Next Article in Journal
Etching Rate Analysis Model Based on Quartz Bond Angle Characteristics
Previous Article in Journal
Mechanical Properties and Interfacial Characterization of Additive-Manufactured CuZrCr/CoCrMo Multi-Metals Fabricated by Powder Bed Fusion Using Pulsed Wave Laser
Previous Article in Special Issue
IGBT Gate Boost Drive Technology for Promoting the Overload Capacity of Traction Converter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Performance Multi-Level Inverter with Symmetry and Simplification

1
Department of Electrical Engineering, Feng Chia University, No. 100, Wenhwa Road, Seatwen, Taichung 40724, Taiwan
2
Department of Electrical Engineering, National Taipei University of Technology, No. 1, Sec. No. 3, Zhongxiao E. Rd., Taipei 10608, Taiwan
*
Author to whom correspondence should be addressed.
Micromachines 2024, 15(6), 766; https://doi.org/10.3390/mi15060766
Submission received: 6 May 2024 / Revised: 1 June 2024 / Accepted: 5 June 2024 / Published: 7 June 2024
(This article belongs to the Special Issue Insulated Gate Bipolar Transistor (IGBT) Modules)

Abstract

:
This paper presents a high-performance, multilevel inverter with symmetry and simplification. This inverter is a single-phase, seven-level symmetric switched-capacitor inverter based on the concept of the double voltage clamping circuit connected to the half-bridge circuit. Above all, only a single DC power supply is used to achieve a single-phase inverter with seven levels and a voltage gain of three. In addition to analyzing the operating principle of the proposed switched-capacitor multilevel inverter in detail, the stability analysis and controller design are carried out by the state-space averaging method. The feasibility and effectiveness of the proposed structure are validated by some simulated results based on the PSIM simulation tool and by some experiments using FPGA as a control kernel, respectively. However, in this study, the switches were implemented by MOSFETs to meet a high switching frequency. These MOSFETs can be replaced by IGBTs in the motor drive applications so that the used switching frequency can be reduced.

1. Introduction

Due to the limitation of fossil fuels and their non-renewable nature after combustion, sustainable energy has become a major topic of discussion in today’s society. Sustainable oil and coal are decreasing year by year, while the amount of renewable energy is increasing year by year, with hydroelectricity being the most important type of renewable energy. Meanwhile, solar power and biomass power generation are also increasing year by year [1].
Renewable energy can help promote the decentralization of energy and electricity and enhance the security and reliability of energy supply. However, the development of renewable energy still faces a number of challenges, such as energy storage and transmission technologies, energy policies, and market uncertainties. Therefore, in the process of promoting the development of renewable energy, it is necessary to strengthen technological research and development and market development and actively formulate and implement policies and regulations favorable to the development of renewable energy. The American Institute of Electrical and Electronics Engineers (Institute of Electrical and Electronics Engineers, IEEE) developed the IEEE 1547 standard [2] is an interconnection standard for distributed energy resources, including solar photovoltaic systems, wind turbines, fuel cells, etc. The standard includes many requirements, such as AC power control, frequency and voltage control, protection and safety, fault and failure mode management, etc. The IEEE 519 standard [3] proposes to protect the equipment in the power system from the adverse effects of harmonics and to protect the harmonic content in the system from exceeding the safe and acceptable limits. In addition, recommendations are provided on the limitation and control of harmonic voltages and currents in power systems. In Europe, the International Electrotechnical Commission (IEC) has issued the IEC 60038 standard [4], which defines the range of voltages and frequencies to be supplied to power systems used at fixed frequencies to ensure compatibility between different types of electrical equipment and stability and reliability under different environmental conditions. At the same time, a series of IEC 61000 standards [5] was promulgated under the title electromagnetic compatibility (EMC). Among these standards, the IEC 61000-3-X standard covers several aspects, including output and input noise limitations for electrical and electronic equipment in power systems, limitations on harmonics and abnormal voltages, etc. Among these, IEC 61000-3-2 [5] aims to limit the effect of electrical equipment (e.g., household appliances, industrial equipment, etc.) on harmonic currents in the public low-voltage power supply network, and IEC 61000-3-3 [6] aims to determine the limits of the input noise of electronic equipment in low-voltage power systems in order to ensure the compatibility of the various types of electronic equipment and to minimize the disturbances and damage in the power grid. In addition, the IEC 62109-1 standard [7] has been developed for the specification of safety requirements for inverters in solar power systems.
Currently, the electricity provided by renewable energy can be categorized as an alternating current (AC) through rotary motors for hydropower and wind power or a direct current (DC) through semiconductor physical and chemical energy conversion for solar power and fuel cells. In order to cope with the AC system of utility, the AC/DC converter combined with the DC/AC converter or DC/AC converter can be used to convert the green energy to AC power. Therefore, the DC/AC converter is one of the most important technologies used for power transmission.
Nowadays, the development of DC–AC converters tends to comprise high efficiency and low harmonic components; therefore, the purpose of the development trend of a multilevel inverter (MLI) is to achieve a large number of levels with a small number of power components to minimize the component loss. The principle of the multilevel inverter is to utilize several capacitors and power switches to change the conduction path to clamp and divide the voltage of several levels so as to make the output voltage present at several levels. The more levels there are, the more the output voltage can reduce the multiple harmonic components of the output waveform without the need for external inductors or filters, and the output voltage can be close to the sinusoidal waveform. In addition, the voltage on part of the switches can be clamped by the level voltage to reduce the voltage stress on them and improve the overall output efficiency. The literature [8] discusses the operating principles of classical multilevel inverter circuits and how to extend the inverter topology to increase the number of levels, including a neutral point clamped inverter (NPC), flying capacitor inverter (FC), and cascaded H-bridge converter (CBC), cascaded H-Bridge converter (CHB), and so on.
As shown in Figure 1, considering the development cost and component selection of multilevel inverters, the output voltage quality can be improved through several modular circuits. This figure discusses four ways to increase the number of voltage levels of multilevel inverters. Circuits A, B, and C can be multilevel inverters, half-bridge or full-bridge circuits composed of switches or switched-capacitor circuits composed of capacitors and switches.
In Figure 1a, the cascaded type is the output of circuit A connected with the output of circuit B, and the system output comes from circuit B. In [9], a module consisting of two independent voltage sources and two power switches is used as the input voltage source of a conventional full-bridge, where the input voltage of the full-bridge is varied by switching the power switches to achieve a five-level voltage output. Study [10] continues the concept of study [9] by combining a boost converter with a diode voltage doubling circuit to boost a single voltage source into a DC voltage with several voltage levels, which can be supplied to the conventional full-bridge for the multilevel AC voltage output. The literature [11] consists of two sets of three-level diode neutral point clamped (DNPC) circuits, where the output of the first set is connected to the neutral point of the second set to achieve a seven-level voltage output and to minimize the impact of the load on the neutral point of voltage. The literature [12] proposes a seventeen-level voltage output, which is characterized by replacing part of the half-bridge circuit in the active neutral point clamped (ANPC) circuit with a high-order flying capacitor (FC) circuit to reduce the majority of the voltage stresses on the power switches to 1/16 of the input voltage. The circuit structures shown in [9,10,11,12] can construct the output capability of circuit B based on the output characteristics of circuit A. However, as the number of levels increases, the number of components increases, leading to an increase in the conduction losses in the circuit.
In Figure 1b, the symmetric type is two identical circuits connected together, and the system output comes from their individual terminals. In [13,14], two sets of three-level DNPC circuits were connected in a pair to achieve a five-level voltage output. In addition, the neutral points of the two identical circuits are connected to the neutral points of the two capacitors, and the current flow to the neutral point of the two circuits was canceled by the unipolar pulse width modulation (PWM) method to solve the neutral point voltage offset problem of the triple-neutral point clamped circuit with the space vector control used to minimize the neutral-point voltage offset. The literature [15] considers the problem of the conduction loss caused by the forward bias of the diode when the DNPC circuit outputs a high current, while a three-level T-type neutral point clamped circuit can be used to improve the conduction loss in the DNPC circuit in [13], instead of this, the voltage stresses on some power switches are increased. Study [16] is based on study [15] using two independent power sources with two switches cross-connected to the two half-bridges to achieve seventeen levels of voltage output by switching the cross-connected switches. In [17], a three-phase F-type neutral point clamping circuit is proposed to improve the total standing voltage (TSV) without increasing the number of components in the T-type circuit, thereby taking into consideration the component selection and cost, and a three-phase system is constructed using three sets of the same circuits, which achieves a five-level line-to-line voltage output. By analyzing the DNPC and T-type circuits above, it can be seen that all the power switches of the former are clamped by the neutral voltage, thereby resulting in the voltage stress being half of the input voltage, which has better-switching characteristics and its output specification tends to be high-voltage with a low-current output, while the latter has a smaller number of components, thereby resulting in a reduction in the total voltage drop at turn-on, which has better turn-on characteristics and its output specification tends to be low-voltage with a high-current output. As for the F-type, its output voltage and current specifications are in between the above two circuits, so it is easier to achieve maximum efficiency when operating at a medium load. In [13,14,15,16], two sets of the same circuit topology are independently controlled so that the output voltage levels are the sum of the levels of the two circuits, and the control complexity can be reduced by the same control strategy. However, two independently controlled circuits mean that more than two power switches are turned on and turned off at the same time, thereby leading to an increase in switching loss.
In Figure 1c, the asymmetric type is two different circuits connected together, and the system output comes from their individual terminals. Study [18] discusses how to reduce the number of components to minimize the total harmonic distortion (THD) by constructing the multilevel inverter asymmetrically. Study [19] presents a classic totem pole structure with power factor correction (PFC), where a power switch is used instead of a diode to realize the bidirectional operation of the converter. In particular, the control strategy of the circuit adopts a mix of high and low speeds to drive two switching circuits: one is a low-speed arm half-bridge, which uses power switches with low on-resistance to switch between positive and negative half-cycles of the sinusoidal waveform, and the other is a high-speed arm, which uses a multilevel inverter circuit equipped with a low-voltage stress feature to switch between high-frequency switches, so as to reduce the THD. Therefore, the switching characteristics of the two sets of switching circuits can be distinguished, and accordingly, by differentiating the switching characteristics of the two switching circuits, the power switches can be suitably selected to achieve high efficiency. The references [20,21,22] present DNPC, T-type, and F-Type circuits, respectively. For these circuits, adding a half-bridge circuit for low-speed control and a PWM control for mixing high and low speeds to a three-level neutral point clamping circuit achieves a five-level voltage output. Compared with the symmetrical structure proposed in studies [13,15,17], the totem pole structure can achieve the same voltage output with fewer components and power loss; however, this structure loses the neutral current cancellation feature and increases the neutral voltage offset. Study [23] uses a T-type circuit with low on-state resistance for c and a DNPC circuit with low voltage stress for high-speed control to achieve a five-level voltage output and reduce the neutral voltage offset by sharing the neutral point. In study [24], a three-level flying capacitor circuit is added to a low-frequency switching half-bridge circuit to construct a totem pole structure to achieve five-level voltage output, in which the flying capacitor is self-balanced by high-frequency charging and discharging based on the output current so that the problem of the neutral voltage can be completely solved. Flying capacitor circuits require additional pre-charging or auxiliary circuits to ensure that the voltages across the power switches can be reduced in advance before the circuit operates. The circuit structures shown in [19,20,21,22,23,24] are based on the difference in the characteristics of circuits A and B to create an additional voltage hierarchy. However, the different characteristics of circuits A and B increase the complexity of the control strategy, and the voltage gains of circuits A and B need to be considered to match each other.
In Figure 1d, the hybrid type is three circuits connected together. Circuit B is paralleled with circuit C at their input sides. After this, the output of circuit A is connected with these input sides. Finally, the system output comes from the individual terminals of the two circuits, B and C. Studies [25,26] use an asymmetric five-level circuit in series with a classical H-bridge circuit to enhance the number of levels and the voltage gain. The five-level circuit mentioned in [25] adopts the asymmetric T-type circuit in [23], while the five-level circuit mentioned in [26] adopts the asymmetric five-level flying capacitor circuit in [24], which is connected to a classical H-bridge circuit to realize the nine-level voltage output. Study [27] follows [23] by using the T-type circuit as the module of the series-connected structure. At the same time, the transformer is used to connect the series-connected module to increase the voltage gain and the number of levels. However, the transformer generates a magnetizing current, leading to an increase in losses, and the neutral point voltage of the T-type circuit varies depending on the load. Study [28] realizes the control of real and virtual power injected into the grid. In addition, the circuit structure must incorporate an independent DC power supply of 0.4 times the input voltage to achieve eleven levels of voltage output from two asymmetric voltage sources. The circuit structures shown in [25,26,27,28] can achieve high levels of the voltage output in more than two ways, as mentioned in Figure 1. Among them, the circuit structure presented in [25,26,28] requires an additional independent power supply to increase the voltage gain, thereby making this circuit structure not applicable to single-input DC voltage source systems.
In this paper, a single-phase seven-level symmetric switched-capacitor inverter is based on the concept of the double voltage clamping circuit connected in series with the half-bridge circuit. Above all, only a single DC power supply is used to achieve a single-phase inverter converter with seven levels and a voltage gain of three. The remainder of this paper is organized as follows: Section 2 introduces the proposed circuit structure of the proposed inverter. Section 3 investigates the basic operating principle of this inverter. Section 4 presents the steady-state and small-single analyses. Section 5 presents the design considerations. Section 6 validates the feasibility and effectiveness of such an inverter by simulated and experimental results, respectively. Finally, Section 7 draws a conclusion.

2. Introduction to the Proposed Circuit Structure

The proposed inverter is shown in Figure 2, where switched-capacitor circuits are composed of capacitors and power switches and are often used to maintain the voltages across the capacitors at several times the input DC voltage. The output voltage of the circuit is controlled by switching the active components to charge and discharge the capacitor to achieve a higher output voltage level.
A symmetrical structure multilevel inverter is composed of two sets of identical multilevel inverter circuits in a pairwise manner, and the main difference between the proposed multilevel inverter and the existing multilevel inverter lies in the symmetry of the output structure of the proposed circuit. In general, the existing symmetric multilevel inverter usually has the circuit components with different currents flowing through them between positive and negative half cycles, so different drive control strategies are required, resulting in control difficulty; however, the proposed symmetrical multilevel inverter with currents flowing through the circuit components between positive and negative half-cycles are the same, so only a unipolar drive control strategy is needed to realize the switching control of this inverter, greatly simplifying the system design and control.
The proposed single-phase seven-level symmetric switched-capacitor DC–AC converter is constructed using the concept of two double voltage clamping circuits connected to individual half-bridge circuits. As shown in Figure 2, the output of the half-bridge circuit consisting of two power switches is connected to the load side, which is called the output half bridge; similarly, the double voltage clamping circuit is composed of the half-bridge circuit consisting of two diodes, two capacitors and two power switches, which is called the clamping half bridge.
In addition, due to the symmetry of the output structure of this converter, the currents flowing through the circuit components at positive and negative half-cycles are the same, so only a unipolar drive control strategy is needed, combined with the level shift-pulse width modulation (LS-PWM) to control the whole circuit.

3. Operating Principle and Associated Analysis of the Proposed Inverter

The proposed inverter shown in Figure 2 uses eight switches, four diodes, four clamping capacitors, two filter inductors, one output capacitor, and one output resistor. In addition, switches S1 to S4 and switches S5 to S8 are symmetrical to each other. Therefore, it is only necessary to analyze the operating principle of the positive half-cycle.

3.1. Symbol Definitions and Assumptions

Before analyzing the circuit behavior, the used symbol definitions and the required assumptions are as follows:
(1)
Vin is the input voltage, vo is the output voltage, N is the reference point of zero potential, and Ro is the output resistor;
(2)
Lo1 and Lo2 are the filter inductors, Co is the filter capacitor, and C1 to C4 are the clamping capacitors;
(3)
iL is the current flowing through the filter inductors Lo1 and Lo2, iCo is the current flowing through the filter capacitor Co, and io is the output current;
(4)
S1 to S8 are the active switches and D1 to D4 are the passive switches;
(5)
By assuming that all the values of the clamping capacitors are large enough, the voltages across them can be viewed as constant, i.e., VC1 = VC2 = VC3 = VC4 = Vin;
(6)
It is assumed that all the components are regarded as ideal.

3.2. Operating Principle Analysis

Figure 3 shows the driving signal waveforms for the switches. There are seven states of circuit behavior which correspond to the seven levels of the voltage output. Among them, vgs1 to vgs8 are the driving signals for the switches S1 to S8, respectively. As the voltage output operates at a positive half-cycle, the converter operating between states I and II is mainly performed by the switch S3 to carry out PWM switching to build up the first voltage level, and at this time, the switch S1 is driven by the complementary signal of S3, and the switches S5 and S7 are kept at normal cutoff; the converter operating between states II and III is mainly performed by the switch S1 to carry out PWM switching to establish the second voltage level, and at this time, the switch S3 is kept at normal conduction, while the switches S5 and S7 are kept at normal cutoff; the converter operating between states III and IV is mainly performed by the switch S5, and the same arm switch is driven by inverting the PWM switching of the switch S5 to establish the third voltage level, and at this time, the switches S1 and S3 are kept at normal conduction, while switch S7 is kept at normal cutoff.
In addition, due to the symmetrical output structure of the converter, the circuit components of the currents flowing through the positive and negative half-cycles of the converter are the same, so only a unipolar drive control strategy is needed, combined with the level shift-pulse width modulation (LS-PWM).
State I. [0 < vctrl < u1]: as shown in Figure 4, the control force vctrl, together with triangular waveforms u1, u2, and u3 performs LS-PWM. When the control force vctrl is located between 0 and the triangular waveform u1, the converter enters into state I, as shown in Figure 5. Since the switches S1, S4, S5, and S8 are on and the switches S2, S3, S6, and S7 are off, the current iL flowing through inductors Lo1, Lo2 forms a loop from the input voltage Vin through the switches S1, S4, S5, S8, the capacitor C2, and the diode D4. Therefore, the voltage between points A and B is zero, i.e., vAB = 0.
At state I, according to Kirchhoff’s voltage law (KVL), the following equation can be obtained:
v L ( t ) = L d i L ( t ) d t = 3 R o n i L ( t ) v o ( t ) + V i n V C V F
At state I, according to Kirchhoff’s current law (KCL), the following equation can be obtained as follows:
i C o ( t ) = C o d v o ( t ) d t = i L ( t ) v o ( t ) R o
where Ron is the conduction resistance of the switches S1 to S8, VF is the forward-biased voltage of the diodes D1 to D4, VC is the voltage of the capacitors C1 to C4, inductance L is the sum of inductances Lo1 and Lo2, and iL1, io(t) is the current flowing through the load resistor Ro, i.e., i o ( t ) = v o ( t ) R o .
After combining Equations (1) and (2), the state equation corresponding to state I can be obtained as follows:
d d t x ( t ) = A 1 x ( t ) + B 1 U
where
x ( t ) = ( i L ( t ) v o ( t ) )
U = ( V i n V F V C )
A 1 = ( 3 R o n L 1 L 1 C o 1 R o C o )
B 1 = ( 1 L 1 L 1 L 0 0 0 )
State II. [u1 < vctrl < u2]: As shown in Figure 6, when the control force vctrl is located between the triangular waves u1 and u2, the converter enters into state II. As shown in Figure 7, since the switches S2, S3, S5, and S8 are on and the switches S1, S4, S6, and S7 are off, the inductor current iL flows from the input voltage Vin through the switches S3, S8, and the diodes D1, D4 to form a loop. Therefore, the voltage between points A and B is the input voltage Vin, i.e., vAB = Vin.
At state II, according to Kirchhoff’s voltage law (KVL), the following equation can be obtained:
v L ( t ) = L d i L ( t ) d t = 2 R o n i L ( t ) v o ( t ) + V i n 2 V F
At state II, according to Kirchhoff’s current law (KCL), the following equation can be obtained:
i C o ( t ) = C o d v o ( t ) d t = i L ( t ) v o ( t ) R o
After combining Equations (8) and (9), the state equation corresponding to state II can be obtained as follows:
d d t x ( t ) = A 2 x ( t ) + B 2 U
where
A 2 = ( 2 R o n L 1 L 1 C o 1 R o C o )
B 2 = ( 1 L 2 L 0 0 0 0 )
State III. [u2 < vctrl < u3]: As shown in Figure 8, the converter enters state III when the control force vctrl is located between the triangular waveforms u2 and u3. As shown in Figure 9, since the switches S1, S3, S5 and S8 are on and the switches S2, S4, S6 and S7 are off, the inductor current iL flows from the input voltage Vin through the switches S1, S3, S5, S8, the capacitor C1 and the diode D4 to form a loop. Therefore, the voltage between points A and B is double the input voltage Vin, i.e., vAB = 2 Vin.
At state III, according to Kirchhoff’s voltage law (KVL), the following equation can be obtained:
v L ( t ) = L d i L ( t ) d t = 3 R o n i L ( t ) v o ( t ) + V i n + V C V F
At state III, according to Kirchhoff’s current law (KCL), the following equation can be obtained:
i C o ( t ) = C o d v o ( t ) d t = i L ( t ) v o ( t ) R o
After combining Equations (13) and (14), the state equation corresponding to state III can be obtained:
d d t x ( t ) = A 3 x ( t ) + B 3 U
where
A 3 = ( 3 R o n L 1 L 1 C o 1 R o C o )
B 3 = ( 1 L 1 L 1 L 0 0 0 )
State IV. [u3 < vctrl]: As shown in Figure 10, when the control force vctrl is larger than the triangular wave u3, the converter enters into state IV. As shown in Figure 11, since the switches S1, S3, S6, and S8 are on and the switches S2, S4, S5, and S7 are off, the inductor current iL forms a loop from the input voltage Vin through the switches S1, S3, S6, S8, and the capacitors C1, C4 form a loop. Therefore, the voltage between points A and B is triple the input voltage Vin, i.e., vAB = 3 Vin.
At stage IV, according to Kirchhoff’s voltage law (KVL), the following equation can be obtained:
v L ( t ) = L d i L ( t ) d t = 4 R o n i L ( t ) v o ( t ) + V i n + 2 V C
At state IV, according to Kirchhoff’s current law (KCL), the following equation can be obtained:
i C o ( t ) = C o d v o ( t ) d t = i L ( t ) v o ( t ) R o
After combining Equations (18) and (19), the state equation corresponding to state III can be obtained as follows:
d d t x ( t ) = A 4 x ( t ) + B 4 U
where
A 4 = ( 4 R o n L 1 L 1 C o 1 R o C o )
B 4 = ( 1 L 0 2 L 0 0 0 )

3.3. Switch Behavior and Clamping Capacitor Behavior

Table 1 shows the switching behavior and clamping capacitor charging and discharging behavior, in which ‘1’ and ‘0’ represent the switch on and off, respectively, ‘C’ and ‘D’ represent charging and discharging, respectively, and ‘---’ means no action is taken.

4. Steady-State Analysis and Small-Signal Analysis

In the following, the quiescent operating point and the small-signal model are discussed.

4.1. State-Space Averaging Model

From Figure 4, Figure 6, Figure 8 and Figure 10, it is clear that there are three switching modes when the output voltage operates at a positive half-cycle. The control force vctrl is compared with the triangular wave u1 to make the converter switch between states I and II, and the corresponding duty cycle in state II is Da; the control force vctrl is compared with the triangular wave u2 to make the converter switch between states II and III, and the corresponding duty cycle in state III is Db; and the control force vctrl is compared with the triangular wave u3 to make the converter switch between states III and IV, and the corresponding duty cycle in state IV is Dc. Hence, the associated equations for the duty cycles Da, Db and Dc are shown below:
D a = v c t r l V m ,   0 < v c t r l < V m
D b = v c t r l V m 1 ,   V m < v c t r l < 2 V m
D c = v c t r l V m 2 ,   2 V m < v c t r l < 3 V m
where Vm is the peak-to-peak value of the triangular waves u1, u2, and u3.
The state-space averaging method [29] is used to obtain the average state matrixes for the three switching modes as shown below:
A a = D a A 2 + D a A 1 ,   B a = D a B 2 + D a B 1
A b = D b A 3 + D b A 2 ,   B b = D b B 3 + D b B 2
A c = D c A 4 + D c A 3 ,   B c = D c B 4 + D c B 3
By rearranging (26) to (28), the corresponding mean state matrices of the three switching modes are shown as follows:
A a = ( ( 3 D a ) R o n L 1 L 1 C o 1 R o C o ) ,   B a = ( 1 L 1 D a L 1 + D a L 0 0 0 )
A b = ( ( 2 + D b ) R o n L 1 L 1 C o 1 R o C o ) ,   B b = ( 1 L 2 + D b L D b L 0 0 0 )
A c = ( ( 3 + D c ) R o n L 1 L 1 C o 1 R o C o ) ,   B c = ( 1 L 1 + D c L 1 + D c L 0 0 0 )

4.2. Steady-State Analysis

For the steady-state analysis, considering the small ripple approximation method [29], the state vector x ( t ) = ( i L ( t ) v o ( t ) ) can be rewritten as X = ( I L V o ) , i.e., the differential state vector d d t X = 0 , and by substituting this result into (32), the relationship between input vector U = ( V i n V F V C ) and state vector X = ( I L V o ) can be obtained as follows:
d d t X = A X + B U = 0
X = ( A ) 1 B U
By assuming that the switch is ideal, i.e., Ron = 0, the diode is an ideal diode, i.e., VF = 0, and the voltages across clamping capacitors are equal to the input voltage Vin. Hence, by substituting (29) to (31) and into (33), the state vectors of the three switching modes, as well as the relationship between the large signals IL and Vo and the input voltage Vin for the three switching modes can be obtained as follows:
I L = V o R o = D a V i n R o ,   V o = D a V i n
I L = V o R o = ( 1 + D b ) V i n R o ,   V o = ( 1 + D b ) V i n
I L = V o R o = ( 2 + D c ) V i n R o ,   V o = ( 2 + D c ) V i n
By substituting (23) to (25) into (34) to (36), the relationship between large signals IL and Vo and the control force vctrl is the same for the three switching methods, as shown in (37) and (38):
I L = V o R o = v c t r l V m V i n R o
V o = v c t r l V m V i n

4.3. Small-Signal Analysis

Firstly, the disturbance is added to the duty cycle Da when the output voltage operates at the first voltage level, and the DC and AC components of the state vector and the input vector are defined as follows:
{ D ( t ) = D a + d ˜ ( t ) D ( t ) = D a d ˜ ( t ) ,   { x ( t ) = X + x ˜ ( t ) u ( t ) = U + u ˜ ( t )
By substituting the duty cycle and the disturbance quantity into (26), the average state matrixes A and B of the small-signal AC model can be obtained as follows:
A = D ( t ) A 2 + D ( t ) A 1 = A a + ( A 2 A 1 ) d ˜ ( t ) B = D ( t ) B 2 + D ( t ) B 1 = B a + ( B 2 B 1 ) d ˜ ( t )
where the matrices Aa and Ba are the DC components of the average state matrices, as shown in (26).
After this, by linearizing (40), the following equation can be obtained:
d d t ( X + x ˜ ( t ) ) = ( A a + ( A 2 A 1 ) d ˜ ( t ) ) ( X + x ˜ ( t ) ) + ( B a + ( B 2 B 1 ) d ˜ ( t ) ) ( U + u ˜ ( t ) )
d d t x ˜ ( t ) = A a x ˜ ( t ) + ( ( A 2 A 1 ) X + ( B 2 B 1 ) U ) d ˜ ( t ) + B a u ˜ ( t )
Therefore, in order to obtain the relationship between the disturbance, the input vector change, and the state vectors of the AC components, transforming the state vectors in the time domain to the state vectors in the frequency domain using the Laplace transform of (42) can be achieved.
d S x ˜ ( s ) = A a x ˜ ( s ) + [ ( A 2 A 1 ) X + ( B 2 B 1 ) U ] d ˜ ( s ) + B a u ˜ ( s )
x ˜ ( s ) = ( d S A a ) 1 { [ ( A 2 A 1 ) X + ( B 2 B 1 ) U ] d ˜ ( s ) + B a u ˜ ( s ) }
where
{ d d t x ˜ ( t ) } = d S x ˜ ( s ) ,   d S = ( s 0 0 s )
Accordingly, the transfer function of the disturbance of the duty cycle to the state vector is shown below:
G x d ( s ) = x ˜ ( s ) d ˜ ( s ) | u ˜ ( s ) = 0 = ( d S A a ) 1 [ ( A 2 A 1 ) X + ( B 2 B 1 ) U ] = ( s + ( 3 D a ) R o n L 1 L 1 C o s + 1 R o C o ) 1 ( ( R o n L 0 0 0 ) X + ( 0 1 L 1 L 0 0 0 ) U ) = ( G i d ( s ) G v d ( s ) )
In (46), since the small-signal equation is the state vector composed of two transfer functions, the transfer function of the duty-cycle disturbance to the output voltage is shown below:
G v d ( s ) = V C V F I L R o n s 2 L C o + s ( L R o + ( 3 D a ) R o n C o ) + ( 1 + R o n R o L C o )
Similarly, when the output voltage operates at the second and third voltage levels, the individual transfer functions of the duty-cycle disturbance to the output voltage are shown in (49) and (51):
G x d ( s ) = ( d S A b ) 1 [ ( A 3 A 2 ) X + ( B 3 B 2 ) U ] = ( s + ( 2 + D b ) R o n L 1 L 1 C o s + 1 R o C o ) 1 ( ( R o n L 0 0 0 ) X + ( 0 1 L 1 L 0 0 0 ) U )
G v d ( s ) = V C + V F I L R o n s 2 L C o + s ( L R o + ( 2 + D b ) R o n C o ) + ( 1 + R o n R o L C o )
G x d ( s ) = ( d S A c ) 1 [ ( A 4 A 3 ) X + ( B 4 B 3 ) U ] = ( s + ( 3 + D c ) R o n L 1 L 1 C o s + 1 R o C o ) 1 ( ( R o n L 0 0 0 ) X + ( 0 1 L 1 L 0 0 0 ) U )
G v d ( s ) = V C + V F I L R o n s 2 L C o + s ( L R o + ( 3 + D c ) R o n C o ) + ( 1 + R o n R o L C o )
Since the effect of the parameters Ron and VF can be omitted, (47), (49) and (51) can be unified as follows:
G v d ( s ) = v ˜ o ( s ) d ˜ ( s ) = V i n s 2 L C o + s L R o + 1
In other words, from (52), one can see that the equivalent small-signal AC model of the proposed inverter is the same for any state and is shown in Figure 12. From this figure, it can be known that the small-signal AC model of the proposed inverter is the same as that of the conventional DC–DC buck converter. As a result, the control design can be carried out easily and systematically.

5. Design Considerations

In the following, the specifications given, the components used, and the controller designed are discussed.

5.1. System Configuration

Figure 13 shows the system block diagram of the switching capacitor DC–AC converter proposed in this paper. The circuit system structure consists of the main circuit, the gate drivers, and the output voltage feedback circuit. The operating principle of the main circuit has already been discussed in the previous section, while the gate drivers use UCC21540 isolation drivers. In terms of the voltage feedback circuit, the output voltage vo is divided into a small voltage signal v , the reference voltage vref is subtracted from it and modulated into a pulse signal vpulse using the sampling technique without the analog-to-digital converter (ADC) [30]. Then, the signal vpulse is sent to the field-programmable gate array (FPGA) controller for conversion into a signal vFB. After this, the signal vFB is compensated to obtain gate-driving signals.
Figure 14 shows how the voltage sampling circuit uses a voltage divider, a differential amplifier, a non-inverting amplifier, and a digital-to-analog converter (DAC). The small voltage signal v is first divided by the voltage divider consisting of resistors R1, R4, R5, and R7 for the output voltage vo, and capacitors C5 and C6 are added to filter out the high-frequency noise, and then sent to the differential amplifier consisting of resistors R2, R3, R6, and R8 for the second division. The reference voltage vref is amplified by a non-inverting amplifier consisting of resistors R9 and R10 to amplify the analog signal from the digital-to-analog converter (DAC). Since the sampling voltage range of the no ADC sampling technique [31] is located between 4 V and 6 V, the reference voltage vref is subtracted from the small voltage signal v o to obtain the sampling signal verr, which is then added with a sawtooth voltage vsaw and a 2.5 V DC voltage, and then passed through a comparator to generate a pulse signal vpulse. Afterward, this signal is sent to the FPGA controller to be converted into a signal vFB.

5.2. System Specifications

Table 2 shows the specifications of the switching capacitor DC–AC converter proposed in this paper. The converter proposed in this paper adopts the utility voltage provided by the Taiwan Power Company, i.e., the rms output voltage is 110 V, its output frequency is 60 Hz, and its output power is 500 W. The prototype of the converter can be used in the AC power system for mobile lighting equipment and small household appliances. Table 3 shows the components and parameters of the voltage sampling circuit in this paper.

5.3. Controller Design

The control architecture used in this paper is a digital PI controller, where the feedback output voltage is sampled based on no ADC [30], and then the result is fed into the controller to obtain a control force; it is then used by the LS-PWM module to generate the gate drive signals required for power switches.
Firstly, by substituting the relevant values of the components used in the converter in Table 2 into (52), the duty cycle to output voltage transfer function Gvd of the converter can be obtained as follows:
G v d ( s ) = v ˜ o ( s ) d ˜ ( s ) = 58 s 2 2.794 × 10 10 + s 1.1734 × 10 5 + 1
Next, the uncompensated loop gain equation is analyzed. Here, Gvd is the transfer function of the analog feedback circuit that converts the output voltage vo to a small voltage signal v o in the voltage sampling circuit, as shown in Figure 14. Substituting the component values in Table 3 yields the following:
H ( s ) = 256 × R 3 R 2 × ( R 4 1 + s C 5 + ( R 2 + R 3 ) 1 ) 1 R 1 + ( R 4 1 + s C 5 + ( R 2 + R 3 ) 1 ) 1 = 256 × 18.2 k 115 k × 1 100 k ( ( 10.5 k ) 1 + s 10 n + ( 18.2 k + 115 k ) 1 ) + 1 = 3 . 594 × 1 s 11274 . 56 + 1
Substituting (53) and (54) into (55), the transfer function of the uncompensated loop gain equation can be obtained. The magnitude and phase at 1 kHz can be obtained by setting the crossover frequency to 1 kHz as follows:
T u 0 ( s ) = H ( s ) G v d ( s ) V m = 5 . 687 × 10 4   6 . 925 × 10 9 s 3 + 3 . 689 × 10 4 s 2 + 28 . 06 s + 2 . 794 × 10 5
| T u 0 ( j 2 π 1000 ) | = 14.9377 dB ,   T u 0 ( j 2 π 1000 ) = 33.4439 °
According to the rule of thumb, the general condition for a stable system is that the phase PM is located between 45° and 75°. Therefore, the compensating phase is set to 60°. In (56), the phase that needs to be compensated by the PI controller at 1 kHz is shown below:
60 ° = θ P I 33.4439 ° ( 180 ° ) θ P I = 60 ° 180 ° + 33.4439 ° = 86.5561 °
From (56) and (57), the magnitude and phase of the transfer function of the PI controller at 1 kHz can be determined to calculate the proportional gain of the time-domain system, as shown in (58) and (59), respectively:
G P I ( j 2 π 1000 ) = 5.5832 86.5561 ° = 0.3353 j 5.5731
G P I ( j 2 π 1000 ) = K P + K i j 2 π 1000 { K p = 0.3353 K i = 35016
The digital PI controller uses the discrete-time system to approximate the time-domain system by quickly sampling the values and summing up the errors. In the case of digital controllers, as mentioned in [29], the data delay affects the phase margin in the high-frequency region, and the sampling frequency determines the Nyquist Frequency. The Nyquist Frequency is the lowest frequency that suppresses aliasing. Accordingly, the sampling frequency fsample is set to be four times the switching frequency fs, i.e., the sampling frequency fsample is 234.4 kHz.
The proportional gain Kpz and integral gain Kiz of the digital PI controller use 8-bit precision, i.e., 1 over 256. By adjusting the gain Kpz and Kiz so that the loopback compensation capability of the digital PI controller is equivalent to that of the PI controller of the continuous time system, the relationships between the gains Kp and Ki of the continuous time system and the gains Kpz and Kiz of the discrete-time system are as follows:
{ K p = 0.3353 = 3 × K p z 256 K i = 35016 = 3 × K i z 256 × f s a m p l e
In practice, the gains are fine-tuned, and the resulting gains are set to K p z = 50 and K i z = 12 .

6. Simulated and Experimental Results

After the design parameters of the proposed inverter are put into the circuit constructed by the simulation software PSIM 9.11 to simulate and verify the feasibility of the closed-loop control, the FPGA is used as the digital control kernel to measure some waveforms and data to demonstrate the effectiveness of the design.
Figure 15 shows the output voltage vo and output current io at 100% load. Figure 16 shows the output voltage vo and unfiltered output voltage vAB at 100% load. Figure 17 shows the gate driving signals vgs1, vgs3, vgs5, vgs7 for the switches S1, S3, S5, S7 under 100% load. Figure 18 shows the gate driving signals vgs2, vgs4, vgs6, vgs8 for the switches S2, S4, S6, S8 operating at 100% load. Figure 19 shows the clamping capacitors C1 and C2 operating at 100% load with cross-voltages vC1 and vC2. Figure 20 shows the clamping capacitance C3 and C4 at 100% load and the cross-voltages vC3 and vC4. Figure 21 shows the voltages vds1, vds3, vds5, vds7 on the switches S1, S3, S5, S7 under 100% load. Figure 22 shows the voltages vds2, vds4, vds6, vds8 on the switches S2, S4, S6, S8 operating at 100% load.
From the waveforms in Figure 15, Figure 16, Figure 17 and Figure 18, it can be known that the simulated results are quite consistent with the experimental results, and the converter generates the gate driving signal of the control switch through the positive and negative half-cycles symmetrically, making the switches operated at high frequency yields the unfiltered output voltage vAB, which is then passed through the second-order low-pass filter to obtain the low-harmonic output voltage vo and output current io. In addition, by comparing the measured waveforms in Figure 17 and Figure 18 with the simulated waveforms, it can be shown that due to the consideration of the non-ideal characteristics of the power switch in practice, it is necessary to add the blanking time in the clamping half-bridge and output the half-bridge to avoid short-through. However, when switching from one stage to another, if the proportion of the switching period controlled in that stage is too small, the switching behavior of the power switch is ignored by the blanking time, and the waveforms show that the gate driving signal of the power switch is normally cutoff for part of the switching period. Figure 19 shows the clamping capacitors C1 and C2, which are switched by the diodes D1 and D2, and the switches S1 and S2 to clamp the voltages on C1 and C2 at the input voltage. Figure 20 shows the clamping capacitors C3 and C4, which are switched by the diodes D3 and D4, and the switches S5 and S6 to clamp the voltages on C3 and C4 at the input voltage. As shown in Figure 19 and Figure 20, the clamping capacitor cross-voltage of the double-voltage clamping circuit is the input voltage, and the simulated results are quite consistent with the experimental results. From Figure 21 and Figure 22, it can be shown that not only are the simulated results quite consistent with the experimental results but also the maximum clamping voltages of the switches S3, S4, S7, and S8 are the sum of the two clamping capacitor cross-voltages, i.e., twice the input voltage, whereas the switches S1, S2, S5, and S6 are all clamped at the input voltage, so they only need to withstand double the input voltage.
Figure 23 shows the waveforms of the output voltage vo and output current io from the 10% load to 100% load. Figure 24 shows the waveforms of the output voltage vo and output current io from 100% load to 10% load. Figure 23 shows that when the output voltage is close to the peak of the negative half-cycle, the output voltage and current can still be stabilized after uploading, and the corresponding recovery time is about 1 ms, and the corresponding output voltage drops by about 50 V due to a larger change in the output current. Figure 24 shows when the output voltage is close to the zero-crossing point, vo and io can be stabilized from 100% load to 10% load. Since io has a smaller variation, a change in vo is less clear. In addition, Figure 25 shows the output voltage harmonics under 100% load, whereas Figure 26 shows the output voltage harmonics under 10% load.
Figure 27 shows the efficiency measurement method, where the current detection resistor (Shunt) is used to obtain the input current; two digital meters (Fluke179) are used to detect the voltage on the current detection resistor and to obtain the average value of the input current and the average value of the input voltage. Accordingly, the corresponding average input power can be calculated. In addition, an AC electronic load (Prodigit 3255) is utilized on the load side, and a power analyzer (PM1000+) is adopted to measure the output current, voltage, average output power, total harmonic distortion, and harmonic magnitude of each order. Finally, the resulting input and output powers are employed to calculate the efficiency of the actual circuit operation.
Figure 28 shows the efficiency curve of the proposed circuit. From this figure, the maximum efficiency of the proposed circuit is 97.28%, and the minimum efficiency is 95.11%. Since the converter is constructed using two double voltage clamping circuits connected symmetrically, the switches have low-voltage stress. As a result, the switches are characterized by low-voltage stress for better switching features. However, the circuit needs to store and release energy through the clamping capacitors to achieve a high-voltage gain. Consequently, the series of parasitic resistance (ESR) leads to additional energy loss during the operation period, resulting in a negative correlation between efficiency and output power, i.e., the higher the output power is, the lower the efficiency.

7. Conclusions

In this study, a single-phase, multilevel symmetric switched-capacitor inverter based on the concept of the double voltage clamping circuit connected to the half-bridge circuit was developed. Only one switch operated at any time, so not only was the corresponding control quite easy but also every switch had a very low switching loss and voltage stress. Furthermore, the proposed seven-level inverter was analyzed in detail by the operating principle as well as the mathematical model for the adopted level-shift sinusoidal pulse width modulation (LS-SPWM), which for this multilevel inverter was successfully developed using the well-known state averaging technique widely employed in the DC–DC converter. As a result, the required proportional-integral (PI) controller used in the closed loop can be designed systematically and easily. In addition, the overall efficiency is above 95.11% and up to 97.28%. At the same time, the total harmonic distortion (THD) under 100% load and 10% load are 0.46% and 0.48%, respectively. In addition, in order to meet a higher switching frequency, the switches were realized herein using MOSFETs. However, in motor drive applications, these MOSFETs can be substituted with IGBTs to obtain a lower switching frequency so as to reduce the switching loss as well as the electromagnetic interference (EMI).

Author Contributions

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

Funding

This research was funded by the National Science and Technology Council, Taiwan, under the Grant Number: NSTC 112-2221-E-027-015-MY2.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bp Statistical Review of World Energy 2022. Available online: https://www.bp.com/content/dam/bp/business-sites/en/global/corporate/pdfs/energy-economics/statistical-review/bp-stats-review-2022-full-report.pdf (accessed on 11 January 2023).
  2. IEEE 1547; Standard for Interconnection and Interoperability of Distributed Energy Resources with Associated Electric Power Systems Interfaces. IEEE Standards Association: Piscataway, NJ, USA, 2018.
  3. IEEE 519; Recommended Practices and Requirements for Harmonic Control in Electrical Power Systems. IEEE Standards Association: Piscataway, NJ, USA, 2014.
  4. IEC60038; Standard Voltages. IEEE Standards Association: Piscataway, NJ, USA, 2009.
  5. IEC61000-3-2; Electromagnetic Compatibility (EMC)-Part 3-2: Limits-Limits for Harmonic Current Emissions (Equipment Input Current ≤ 16 A per Phase). IEEE Standards Association: Piscataway, NJ, USA, 2005.
  6. IEC61000-3-3; Electromagnetic Compatibility (EMC)-Part 3-3: Limits-Limitation of Voltage Changes, Voltage Fluctuations and Flicker in Public Low-Voltage Supply Systems, for Equipment with Rated Current ≤ 16 A per Phase and Not Subject to Conditional Connection. IEEE Standards Association: Piscataway, NJ, USA, 2013.
  7. IEC62109-1; Safety of Power Converters for Use in Photovoltaic Power Systems-Part 1: General Requirements. IEEE Standards Association: Piscataway, NJ, USA, 2010.
  8. Koshti, A.K.; Rao, M.N. A brief review on multilevel inverter topologies. In Proceedings of the 2017 International Conference on Data Management, Analytics and Innovation (ICDMAI), Pune, India, 24–26 February 2017; pp. 1–7. [Google Scholar]
  9. Kumar, N.K.; Muthyala, U.R.; Dhabale, A. Optimized multilevel inverters derived from generalized asymmetrical topology. In Proceedings of the 2019 IEEE 4th International Future Energy Electronics Conference (IFEEC), Singapore, 24–28 November 2019; pp. 1–6. [Google Scholar]
  10. Priyadarshi, A.; Kar, P.K.; Karanki, S.B. A single source transformer-less boost multilevel inverter topology with self-voltage balancing. IEEE Trans. Ind. Appl. 2020, 56, 3954–3965. [Google Scholar] [CrossRef]
  11. Escobar, G.; Martinez-Garcia, J.F.; Vazquez-Guzman, G.; Martinez-Rodriguez, P.R.; Valdez-Fernandez, A.A.; Sosa-Zuniga, J.M. Step-up seven-level neutral-point-clamped inverter based topology for TL-PVS. IET Power Electron. 2020, 13, 2847–2853. [Google Scholar]
  12. Dargahi, V.; Abarzadeh, M.; Corzine, K.A.; Enslin, J.H.; Sadigh, A.K.; Rodriguez, J.; Blaabjerg, F.; Maqsood, A. Fundamental circuit topology of duo-active-neutral-point-clamped, duo-neutral-point-clamped, and duo-neutral-point-piloted multilevel converters. IEEE J. Emerg. Sel. Top. Power Electron. 2019, 7, 1224–1242. [Google Scholar] [CrossRef]
  13. Becker, F.; Jamshidpour, E.; Poure, P.; Saadate, S. Modulation strategy with a minimal number of commutations for a five-level H-bridge NPC inverter. Electronics 2019, 8, 454. [Google Scholar] [CrossRef]
  14. Stala1, R. Natural dc-link voltage balance in a single-phase npc inverter with four-level legs and novel modulation method. IET Power Electron. 2020, 13, 3764–3776. [Google Scholar] [CrossRef]
  15. Woldegiorgis, D.; Wei, Y.; Mhiesan, H.; Mantooth, A. New modulation technique for five-level interleaved T-type inverters for switching loss reduction. In Proceedings of the 2020 IEEE Energy Conversion Congress and Exposition (ECCE), Virtual, 11–15 October 2020; pp. 6253–6257. [Google Scholar]
  16. Meraj, S.T.; Hasan, K.; Masaoud, A. A novel configuration of cross-switched T-type (CT-type) multilevel inverter. IEEE Trans. Power Electron. 2020, 35, 3688–3696. [Google Scholar] [CrossRef]
  17. Odeh, C.I.; Lewicki, A.M.M.; Kondratenko, D. Three-level F-type inverter. IEEE Trans. Power Electron. 2021, 36, 11265–11275. [Google Scholar] [CrossRef]
  18. Batool, B.A.; Aliya; Noor, M.; Athar, S.O. A study of asymmetrical multilevel inverter topologies with less number of devices and low THD: A Review. In Proceedings of the 2018 International Conference on Power Generation Systems and Renewable Energy Technologies (PGSRET), Islamabad, Pakistan, 10–12 September 2018; pp. 1–6. [Google Scholar]
  19. Vu, T.T.; Beinarys, R. Feasibility study of compact high-efficiency bidirectional 3-level bridgeless totem-pole PFC/inverter at low cost. In Proceedings of the 2020 IEEE Applied Power Electronics Conference and Exposition (APEC), Long Beach, CA, USA, 15–19 March 2020; pp. 3397–3404. [Google Scholar]
  20. Lopez-Sanchez, M.J.; Pool, E.I.; Cruz-Chan, R.; Escobar, G.; Martinez-Garcia, J.F.; Valdez-Fernandez, A.A.; Villanueva, C. A single phase asymmetrical NPC inverter topology. In Proceedings of the 2016 13th International Conference on Power Electronics (CIEP), Guanajuato, Mexico, 20–23 June 2016. [Google Scholar]
  21. Woldegiorgis, D.; Mantooth, A. Five-level hybrid T-type inverter topology with mixed Si+SiC semiconductor device configuration. In Proceedings of the Thirtheenth Annual Energy Conversion Congress and Exposition, Vancouver, BC, Canada, 10–14 October 2021; pp. 991–996. [Google Scholar]
  22. Afjeh, M.G.; Babaei, M. Model predictive control of a modified five-level inverter with decreased device counts. Int. J. Circuit Theory Appl. 2021, 49, 1987–2006. [Google Scholar] [CrossRef]
  23. Choi, U.-M.; Lee, J.-S. Single-phase five-level IT-Type NPC Inverter with improved efficiency and reliability in photovoltaic systems. IEEE J. Emerg. Sel. Top. Power Electron. 2022, 10, 5226–5239. [Google Scholar] [CrossRef]
  24. Ye, M.; Wei, Q.; Li, S.; Ren, W.; Song, G. Research on balance control strategy of single capacitor clamped five-level inverter clamping capacitor. IET Power Electron. 2021, 14, 280–289. [Google Scholar] [CrossRef]
  25. Rao, B.N.; Suresh, Y.; Naik, B.S.; Venkataramanaiah, J.; Aditya, K.; Panda, A.K. A novel single source multilevel inverter with hybrid switching technique. Int. J. Circuit Theory Appl. 2021, 50, 794–811. [Google Scholar]
  26. Sun, X.; Yun, Z. Hybrid control strategy for a novel asymmetrical multilevel inverter. In Proceedings of the 2010 International Conference on Intelligent System Design and Engineering Application (ISDEA), Changsha, China, 13–14 October 2010; pp. 1–8. [Google Scholar]
  27. Abu Sneineh, A.; Wang, M.-Y. Novel hybrid flying-capacitor-half-bridge 9-level inverter. In Proceedings of the TENCON 2006—2006 IEEE Region 10 Conference, Hong Kong, China, 14–17 November 2006; pp. 1–4. [Google Scholar]
  28. Tayyab, M.; Sarwar, A.; Murshid, S.; Tariq, M.; Al-Durra, A.; Bakhsh, F.I. Active and reactive power control of grid-connected single-phase asymmetrical eleven-level inverter. IET Gener. Transm. Distrib. 2022, 17, 632–644. [Google Scholar] [CrossRef]
  29. Erickson, R.W.; Maksimovic, D. Fundamental Power Electronics, 2nd ed.; Electronic Services: London, UK, 2001. [Google Scholar]
  30. Hwu, K.I.; Yau, Y.T. Applying one-comparator counter-based PWM control strategy to DC-AC converter with voltage reference feedforward control considered. In Proceedings of the 2011 IEEE Applied Power Electronics Conference and Exposition—APEC 2011, Fort Worth, TX, USA, 6–11 March 2011; pp. 1811–1815. [Google Scholar]
  31. Phillips, C.L.; Nagle, H.T. Digital Control System Analysis and Design, 4th ed.; Prentice-Hall Inc.: Raleigh, NC, USA, 2015. [Google Scholar]
Figure 1. Four types of increasing numbers of voltage levels: (a) cascaded; (b) symmetrical; (c) asymmetrical; and (d) hybrid.
Figure 1. Four types of increasing numbers of voltage levels: (a) cascaded; (b) symmetrical; (c) asymmetrical; and (d) hybrid.
Micromachines 15 00766 g001
Figure 2. Circuit structure of the proposed DC–AC converter.
Figure 2. Circuit structure of the proposed DC–AC converter.
Micromachines 15 00766 g002
Figure 3. Driving signal waveforms for the switches.
Figure 3. Driving signal waveforms for the switches.
Micromachines 15 00766 g003
Figure 4. Driving signals for state I.
Figure 4. Driving signals for state I.
Micromachines 15 00766 g004
Figure 5. Current flow for state I.
Figure 5. Current flow for state I.
Micromachines 15 00766 g005
Figure 6. Driving signals for state II.
Figure 6. Driving signals for state II.
Micromachines 15 00766 g006
Figure 7. Current flow for state II.
Figure 7. Current flow for state II.
Micromachines 15 00766 g007
Figure 8. Driving signals for state III.
Figure 8. Driving signals for state III.
Micromachines 15 00766 g008
Figure 9. Current flow for state III.
Figure 9. Current flow for state III.
Micromachines 15 00766 g009
Figure 10. Driving signals for state IV.
Figure 10. Driving signals for state IV.
Micromachines 15 00766 g010
Figure 11. Current flow for state IV.
Figure 11. Current flow for state IV.
Micromachines 15 00766 g011
Figure 12. Equivalent small-signal AC model of the proposed inverter.
Figure 12. Equivalent small-signal AC model of the proposed inverter.
Micromachines 15 00766 g012
Figure 13. System block diagram of the proposed single-phase switched-capacitor multilevel inverter.
Figure 13. System block diagram of the proposed single-phase switched-capacitor multilevel inverter.
Micromachines 15 00766 g013
Figure 14. Voltage sampling circuit.
Figure 14. Voltage sampling circuit.
Micromachines 15 00766 g014
Figure 15. Output voltage and output current under 100% load: (1) vo; (2) io. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 15. Output voltage and output current under 100% load: (1) vo; (2) io. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g015
Figure 16. Output voltage and output current under 100% load: (1) vo; (2) vAB. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 16. Output voltage and output current under 100% load: (1) vo; (2) vAB. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g016
Figure 17. Gate driving signals for S1, S3, S5, and S7 under 100% load: (1) vgs1; (2) vgs3; (3) vgs5; and (4) vgs7. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 17. Gate driving signals for S1, S3, S5, and S7 under 100% load: (1) vgs1; (2) vgs3; (3) vgs5; and (4) vgs7. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g017
Figure 18. Gate driving signals for S2, S4, S6, and S8 under 100% load: (1) vgs2; (2) vgs4; (3) vgs6; and (4) vgs8. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 18. Gate driving signals for S2, S4, S6, and S8 under 100% load: (1) vgs2; (2) vgs4; (3) vgs6; and (4) vgs8. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g018
Figure 19. Voltages on C1 and C2 under 100% load: (1) vC1; (2) vC2. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 19. Voltages on C1 and C2 under 100% load: (1) vC1; (2) vC2. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g019
Figure 20. Voltages on C3 and C4 under 100% load: (1) vC3; (2) vC4. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 20. Voltages on C3 and C4 under 100% load: (1) vC3; (2) vC4. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g020
Figure 21. Voltages across S1, S3, S5, and S7: (1) vds1; (2) vds3; (3) vds5; and (4) vds7. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 21. Voltages across S1, S3, S5, and S7: (1) vds1; (2) vds3; (3) vds5; and (4) vds7. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g021
Figure 22. Voltages across S2, S4, S6, and S8: (1) vds2; (2) vds4; (3) vds6; and (4) vds8. (a) Simulated waveforms, (b) Experimental waveforms.
Figure 22. Voltages across S2, S4, S6, and S8: (1) vds2; (2) vds4; (3) vds6; and (4) vds8. (a) Simulated waveforms, (b) Experimental waveforms.
Micromachines 15 00766 g022
Figure 23. Experimental waveforms close to the peak of the negative half-cycle from the 10% load to 100% load: (1) vo; (2) io.
Figure 23. Experimental waveforms close to the peak of the negative half-cycle from the 10% load to 100% load: (1) vo; (2) io.
Micromachines 15 00766 g023
Figure 24. Experimental waveforms close to the zero-crossing point from 100% load to 10% load: (1) vo; (2) io.
Figure 24. Experimental waveforms close to the zero-crossing point from 100% load to 10% load: (1) vo; (2) io.
Micromachines 15 00766 g024
Figure 25. Output voltage harmonics under 100% load.
Figure 25. Output voltage harmonics under 100% load.
Micromachines 15 00766 g025
Figure 26. Output voltage harmonics under 10% load.
Figure 26. Output voltage harmonics under 10% load.
Micromachines 15 00766 g026
Figure 27. Block diagram for efficiency measurement.
Figure 27. Block diagram for efficiency measurement.
Micromachines 15 00766 g027
Figure 28. Curve of efficiency versus output power.
Figure 28. Curve of efficiency versus output power.
Micromachines 15 00766 g028
Table 1. Switching behavior and clamping capacitor charging and discharging behavior.
Table 1. Switching behavior and clamping capacitor charging and discharging behavior.
StatesSwitchesCapacitorsvANvBNvAB
S1S2S3S4S5S6S7S8C1C2C3C4
I10011001---C---C0 Vin0 Vin0 Vin
II01101001C------C1 Vin01 Vin
III10101001D------C2 Vin02 Vin
IV10100101D------D2 Vin−1 Vin3 Vin
V10010110---C---C0 Vin1 Vin−1 Vin
VI10011010---C---C0 Vin2 Vin−2 Vin
VII01011010------DD−1 Vin2 Vin−3 Vin
Table 2. Specifications and components of the proposed switched-capacitor inverter.
Table 2. Specifications and components of the proposed switched-capacitor inverter.
Specification/ComponentModel Name/Value
DC Input Voltage (Vin)58 V
AC Output Voltage (vo)110 Vrms
Output Frequency (fline)60 Hz
Rated Power (Po,rated)500 W
Clamped Capacitors (C1 to C4)3.3 mF/100 V
Output Filter Capacitor (Co)1 μF/275 V (MKP)
Output filter Inductors (Lo1 and Lo2)142 μH (Ring Core)
Switching Frequency (fs)58.6 kHz
Table 3. Components used in the voltage sampling circuit.
Table 3. Components used in the voltage sampling circuit.
Geometric TypeComponent SymbolModel Name/Value
Four-Channel OPAOPTLV2374
Dip ResistorsR1 and R5100 kΩ
SMD 0805 ResistorsR2 and R6115 kΩ
SMD 0805 ResistorsR3 and R818.2 kΩ
SMD 0805 ResistorsR4 and R710.5 kΩ
SMD 0805 ResistorsR9 and R102 kΩ
0805 MLCCC5 and C610 nF
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

Shieh, J.-J.; Hwu, K.-I.; Chen, S.-J. High-Performance Multi-Level Inverter with Symmetry and Simplification. Micromachines 2024, 15, 766. https://doi.org/10.3390/mi15060766

AMA Style

Shieh J-J, Hwu K-I, Chen S-J. High-Performance Multi-Level Inverter with Symmetry and Simplification. Micromachines. 2024; 15(6):766. https://doi.org/10.3390/mi15060766

Chicago/Turabian Style

Shieh, Jenn-Jong, Kuo-Ing Hwu, and Sheng-Ju Chen. 2024. "High-Performance Multi-Level Inverter with Symmetry and Simplification" Micromachines 15, no. 6: 766. https://doi.org/10.3390/mi15060766

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