 2T3  T 4T5  T 6
VcbP1Vab/Vcb01
VabP210Vcb/Vab
VacP31Vbc/Vac0
VbcP4Vac/Vbc10
VbaP501Vca/Vba
VcaP60Vba/Vca1
Table 2. MUX partial encoding rules.
Table 2. MUX partial encoding rules.
IntervalT1  T 2Signal1Signal2Encoded Signal
P1Vab/Vcb0101
P210000
P310000
P4Vac/Vbc1010
P501111
P601111
Table 3. Simulation parameters in simulink.
Table 3. Simulation parameters in simulink.
ParametersSymbolValue
Battery nominal voltageVdc600 V
Rated capacity of batteryCr60 Ah
DC-side capacitanceC3 mF
DC-side inductanceL500 μH
Carrier switching frequencyfc20 kHz
IGBT switching frequencyfs10 kHz
The voltage ratio of HFTn1
AC-side filtering inductanceLf5 mH
AC-side filtering capacitanceCf20 μF
Phase voltageVa, Vb, Vc220 V
Phase voltage initial phaseΦ0
AC-side resistanceR0.1 Ω
Table 4. Simulation parameters in PSIM.
Table 4. Simulation parameters in PSIM.
ParametersSymbolValue
DC-side voltageVdc600 V
DC-side capacitanceC1 mF
DC-side inductanceL1 mH
Carrier switching frequencyfc20 kHz
IGBT switching frequencyfs10 kHz
The voltage ratio of HFTn1
AC-side filtering inductanceLf1 mH
AC-side filtering capacitanceCf20 μF
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

Sun, L.; Wang, X.; Ma, C. Bidirectional Power Control Strategy for On-Board Charger Based on Single-Stage Three-Phase Converter. Electronics 2024, 13, 1041. https://doi.org/10.3390/electronics13061041

AMA Style

Sun L, Wang X, Ma C. Bidirectional Power Control Strategy for On-Board Charger Based on Single-Stage Three-Phase Converter. Electronics. 2024; 13(6):1041. https://doi.org/10.3390/electronics13061041

Chicago/Turabian Style

Sun, Li, Xin Wang, and Chenglian Ma. 2024. "Bidirectional Power Control Strategy for On-Board Charger Based on Single-Stage Three-Phase Converter" Electronics 13, no. 6: 1041. https://doi.org/10.3390/electronics13061041

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