Next Article in Journal
Cost Breakeven Point of Offshore Wind Energy in Brazil
Previous Article in Journal
Optimal Siting and Sizing of Hybrid Energy Storage Systems in High-Penetration Renewable Energy Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Voltage Sensorless Control for a PWM Converter with DSOGI-FLL Under Grid Disturbances

1
Department of Electrical Engineering, Pusan National University, Busan 46241, Republic of Korea
2
Department of Research and Development, Hyowon Powertech Company, Busan 46241, Republic of Korea
3
Department of Electrical and Electronics Engineering, Pusan National University, Busan 46241, Republic of Korea
*
Author to whom correspondence should be addressed.
Energies 2025, 18(9), 2199; https://doi.org/10.3390/en18092199
Submission received: 6 April 2025 / Revised: 18 April 2025 / Accepted: 23 April 2025 / Published: 25 April 2025

Abstract

:
This paper presents Enhanced Voltage Sensorless Control for PWM converter with DSOGI-FLL under grid disturbances. Even without grid voltage sensors, the proposed method accurately estimates the grid angle and voltage, which are necessary for power transfer between the DC link of the PWM converter and the grid. The estimated grid voltage obtained through observer design is separated into positive and negative sequence components, and the grid frequency is estimated using the Dual Second-Order Generalized Integrator Quadrature Signal Generator (DSOGI-QSG) and Dual Second-Order Generalized Integrator Frequency-Locked Loop (DSOGI-FLL). The estimated positive and negative sequence voltages were effectively controlled using a dual current controller. The method operates effectively under normal, balanced AC source conditions and in various grid fault scenarios, including unbalanced voltage, harmonic distortion, voltage sag, and frequency step changes. The validity of the proposed method was evaluated through experimental results by using a grid simulator to implement the fault condition.

1. Introduction

With the increasing deployment of distributed energy resources (DERs) and growing electricity demand, conventional distribution networks are facing significant challenges in maintaining power quality. In particular, bidirectional power flow, along with voltage and frequency fluctuations, is becoming more widespread [1]. Moreover, the growing integration of direct current (DC) loads imposes additional burdens on traditional alternating current (AC) distribution systems [2]. To address these issues, three-phase PWM converters have been implemented in industrial systems, offering bidirectional power regulation and harmonic reduction [3].
Usually, the control scheme of the three-phase PWM converter needs a grid voltage sensor, an AC-line current sensor, and a DC-link voltage sensor. A detailed description of the PWM converter is introduced in [4,5,6]. Different control methods, such as sensorless control, have been proposed to reduce the cost of a PWM converter. In [7,8,9,10,11], a source voltage sensorless control method that estimates source voltages to reduce the source voltage sensors was investigated. In addition to the obvious benefits of sensor reduction, there are several other advantages, like the elimination of noise, resolution limitations, offset, various disturbances related to sensors, and a decrease in hardware complexity [12,13,14]. Figure 1 illustrates the structure of a three-phase source-voltage sensorless PWM converter.
To implement sensorless control, various strategies have been developed. In [15], a Kalman filter was adopted. The nonlinearity of these methods makes control design and stability analysis more difficult. As an alternative, linear estimators have also been used to predict grid voltage. Among these, methods based on a change in variables [16] and virtual flux [17] have been reported. A voltage sensorless controller using a source voltage observer was proposed in [18] to account for grid impedance and abnormal grid voltages. It is effective in handling grid disturbances during steady-state operation. However, it experiences significant current overshoots during grid voltage disturbances.
Generally, when source voltage unbalance occurs in a three-phase PWM converter, it causes DC-link ripples, higher reactive power, and poorer current regulator performance [19]. These conditions result in greater system losses, output current distortion, and electromagnetic interference [20]. Furthermore, the resulting distortions can also lead to estimation errors, as the source voltage observer designed for sensorless operation is based on the mathematical model of the overall system and relies on measured current and frequency for voltage estimation.
To address this problem, control methods are required to separate positive and negative sequences in an unbalanced grid condition. To separate the positive and negative sequences, DSOGI-QSG can be used [21]. A sequence calculator positioned prior to the PLL effectively suppresses low-order harmonics and enables accurate estimation of positive- and negative-sequence components [22]. Although many PLL techniques have been studied, synchronous-reference frame phase-locked loop (SRF-PLL) is arguably the most widely used synchronization technique in three-phase power systems. However, its phase estimation may become inaccurate in the presence of low-order harmonics such as the 5th and 7th. Moreover, it tends to show poor performance under grid fault conditions [23].
According to [3], sensorless control is proposed under unbalanced grid conditions using a voltage observer along with SRF-PLL. However, a voltage observer based on the mathematical model of the overall system has difficulty in accurately estimating grid voltage when harmonics are present due to current distortion. Furthermore, since voltage estimation relies on parameters such as estimated frequency, grid disturbance conditions can lead to estimation errors.
To address similar issues, research on sensorless control has been actively progressing in the motor control field, which shares similarities with grid models. In recent years, PLL and FLL have gained significant attention for their simple and fast response characteristics in position and speed estimation [24,25,26,27]. In [28], a second-order SOGI-frequency-locked loop (SOGI-FLL)-based flux observer for PMSM sensorless control was proposed.
Specifically, the authors of [29] introduced a sensorless observer based on Sliding Mode Observer (SMO) operating with SOGI-FLL. SOGI-FLL enables phase synchronization by providing positive sequence extraction, phase angle estimation, and inherent filtering capabilities. Unlike conventional methods, the proposed control method enables reliable inverter control under various grid conditions, owing to its stable and fast transient characteristics [30,31].
In this paper, a three-phase voltage sensorless PWM converter with DSOGI-FLL is proposed. This method enhanced sensorless control to improve dynamic performance under grid disturbance conditions. Additionally, a full-state source voltage observer based on the stationary reference frame mathematical model of the overall system, along with a dual current controller for controlling the positive and negative sequences, is designed.

2. Conventional Sensorless Control for a Three-Phase PWM Converter

2.1. Source Voltage Observer

Conventional voltage sensorless control [32] for a three-phase PWM converter uses only an AC line current sensor and a DC voltage sensor. By using a full-state source voltage observer, the AC line voltage can be estimated. The following Equation (1) represents the voltage equation for the three-phase PWM converter in the stationary reference frame, where V a b c is the converter output voltage, E ^ a b c is the grid phase estimated voltage, I a b c is the phase current, and R s and L s represent the resistance and inductance of the input reactor, respectively.
V a b c = R s i a b c + L s d i a b c d t + E ^ a b c
When the above Equation (1) is transformed into the stationary reference frame through coordinate transformation, it can be expressed as follows:
V d q s = R s i d q s + L s d i d q s d t + E ^ d q s
When the above Equation (2) is expressed (3) as a state-space equation [32], we obtain
d d x x ^ = [ R s L s 0 1 L s 0 0 R s L s 0 1 L s 0 0 0 ω ^ 0 0 ω ^ 0 ] x ^ + [ 1 L s 0 0 1 L s 0 0 0 0 ] [ V d s V q s ] y = C x = [ 1 0 0 0 0 1 0 0 ] [ i d s i q s E ^ d s E ^ q s ]
where x = [ i d s ,   i q s ,     E ^ d s ,   E ^ q s ] ,   y = [ i d s ,   i q s   ] . When the state-space equation of the system is observable, a full-state source voltage observer is represented (Equation (4)). Figure 2 shows a block diagram of a source voltage observer.
d d t x ^ = [ R s L s 0 1 L s 0 0 R s L s 0 1 L s 0 0 0 ω ^ 0 0 ω ^ 0 ] x ^ + [ 1 L s 0 0 1 L s 0 0 0 0 ] [ V d s V q s ] + [ L 1 L 2 ] ( y y ^ )
Here, L 1 ,   L 2     is the proportional gain of the full-state source voltage observer. If there is no error in the system matrix used in the state observer, the dynamic characteristics of the state variable estimation error are determined by Equation (5).
det [ s I ( A L C ) ] = = det [ s R s L s + l 11 l 12 1 L s 0 l 21 s R s L s + l 22 0 1 L s l 31 l 32 s ω ^ l 41 l 42 ω ^ s ]
The characteristic polynomial that determines the dynamic characteristics of the state variable estimation error can be obtained from Equation (5). Once the desired characteristic polynomial is defined, the coefficients of the two characteristic polynomials can be compared to determine the elements of the gain matrix. Since the target system is a fourth-order system, it is necessary to specify four poles. In the given paper, the poles are determined as conjugate complex double poles, as shown in Equation (6), and the required gain is calculated accordingly.
S 1 , 2 = ς ω o b s + j ω o b s 1 ς 2 S 3 , 4 = ς ω o b s j ω o b s 1 ς 2
C ( s ) = ( s 2 + 2 ς ω o b s s + ω o b s ) 2
When the poles are determined as in Equation (6), the characteristic polynomial is given by Equation (7), and the corresponding gain can be obtained, as shown in Equation (8).
L = [ L 11 L 21 L 31 L 41 L 12 L 22 L 32 L 42 ] T = = [ R s L s + 2 ς ω o b s ω ^ ω ^ R s L s + 2 ς ω o b s ( ω ^ 2 ω o b s 2 ) L s 2 ς ω ^ ω o b s L s 2 ς ω ^ ω o b s L s ( ω ^ 2 ω o b s 2 ) L s ]

2.2. Grid Synchronization

The SRF PLL is likely the most popular type of PLL for grid synchronization techniques. Figure 3 shows a diagram of the SRF PLL. It consists of a standard PI feedback controller with a feedforward nominal grid frequency. The PI controller is used to nullify the quadrature component of the grid voltage E d , ensuring continuous synchronization between the d–q reference frame and the rotating voltage vector [33].

2.3. Grid Synchronization Sensorless Control for a Three-Phase PWM Converter

Figure 4 shows the three-phase PWM converter sensorless control structure using a source voltage observer. The grid voltage is estimated by the observer, and the phase angle is extracted using the SRF-PLL.
However, the conventional voltage sensorless control method has disadvantages. When grid harmonics are present, though the phase angle is tracked precisely, the harmonics are not completely canceled. Moreover, frequency deviations occur, and a small error is obtained in the output signal. Significantly, under an unbalanced grid condition, SRF PLL produces high errors since it can track only the positive sequence component [34].
Hence, though SRF PLL is most suitable under balanced conditions, it is unsuitable for operation under grid disturbance conditions. To overcome this limitation, frequency-locked loop (FLL) has been proposed [35]. Frequency-locked loop (FLL) can estimate the input signal frequency without being influenced by instantaneous changes in phase angle, offering performance advantages over PLL-based methods under unbalanced conditions.

3. Proposed Sensorless Control of a Three-Phase PWM Converter Under Grid Disturbance Conditions

The proposed control method enables sensorless control even in grid disturbance conditions such as unbalanced, harmonics, voltage sag, or frequency sag. To achieve a good dynamic response under grid disturbance conditions, DSOGI-FLL is applied [35]. DSOGI-FLL enables the precise extraction of the positive and negative voltage components by effectively eliminating the 2ω oscillation that occurs under unbalanced voltage conditions. Additionally, it exhibits fast dynamics when the phase angle changes.
In this paper, the voltage sensorless strategy for the proposed three-phase PWM converter is as follows. First, for sensorless control of a three-phase PWM converter under grid disturbance conditions, the source voltage observer individually estimates the positive and negative sequences of the grid voltage. To implement this, DSOGI-QSG is used to separate the positive and the negative sequences of the current and output voltage of the converter. Second, DSOGI-FLL was adopted to improve stability and transient characteristics when the grid is unbalanced or when grid voltage fluctuations, frequency variations, and voltage drops occur. Third, to address grid harmonic conditions, a harmonic compensator is implemented within the current controller. Fourth, Dual Vector Current Control is adopted to implement sensorless control for the PWM converter under grid disturbance conditions. Figure 5 shows the proposed sensorless control of a three-phase PWM converter.
In relation to this study, grid parameters such as resistance and inductance also influence the observer, as they are incorporated into the estimation algorithm. In [36], the impact of parameter variations was analyzed using a state observer, assuming an L-type grid filter. It was demonstrated that the estimation performance remains stable even with a 20% variation in grid inductance. However, since the control method adopted in this study differs from the approach presented in [36], the robustness of the proposed method against parameter variations cannot be ensured. Therefore, this study assumes that the grid parameters are precisely known.

3.1. Source Voltage Observer for Proposed Sensorless Control of Three-Phase PWM Converter

Under grid disturbance conditions, the input of the phase-locked loop fluctuates, which makes the estimated grid angle fluctuate. This fluctuation in the angle leads to an AC-side harmonic current with the current regulation loop of the PWM converter, where a positive and negative sequence voltage observer has been devised.
To address this issue, the positive and negative sequence voltages are estimated separately. By applying the source observer equation, the positive sequence and negative sequence voltages can be separated and defined to estimate them individually as follows. The proposed source voltage observer is represented by the following Equations (9) and (10).
d d t x ^ = [ R s L s 0 1 L s 0 0 R s L s 0 1 L s 0 0 0 ω 0 0 ω 0 ] x ^ + [ 1 L s 0 0 1 L s 0 0 0 0 ] [ V d s + V q s + ] + [ L 1 L 2 ] ( y y ^ )
where x = [ i d s + , i q s + , E d s +   E q s + ] ,   y = [ i d s + , i q s + ]
d d t x ^ = [ R s L s 0 1 L s 0 0 R s L s 0 1 L s 0 0 0 ω 0 0 ω 0 ] x ^ + [ 1 L s 0 0 1 L s 0 0 0 0 ] [ V d s V q s ] + [ L 1 L 2 ] ( y y ^ )
where x = [ i d s , i q s , E d s   E q s ] ,   y = [ i d s , i q s ] .
To achieve this, it is necessary to have the currents and converter voltages separated into positive and negative sequences, which is achieved using DSOGI-QSG, as described in Section 3.2.

3.2. Separation of Positive and Negative Sequences Using DSOGI-QSG for the Proposed Sensorless Control of a Three-Phase PWM Converter

In the previously proposed positive and negative sequence voltage observers, it is essential to separate the positive and negative sequence components of the current. To achieve this, this paper adopts DSOGI-QSG [37].
DSOGI-QSG consists of two quadrature signal generators (QSGs) based on a second-order generalized integrator (SOGI). Figure 6 shows the structure of SOGI, in which k is the damping ratio and ω ^ is the undamped natural frequency. The block diagram of SOGI is expressed in Figure 6, and its transfer functions are given in Equations (11) and (12). In these equations, the bandwidth of SOGI is decoupled from the natural frequency ω ^ , but it is affected by the damping ratio gain k . Additionally, q I and I have a phase difference of 90 degrees. Figure 7 is a bode plot for the above equation. The top part is the magnitude bode plot, and the bottom part is the phase bode plot. From the figure, it can be seen that a lower gain k leads to improved filtering performance but reduces stability in response to frequency variations. At the natural frequency, the amplitude reaches its maximum, and the phase is zero. As the frequency increases, the amplitude gradually decreases.
D ( s ) = I I ( s ) = k ω ^ s s 2 + k ω ^ s + ω 2
Q ( s ) = q I I ( s ) = k ω ^ 2 s 2 + k ω ^ s + ω 2
DSOGI-QSG consists of two SOGI-QSGs, each formed by a second-order generalized integrator with unity feedback. By applying the previously described characteristics of SOGI-QSG, DSOGI-QSG can effectively separate the positive and negative sequence components. To achieve this, the symmetrical components method is applied to extract the respective components from the balanced three-phase current, as shown in Equations (13)–(15).
[ I a + I b + I c + ] = 1 3 [ 1 α α 2 α 2 1 α α α 2 1 ] [ I a I b I c ] ( I a b c + = [ T + ] I a b c )
[ I a I b I c ] = 1 3 [ 1 α 2 α α 1 α 2 α 2 α 1 ] [ I a I b I c ] ( I a b c = [ T ] I a b c )
[ I a 0 I b 0 I c 0 ] = 1 3 [ 1 1 1 1 1 1 1 1 1 ] [ I a I b I c ] ( I a b c 0 = [ T 0 ] I a b c )
where a = e j 2 π 3 = 1 2 + j 3 2 ,   a 2 = e j 4 π 3 = 1 2 j 3 2 .
In this paper, it is assumed that there are no zero sequence components. Therefore, the balanced three-phase current in the stationary reference frame is given by Equation (16).
[ I α I β I 0 ] = 2 3 [ 1 1 2 1 2 0 3 2 3 2 1 2 1 2 1 2 ] [ I a I b I c ] ( I α β 0 = [ T α β 0 ] I a b c )
Based on Equations (13)–(16), the positive and negative sequences in the stationary reference frame are derived as shown in Equations (17) and (18).
I α β + = [ T α β ] I a b c + = [ T α β ] [ T + ] I a b c = [ T α β ] [ T + ] [ T α β ] T I α β = 1 2 [ 1 e j π 2 e j π 2 1 ] I α β
I α β = [ T α β ] I a b c = [ T α β ] [ T ] I a b c = [ T α β ] [ T ] [ T α β ] T I α β = 1 2 [ 1 e j π 2 e j π 2 1 ] I α β
In Equations (17) and (18), e j π 2 represents a 90 °   phase delay. Therefore, by using the characteristics of DSOGI-QSG along with Equations (17) and (18), a block diagram for separating the positive and negative sequence components can be created, as shown in Figure 8.

3.3. DSOGI-QSG-Based Frequency-Locked Loop for the Proposed Sensorless Control of a Three-Phase PWM Converter

To control the three-phase PWM converter, the full-state source voltage observer needs to be used to estimate the phase angle of the grid. In conventional SRF-PLL systems used for grid phase angle synchronization, the grid voltage value is measured to extract the grid phase angle. This method can efficiently extract the grid phase angle and accurately track it even under parameter and frequency variations. However, under grid fault conditions, transient oscillations in the frequency measurement can occur, which may affect DC link control and power factor control. However, DSOGI-FLL is capable of providing stable grid angle estimation with filtering even under grid fault conditions. In addition, it demonstrates reliable frequency tracking performance. Figure 9 shows the DSOGI-QSG-based frequency-locked loop.
In the case of the proposed DSOGI-FLL, it can stably and accurately track grid frequency, even under grid disturbance conditions. The transfer function relating to the input and error signals is presented in Equation (19), and Figure 10 is a bode plot of Equation (19). As shown in Figure 10, the center frequency ω ^ should be aligned with the input frequency ω f f in order to generate a balanced set of in-quadrature outputs with equal amplitudes. To enable automatic tuning of the SOGI structure, it is necessary to analyze the error signal ε v , which is used to regulate the center frequency ω ^ .
E ( s ) = ε v v ^ ( s ) = s 2 + ω ^ 2 s 2 + k ω ^ s + ω ^ 2
It can function as an adaptive filter if an external circuit or algorithm is capable of measuring or detecting this frequency. According to [22], The sign of the frequency error ε f indicates whether the estimated frequency ω ^ is below, equal to, or above the reference frequency ω f f . This sign information is applied by an integral controller with a negative gain γ to continuously adjust ω ^ to match ω f f . According to its operational characteristics, FLL uses γ as its only tuning parameter. To define this parameter, ref. [38] demonstrates that, with gain normalization, FLL exhibits a first-order transfer function. The parameter γ can be selected based on the expected settling time of the system response. In this paper, γ is set to 40. This control provides faster dynamics compared to other methods in the case of a grid fault condition, such as a phase angle jump or transient grid unbalanced conditions [31].

3.4. Current Control Strategy for a Three-Phase PWM Converter Under Grid Disturbance

An unbalanced three-phase voltage without a zero-sequence component can be described by Equations (20) and (21), which are expressed in the positive and negative synchronous reference frames, respectively [39].
E ^ d q p = L s d I d q p d t + R s I d q p + j ω ^ L s I d q p + V d q p
E ^ d q n = L s d I d q n d t + R s I d q n j ω ^ L s I d q n + V d q n
Here, E ^ p d q , E ^ n d q represent the estimated grid voltage using the observer. I ^ p d q , I ^ n d q ,   V ^ p d q , V ^ n d q represents the positive and the negative sequences of components using DSOSI-QSG. Under unbalanced grid conditions, the apparent power can be expressed as follows:
S = ( e j ω ^ t E ^ d q p + e j ω ^ t E ^ d q n ) ( e j ω ^ t I d q p + e j ω ^ t I d q n ) *
Here, the real power P and reactive power Q are expressed as follows. Thus, the active and reactive power can be obtained as shown in [40].
P ( t ) = P 0 + P c 2 cos ( 2 ω t ) + P s 2 sin ( 2 ω t )
Q ( t ) = Q 0 + Q c 2 cos ( 2 ω t ) + Q s 2 sin ( 2 ω t )
where P c 2 , P s 2 ,   Q c 2 , Q s 2   are caused by unbalanced grid conditions.
The real power P is delivered to the DC link and affects the DC voltage level. However, if P c 2 , P s 2 (which are caused by unbalanced grid conditions) are not controlled to zero, P(t) will vary with time, resulting in a DC voltage ripple.
Therefore, a control strategy is required to bring P c 2 , P s 2 to zero. Similarly, to achieve a power factor of 1, control is necessary to reduce Q(t) to zero. This can be expressed in equation form as follows:
[ 2 3 P 0 2 3 Q 0 2 3 P s 2 2 3 P c 2 ] = [ E ^ d p E ^ q p E ^ d n E ^ q n E ^ q p E ^ d p E ^ q n E ^ d n E ^ q n E ^ d n E ^ q p E ^ d p E ^ d n E ^ q n E ^ d p E ^ q p ] [ I d p ( t ) I q p ( t ) I d n ( t ) I q n ( t ) ]
To eliminate DC link voltage ripple and achieve a power factor of 1, the following current reference values can be obtained:
[ I d p ( t ) I q p ( t ) I d n ( t ) I q n ( t ) ] = [ E ^ d p E ^ q p E ^ d n E ^ q n E ^ q p E ^ d p E ^ q n E ^ d n E ^ q n E ^ d n E ^ q p E ^ d p E ^ d n E ^ q n E ^ d p E ^ q p ] 1 [ 2 3 P 0 0 0 0 ] = 2 P 0 3 D [ E ^ d p E ^ q p E ^ d n E ^ q n ]
where   D = [ ( E ^ d p ) 2 + ( E ^ q p ) 2 ] [ ( E ^ d n ) 2 + ( E ^ q n ) 2 ] , and D 0 is assumed. Here, P can be considered as the product of the voltage controller output and the DC link voltage reference value. According to the above equation, the negative-sequence current is included in the case of an unbalanced grid. Therefore, the current controller command can be determined as follows through Equation (27):
[ I d p * ( t ) I q p * ( t ) I d n * ( t ) I q n * ( t ) ] = 2 3 D P * [ E ^ d p E ^ q p E ^ d n E ^ q n ]
Figure 11 shows the control block diagram, which employs two independent current regulators.
In the current controller, the separation of positive and negative sequences is not performed using DSOGI-QSG. If DSOGI-QSG is used for separation, its filtering effect can degrade the transient response characteristics of the current controller. Therefore, the separation of the positive and negative sequences for current control was conducted according to the equations described earlier. For this reason, in grid harmonic conditions, a harmonic compensator is required to compensate the current controller command voltage. Thus, a harmonic compensator consisting of 5th and 7th PR controllers was used [41].

3.5. Grid Synchronization Strategy During Initial Operation

When the converter operation starts, the nominal voltage of the grid is already known; however, the initial grid angle error may result in poor source voltage feedforward. This may invoke a high in-rush current and an undesirable boost of the DC link voltage. Thus, the initial grid angle has to be estimated to prevent overcurrent and DC link overvoltage trip. To initialize grid angle estimation, the output voltage reference of the PWM converter is temporarily set to zero at the start of switching, lasting briefly to avoid triggering overcurrent protection. During this zero-voltage period, the voltage equation can be simplified, as expressed in Equation (28):
E ^ d q s L s d d t i d q s + v d q s   E ^ d q s L s Δ i d q s Δ t
where   Δ t is the PWM zero-voltage period. From this equation, the initial source voltage can be estimated [3], and the corresponding grid angle can be estimated using Equation (29) based on the current measurements.
θ ^ e . initial arctan ( Δ i d s Δ i q s )
Additionally, if there is an initial charging circuit, the initial grid angle can be obtained using Equation (29) through the DSOGI-QSG current flowing during initial charging. However, before errors in the initial grid angle occur, the PWM converter should be operated and controlled using the grid angle obtained from the observer. Therefore, as soon as the initial charging is completed, a signal is received to operate the PWM converter, allowing for proper control. This paper implements control using this method.

4. Experimental Results

To validate the proposed control method, an experiment was performed. The target three-phase PWM converter was connected to an isolated AC power supply (TC.ACS, Regatron, Rorschach, Switzerland) [42] to simulate grid disturbance conditions. The parameters of the experiment system are shown in Table 1.
Table 2 shows the experimental conditions used for experimenting with the three-phase PWM converter control.
The experiments were conducted under balanced grid conditions, unbalanced grid conditions, harmonic conditions, and frequency step changes. The test condition of No. 1 represents the balanced grid condition, which refers to a three-phase balanced system with a line-to-line voltage of 220 Vrms at 60 Hz. For the unbalanced grid condition in No. 2, a voltage deviation of ±20% from the nominal voltage was assumed. In No. 3, harmonic injection was set based on the limits specified in EN 50160 [43], with 5th harmonics at 6% and 7th harmonics at 5%. The frequency step change in No. 4 was assumed to be ±10% of the nominal frequency. For the voltage sag under unbalanced conditions in No. 5, a sag of up to 50% lasting for more than two cycles was assumed.
Figure 12 presents the Equipment Under Test (EUT) and the load equipment utilized in the experiments. During the experiments, an oscilloscope (Model: HDO6054, Firmware Version: 9.2.0.4) from Teledyne LeCroy was used.
Figure 13 shows the estimated initial angle obtained using Equation (29). It can be observed that the estimated grid angle matches the actual grid angle calculated using the grid voltage sensor. It is important to achieve rapid synchronization and stabilization when starting up a PWM converter under a grid voltage sensorless scheme due to the absence of a voltage sensor. As mentioned in Section 3.5 in this paper, the initial value of the grid angle was obtained through the phase of the current flowing during the initial charging.
Figure 14 shows the experimental results of grid voltage and current under balanced grid conditions with a line-to-line voltage of 220 Vrms at 60 Hz. Figure 15 shows the experimental results of the estimated positive source voltage and estimated grid angle. The negative sequence voltage is not attacheds because it does not exist under balanced grid conditions. In Figure 16, the actual grid angle measured by the voltage sensor is compared with the estimated grid angle. The estimated grid angle demonstrates close tracking of the actual angle, with a time difference of approximately 200 μ s .
Figure 17 shows the experimental results under unbalanced grid conditions. The unbalanced grid condition is defined with phase B as the reference, where phase A is reduced by 20%, phase C is increased by 10%, and the system operates at 60 Hz.
Figure 18a shows the estimated positive sequence voltage. Due to the extraction of only the positive sequence under unbalanced conditions, a difference in voltage magnitude is observed. Figure 18b shows the estimated positive sequence voltage and estimated positive grid angle. Figure 18c shows the estimated negative sequence voltage, which is detected under unbalanced conditions.
Figure 19a shows the estimated grid voltage and current in the synchronous reference frame using the conventional control method under unbalanced grid conditions.
As previously discussed, when the positive and negative sequence components are not decoupled and conventional control strategies are adopted under unbalanced grid conditions, ripples are observed in the DC link voltage, grid voltage, and current in the synchronous reference frame. These disturbances result in oscillations in the active power, as shown in Figure 19b. However, as shown in Figure 19c,d, the proposed control method adopted to estimate the positive and negative sequence voltages suppresses active power ripples.
Figure 20 presents the experimental results under harmonic distortion conditions. In Figure 20a, the phase voltage is distorted due to the injection of 5th and 7th harmonics, which leads to a distorted current waveform, as shown in Figure 20c. Figure 21a illustrates both the measured grid voltage and the estimated positive sequence voltage in the stationary dq frame. As shown in Figure 21b, the proposed method is capable of accurately estimating the positive sequence voltage even in the presence of harmonic components in the grid voltage.
Figure 22 shows the experimental results under frequency step changes. Figure 22a,b show the grid voltage frequency change and estimated grid phase angle of the positive sequence when the grid frequency changes. It can be observed that the proposed control method exhibits a fast response to a frequency step change. In Figure 22c,d, the displayed dynamic response of frequency estimation is shown using the proposed DSOGI-FLL and SRF-PLL. Additionally, since no ripple is observed in the DSOGI-FLL output, it facilitates more effective control. In Figure 23, when the frequency changes from 60 Hz to 64 Hz and 56 Hz, the voltage in the state observer is estimated using the updated frequency parameter. Due to the fast dynamic response of DSOGI-FLL, no overshoot is observed during this process.
Figure 24 shows the experimental results under voltage drop and unbalanced grid conditions. At 60 Hz, for a duration of four cycles, phase A decreased by 50%, phase B increased by 10%, and phase C decreased by 35%. In Figure 25b,c, from the moment the voltage drops and unbalanced conditions occur, the estimated negative grid voltage in the stationary dq frame and estimated negative grid angle can be observed. In Figure 26, the estimated positive and negative sequence grid voltage in the synchronous reference frame and the active power are shown when the voltage drops and unbalanced conditions occur.

5. Conclusions

In this paper, an Enhanced Voltage Sensorless Control for a PWM converter with DSOGI-FLL is proposed and analyzed. To respond to grid disturbance, a DSOGI-QSG-based frequency-locked loop was adopted. The performance of the proposed method has been verified through experimental results. The operation of the PWM converter was validated not only under a three-phase balanced grid condition but also under a 20% unbalanced condition, severe harmonic conditions with 5th harmonics at 6% and 7th harmonics at 5%, a 6% frequency step change, and voltage sag conditions with a maximum imbalance of 50%. The experiment was tested under various grid conditions, and the results confirmed its capability for current control and voltage estimation.

Author Contributions

Conceptualization, S.-P.K., D.-Y.K. and J.-M.K.; Methodology, S.-P.K. and J.-M.K.; Software, S.-P.K.; Validation, S.-P.K. and D.-Y.K.; Data curation, S.-P.K. and D.-Y.K.; Writing—original draft, S.-P.K.; Writing—review & editing, D.-Y.K.; Supervision, J.-M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Innovation Program (or Industry Strategic Technology Development Program) (20026552, Development of Electric Propulsion Outboard Engine Technology for Small Boats) funded by the Ministry of Trade, Industry & Energy (MOTIE, Republic of Korea).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Author Seung-Pyo Kang was employed by the Hyowon Powertech Company. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Chen, J.; Zhang, J.; Zhou, J.; Shi, G.; Jia, Y.; Wang, H.; Cai, X. An Enhanced Modular Multilevel Converter with Multiple MVAC Ports Based on Active Fundamental-Frequency Circulating Current Injection to Realize Full-Range Operation. IEEE Trans. Power Electron. 2024, 40, 5423–5439. [Google Scholar] [CrossRef]
  2. Zhang, J.; Li, H.; Kong, X.; Zhou, J.; Shi, G.; Zang, J.; Wang, J. A Novel Multiple-Medium-AC-Port Power Electronic Transformer. IEEE Trans. Ind. Electron. 2024, 71, 6568–6578. [Google Scholar] [CrossRef]
  3. Jung, E.; Kim, M.; Sul, S.K. Control scheme for source voltage sensorless PWM converters under source voltage unbalance. In Proceedings of the 2011 14th European Conference on Power Electronics and Applications, Birmingham, UK, 30 August–1 September 2011; pp. 1–10. [Google Scholar]
  4. Wu, R.; Dewan, S.; Slemon, G. Analysis of an AC-to-DC voltage source converter using PWM with phase and amplitude control. IEEE Trans. Ind. Appl. 1991, 27, 355–364. [Google Scholar] [CrossRef]
  5. Kaura, V.; Blasko, V. Operation of a voltage source converter at increased utility voltage. IEEE Trans. Power Electron. 1997, 12, 132–137. [Google Scholar] [CrossRef]
  6. Blasko, V.; Kaura, V. A new mathematical model and control of a three-phase AC-DC voltage source converter. IEEE Trans. Power Electron. 1997, 12, 116–123. [Google Scholar] [CrossRef]
  7. Ohnishi, T.; Fujii, K. Line voltage sensorless three phase PWM converter by tracking control of operating frequency. In Proceedings of the Power Convers. Conf.—PCC ’97, Nagaoka, Japan, 3–6 August 1997; Volume 1, pp. 247–252. [Google Scholar]
  8. Noguchi, T.; Tomiki, H.; Kondo, S.; Takahashi, I. Direct power control of PWM converter without power source voltage sensors. In Proceedings of the IAS ’96. Conference Record of the 1996 IEEE Industry Applications Conference Thirty-First IAS Annual Meeting, San Diego, CA, USA, 6–10 October 1996; Volume 2, pp. 941–946. [Google Scholar]
  9. Hansen, S.; Malinowski, M.; Blaabjerg, F.; Kazmierkowski, M.P. Sensorless control strategies for PWM rectifier. In Proceedings of the APEC 2000. Fifteenth Annual IEEE Applied Power Electronics Conference and Exposition (Cat. No.00CH37058), New Orleans, LA, USA, 6–10 February 2000; Volume 2, pp. 832–838. [Google Scholar]
  10. Song, H.S.; Joo, I.W.; Nam, K. Source voltage sensorless estimation scheme for PWM rectifiers under unbalanced conditions. IEEE Trans. Ind. Electron. 2003, 50, 1238–1245. [Google Scholar] [CrossRef]
  11. Yoo, H.; Kim, J.H.; Sul, S.K. Sensorless Operation of a PWM Rectifier for a Distributed Generation. IEEE Trans. Power Electron. 2007, 22, 1014–1018. [Google Scholar] [CrossRef]
  12. Ghodke, A.; Chatterjee, K. One-cycle-controlled bidirectional three-phase unity power factor ac–dc converter without having voltage sensors. IET Power Electron. 2012, 5, 1944–1955. [Google Scholar] [CrossRef]
  13. Ketzer, M.B.; Jacobina, C.B. Sensorless control technique for PWM rectifiers with voltage disturbance rejection and adaptive power factor. IEEE Trans. Ind. Electron. 2014, 62, 1140–1151. [Google Scholar] [CrossRef]
  14. Rahoui, A.; Bechouche, A.; Seddiki, H.; Abdeslam, D.O. Grid Voltages Estimation for Three-Phase PWM Rectifiers Control Without AC Voltage Sensors. IEEE Trans. Power Electron. 2017, 33, 859–875. [Google Scholar] [CrossRef]
  15. Lokesh, N.; Mishra, M.K.; Ismail, N.M. Variable Structure Control for Three Phase-Three Wire Nine Switch Converter with LCL Filter. In Proceedings of the 2019 IEEE 13th International Conference on Power Electronics and Drive Systems (PEDS), Toulouse, France, 9–12 July 2019. [Google Scholar]
  16. Wang, B.; Xu, Y.; Shen, Z.; Zou, J.; Li, C.; Liu, H. Current Control of Grid-Connected Inverter with LCL Filter Based on Extended-State Observer Estimations Using Single Sensor and Achieving Improved Robust Observation Dynamics. IEEE Trans. Ind. Electron. 2017, 64, 5428–5439. [Google Scholar] [CrossRef]
  17. Roslan, N.F.; Suul, J.A.; Rocabert, J.; Rodriguez, P. A comparative study of methods for estimating virtual flux at the point of common coupling in grid connected voltage source converters with LCL filter. In Proceedings of the 2016 IEEE Energy Conversion Congress and Exposition (ECCE), Milwaukee, WI, USA, 18–22 September 2016. [Google Scholar]
  18. Lai, N.B.; Baltas, G.N.; Marin, L.; Tarasso, A.; Rodriguez, P. Voltage Sensorless Control for Grid-connected Power Converters based on State Feedback and State Observer. In Proceedings of the 2020 IEEE 21st Workshop on Control and Modeling for Power Electronics (COMPEL), Aalborg, Denmark, 9–12 November 2020. [Google Scholar]
  19. Enjeti, P.N.; Choudhury, S.A. A new control strategy to improve the performance of a PWM AC to DC converter under unbalanced operating conditions. IEEE Trans. Power Electron. 1993, 8, 493–500. [Google Scholar] [CrossRef]
  20. Li, S.; Zhou, J.; Zhou, F.; Niu, F.; Deng, W. A Reduced Current Ripple Overmodulation Strategy for Indirect Matrix Converter. IEEE Trans. Ind. Electron. 2024, 72, 3768–3777. [Google Scholar] [CrossRef]
  21. Peng, L.; Fu, Z.; Xiao, T.; Qian, Y.; Zhao, W.; Zhang, C. An Improved Dual Second-Order Generalized Integrator Phased-Locked Loop Strategy for an Inverter of Flexible High-Voltage Direct Current Transmission Systems under Nonideal Grid Conditions. Processes 2023, 11, 2634. [Google Scholar] [CrossRef]
  22. Nazib, A.A.; Holmes, D.G.; McGrath, B.P. Decoupled DSOGI-PLL for Improved Three Phase Grid Synchronisation. In Proceedings of the 2018 International Power Electronics Conference (IPEC-Niigata 2018 -ECCE Asia), Niigata, Japan, 20–24 May 2018; pp. 3670–3677. [Google Scholar]
  23. Kang, J.-W.; Shin, K.-W.; Lee, H.; Kang, K.-M.; Kim, J.; Won, C.-Y. A Study on Stability Control of Grid Connected DC Distribution System Based on Second Order Generalized Integrator-Frequency Locked Loop (SOGI-FLL). Appl. Sci. 2018, 8, 1387. [Google Scholar] [CrossRef]
  24. Abdelrahem, M.; Hackl, C.M.; Kennel, R. Finite position set-phase locked loop for sensorless control of direct-driven permanent-magnet synchronous generators. IEEE Trans. Power Electron. 2017, 33, 3097–3105. [Google Scholar] [CrossRef]
  25. Ahmed, H.; Amamra, S.-A.; Bierhoff, M.H. Frequency-locked loop-based estimation of single-phase grid voltage parameters. IEEE Trans. Ind. Electron. 2018, 66, 8856–8859. [Google Scholar] [CrossRef]
  26. Wang, G.; Zhan, H.; Zhang, G.; Gui, X.; Xu, D. Adaptive compensation method of position estimation harmonic error for EMF-based observer in sensorless IPMSM drives. IEEE Trans. Power Electron. 2013, 29, 3055–3064. [Google Scholar] [CrossRef]
  27. Xu, W.; Jiang, Y.; Mu, C.; Blaabjerg, F. Improved nonlinear flux observer-based second-order SOIFO for PMSM sensorless control. IEEE Trans. Power Electron. 2018, 34, 565–579. [Google Scholar] [CrossRef]
  28. Sreejith, R.; Singh, B. Sensorless Predictive Current Control of PMSM EV Drive Using DSOGI-FLL Based Sliding Mode Observer. IEEE Trans. Ind. Electron. 2020, 68, 5537–5547. [Google Scholar] [CrossRef]
  29. Kivanc, O.C.; Ozturk, S.B. Sensorless PMSM drive based on stator feedforward voltage estimation improved with MRAS multiparameter estimation. IEEE/ASME Trans. Mechatron. 2018, 23, 1326–1337. [Google Scholar] [CrossRef]
  30. Pinto, J.; Carvalho, A.; Rocha, A.; Araújo, A. Comparison of DSOGI-Based PLL for Phase Estimation in Three-Phase Weak Grids. Electricity 2021, 2, 244–270. [Google Scholar] [CrossRef]
  31. Fuad, K. Grid-Voltage Synchronization Algorithms Based on Phase-Locked Loop and Frequency-Locked Loop for Power Converters. Master’s Thesis, Aalto University, Espoo, Finland, 2014. [Google Scholar]
  32. Jang, J. Sensorless Control of PMSMs for Wide Speed Operation Range. Ph.D. Dissertation, Seoul National University, Seoul, Republic of Korea, 2006. [Google Scholar]
  33. Setiawan, I.; Facta, M.; Priyadi, A.; Purnomo, M.H. Comparison of three popular PLL schemes under balanced and unbalanced grid voltage conditions. In Proceedings of the 2016 8th International Conference on Information Technology and Electrical Engineering (ICITEE), Yogyakarta, Indonesia, 5–6 October 2016. [Google Scholar]
  34. Priyanka, B.; V, S.K.; Lavanya, M.C.; Krishnan V, M. Phase-Locked Loop (PLL) Techniques for Grid Synchronization: A Comprehensive Review. In Proceedings of the 2024 Second International Conference on Emerging Trends in Information Technology and Engineering (ICETITE), Vellore, India, 22–23 February 2024. [Google Scholar]
  35. Rodríguez, P.; Luna, A.; Muñoz-Aguilar, R.S.; Etxeberria-Otadui, I.; Teodorescu, R.; Blaabjerg, F. A Stationary Reference Frame Grid Synchronization System for Three-Phase Grid-Connected Power Converters Under Adverse Grid Conditions. IEEE Trans. Power Electron. 2011, 27, 99–112. [Google Scholar] [CrossRef]
  36. Kim, H.-S.; Kim, K.-H. Voltage-Sensorless Control Scheme for a Grid Connected Inverter Using Disturbance Observer. Energies 2017, 10, 166. [Google Scholar] [CrossRef]
  37. Qiming, C.; Fengren, T.; Jie, G.; Yu, Z.; Deqing, Y. The separation of positive and negative sequence component based on SOGI and cascade DSC and its application at unbalanced PWM rectifier. In Proceedings of the 2017 29th Chinese Control and Decision Conference (CCDC), Chongqing, China, 28–30 May 2017. [Google Scholar]
  38. Rodríguez, P.; Luna, A.; Candela, I.; Mujal, R.; Teodorescu, R.; Blaabjerg, F. Multiresonant frequency-locked loop for grid synchronization of power converters under distorted grid conditions. IEEE Trans. Ind. Electron. 2011, 58, 127–138. [Google Scholar] [CrossRef]
  39. Rioual, P.; Pouliquen, H.; Louis, J.P. Regulation of a PWM rectifier in the unbalanced network state using a generalized model. IEEE Trans. Power Electron. 1996, 11, 495–502. [Google Scholar] [CrossRef]
  40. Song, H.-S.; Nam, K. Dual current control scheme for PWM converter under unbalanced input voltage conditions. IEEE Trans. Ind. Electron. 1999, 46, 953–959. [Google Scholar] [CrossRef]
  41. Park, H.-S.; Heo, H.-J.; Choi, B.-S.; Kim, K.C.; Kim, J.-M. Speed Control for Turbine-Generator of ORC Power Generation System and Experimental Implementation. Energies 2019, 12, 200. [Google Scholar] [CrossRef]
  42. Regatron, Technical Datasheets, TC.ACS.50.528.4WR.S.LC. Programmable Regenerative AC Power Supply. Available online: https://www.regatron.com/product/overview/programmable-bidirectional-ac-power-sources/tc-acs-series/#technical-datasheets (accessed on 1 March 2025).
  43. BS EN 50160:2022; Voltage Characteristics of Electricity Supplied by Public Electricity Networks. European Standards: Brussels, Luxembourg, 2022.
Figure 1. Structure of a grid voltage sensorless PWM voltage source converter.
Figure 1. Structure of a grid voltage sensorless PWM voltage source converter.
Energies 18 02199 g001
Figure 2. Block diagram of a source voltage observer.
Figure 2. Block diagram of a source voltage observer.
Energies 18 02199 g002
Figure 3. Block diagram of the SRF PLL.
Figure 3. Block diagram of the SRF PLL.
Energies 18 02199 g003
Figure 4. Block diagram of the conventional sensorless control for a three-phase PWM converter.
Figure 4. Block diagram of the conventional sensorless control for a three-phase PWM converter.
Energies 18 02199 g004
Figure 5. Block diagram of the proposed control scheme.
Figure 5. Block diagram of the proposed control scheme.
Energies 18 02199 g005
Figure 6. Block diagram of SOGI-QSG.
Figure 6. Block diagram of SOGI-QSG.
Energies 18 02199 g006
Figure 7. Bode plot of SOGI-QSG for different gain k values at a center frequency of 60 Hz.
Figure 7. Bode plot of SOGI-QSG for different gain k values at a center frequency of 60 Hz.
Energies 18 02199 g007
Figure 8. Block diagram of the current DSOGI-QSG.
Figure 8. Block diagram of the current DSOGI-QSG.
Energies 18 02199 g008
Figure 9. Block diagram of the DSOGI-QSG-based frequency-locked loop.
Figure 9. Block diagram of the DSOGI-QSG-based frequency-locked loop.
Energies 18 02199 g009
Figure 10. Block diagram of the voltage DSOGI-QSG-based frequency-locked loop.
Figure 10. Block diagram of the voltage DSOGI-QSG-based frequency-locked loop.
Energies 18 02199 g010
Figure 11. Block diagram of the dual controller.
Figure 11. Block diagram of the dual controller.
Energies 18 02199 g011
Figure 12. Experimental setup with PWM converter.
Figure 12. Experimental setup with PWM converter.
Energies 18 02199 g012
Figure 13. Experimental result for obtaining the initial grid angle.
Figure 13. Experimental result for obtaining the initial grid angle.
Energies 18 02199 g013
Figure 14. Experimental results under balanced grid conditions: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Figure 14. Experimental results under balanced grid conditions: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Energies 18 02199 g014
Figure 15. Experimental results under balanced grid conditions: (a) measured grid voltage and estimated grid voltage in the stationary dq frame; (b) estimated grid voltage in the stationary dq frame and estimated grid angle.
Figure 15. Experimental results under balanced grid conditions: (a) measured grid voltage and estimated grid voltage in the stationary dq frame; (b) estimated grid voltage in the stationary dq frame and estimated grid angle.
Energies 18 02199 g015aEnergies 18 02199 g015b
Figure 16. Experimental results under balanced grid conditions: (a) comparison of estimated grid angles and actual grid angle; (b) detailed view of estimated positive grid angles and actual grid angle.
Figure 16. Experimental results under balanced grid conditions: (a) comparison of estimated grid angles and actual grid angle; (b) detailed view of estimated positive grid angles and actual grid angle.
Energies 18 02199 g016
Figure 17. Experimental results of the grid source voltage and line current under unbalanced grid conditions: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Figure 17. Experimental results of the grid source voltage and line current under unbalanced grid conditions: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Energies 18 02199 g017
Figure 18. Experimental results of the grid source voltage and line current under unbalanced grid conditions: (a) measured grid voltage and estimated positive grid voltage in the stationary dq frame; (b) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle; (c) measured grid voltage and estimated negative grid voltage in the stationary dq frame.
Figure 18. Experimental results of the grid source voltage and line current under unbalanced grid conditions: (a) measured grid voltage and estimated positive grid voltage in the stationary dq frame; (b) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle; (c) measured grid voltage and estimated negative grid voltage in the stationary dq frame.
Energies 18 02199 g018aEnergies 18 02199 g018b
Figure 19. Experimental results under unbalanced grid conditions: (a) grid phase current and estimated positive grid voltage in the synchronous reference frame under conventional sensorless control; (b) active input power under conventional sensorless control; (c) grid phase current and estimated positive grid voltage in the synchronous reference frame under the proposed sensorless control; (d) active input power under proposed sensorless control.
Figure 19. Experimental results under unbalanced grid conditions: (a) grid phase current and estimated positive grid voltage in the synchronous reference frame under conventional sensorless control; (b) active input power under conventional sensorless control; (c) grid phase current and estimated positive grid voltage in the synchronous reference frame under the proposed sensorless control; (d) active input power under proposed sensorless control.
Energies 18 02199 g019aEnergies 18 02199 g019b
Figure 20. Experimental results under harmonic grid condition: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Figure 20. Experimental results under harmonic grid condition: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Energies 18 02199 g020aEnergies 18 02199 g020b
Figure 21. Experimental results under harmonic grid conditions: (a) measured grid voltage and estimated positive grid voltage in the stationary dq frame; (b) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle.
Figure 21. Experimental results under harmonic grid conditions: (a) measured grid voltage and estimated positive grid voltage in the stationary dq frame; (b) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle.
Energies 18 02199 g021
Figure 22. Experimental results under frequency step change conditions: (a) measured grid voltage and estimated positive grid angle when the grid frequency changes to 56 Hz; (b) measured grid voltage and estimated positive grid angle when the grid frequency changes to 64 Hz; (c) measured grid angle and estimated positive grid angle when the grid frequency changes to 56 Hz; (d) measured grid angle and estimated positive grid angle when the grid frequency changes to 64 Hz.
Figure 22. Experimental results under frequency step change conditions: (a) measured grid voltage and estimated positive grid angle when the grid frequency changes to 56 Hz; (b) measured grid voltage and estimated positive grid angle when the grid frequency changes to 64 Hz; (c) measured grid angle and estimated positive grid angle when the grid frequency changes to 56 Hz; (d) measured grid angle and estimated positive grid angle when the grid frequency changes to 64 Hz.
Energies 18 02199 g022
Figure 23. Experimental results under frequency step change conditions: (a) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle when the grid frequency changes from 56 Hz to 60 Hz; (b) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle when the grid frequency changes from 60 Hz to 64 Hz.
Figure 23. Experimental results under frequency step change conditions: (a) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle when the grid frequency changes from 56 Hz to 60 Hz; (b) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle when the grid frequency changes from 60 Hz to 64 Hz.
Energies 18 02199 g023
Figure 24. Experimental results under voltage drop and unbalanced conditions: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Figure 24. Experimental results under voltage drop and unbalanced conditions: (a) measured grid phase voltage; (b) measured grid line-to-line voltage; (c) grid phase current.
Energies 18 02199 g024aEnergies 18 02199 g024b
Figure 25. Experimental results under voltage drop and unbalanced conditions: (a) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle; (b) estimated negative grid voltage in the stationary dq frame; (c) estimated negative grid voltage in the stationary dq frame and estimated negative grid angle.
Figure 25. Experimental results under voltage drop and unbalanced conditions: (a) estimated positive grid voltage in the stationary dq frame and estimated positive grid angle; (b) estimated negative grid voltage in the stationary dq frame; (c) estimated negative grid voltage in the stationary dq frame and estimated negative grid angle.
Energies 18 02199 g025
Figure 26. Experimental results under voltage drop and unbalanced conditions: (a) estimated positive and negative sequence grid voltage in the synchronous reference frame; (b) estimated positive and negative sequence currents in the synchronous reference frame; (c) estimated negative sequence grid voltage in the synchronous reference frame and active input power.
Figure 26. Experimental results under voltage drop and unbalanced conditions: (a) estimated positive and negative sequence grid voltage in the synchronous reference frame; (b) estimated positive and negative sequence currents in the synchronous reference frame; (c) estimated negative sequence grid voltage in the synchronous reference frame and active input power.
Energies 18 02199 g026
Table 1. Parameters of the experiment system.
Table 1. Parameters of the experiment system.
SpecificationValue
Rated Grid Voltage220 Vrms/60 Hz
DC Link Voltage350 V
Interface Inductor2 mH
DC Link Capacitor4700 μF
Switching Frequency5 kHz
Rated Power10 kW
Table 2. Experiment condition.
Table 2. Experiment condition.
No.Test Condition
1Balanced grid condition
2Unbalanced grid condition
3Harmonic injection
4Frequency step change
5Voltage sag with unbalance
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

Kang, S.-P.; Kim, D.-Y.; Kim, J.-M. Enhanced Voltage Sensorless Control for a PWM Converter with DSOGI-FLL Under Grid Disturbances. Energies 2025, 18, 2199. https://doi.org/10.3390/en18092199

AMA Style

Kang S-P, Kim D-Y, Kim J-M. Enhanced Voltage Sensorless Control for a PWM Converter with DSOGI-FLL Under Grid Disturbances. Energies. 2025; 18(9):2199. https://doi.org/10.3390/en18092199

Chicago/Turabian Style

Kang, Seung-Pyo, Dong-Youn Kim, and Jang-Mok Kim. 2025. "Enhanced Voltage Sensorless Control for a PWM Converter with DSOGI-FLL Under Grid Disturbances" Energies 18, no. 9: 2199. https://doi.org/10.3390/en18092199

APA Style

Kang, S.-P., Kim, D.-Y., & Kim, J.-M. (2025). Enhanced Voltage Sensorless Control for a PWM Converter with DSOGI-FLL Under Grid Disturbances. Energies, 18(9), 2199. https://doi.org/10.3390/en18092199

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