Next Article in Journal
Analysing the Economic Viability of Implicit Demand Response Control of Thermal Energy Storage in Hot Water Tanks
Next Article in Special Issue
A Comprehensive Analysis of the Penetration of Detailed Type 4 Wind Turbine Generators in the Two-Area Benchmark System
Previous Article in Journal
Direct Contact Condensers: A Comprehensive Review of Experimental and Numerical Investigations on Direct-Contact Condensation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of a High-Order Stationary Frame Controller for Two-Level and Three-Level NPC Grid-Connected Inverters

by
Nawaf O. Almatani
1,
Abdullah Ali Alhussainy
1,2,
Sultan Alghamdi
1,2,
Hossam Kotb
3,
Kareem M. AboRas
3,*,
Mahendiran Vellingiri
1,2 and
Muhyaddin Rawa
1,2,*
1
Department of Electrical and Computer Engineering, Faculty of Engineering, K. A. CARE Energy Research and Innovation Center, King Abdulaziz University, Jeddah 21589, Saudi Arabia
2
Smart Grids Research Group, Center of Research Excellence in Renewable Energy and Power Systems, King Abdulaziz University, Jeddah 21589, Saudi Arabia
3
Department of Electrical Power and Machines, Faculty of Engineering, Alexandria University, Alexandria 21544, Egypt
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(24), 9313; https://doi.org/10.3390/en15249313
Submission received: 3 November 2022 / Revised: 29 November 2022 / Accepted: 6 December 2022 / Published: 8 December 2022

Abstract

:
Most grid-connected DC/AC inverters use traditional proportional–integral (PI) controllers in a synchronous frame. In addition to poor disturbance rejection capabilities, these PI controllers also exhibit steady-state errors for sinusoidal reference signals. To address these drawbacks, this article investigates the use of a high-order controller in the stationary frame and then compares it with the standard PI controller. The effectiveness of the high-order controller in the stationary frame has been examined by providing an infinite gain at a resonance frequency. In this work, the design of high-order and PI controllers and tuning instructions are given. Furthermore, both high-order and PI current-controlled two-level and three-level neutral point clamped (NPC) inverters are compared. Various operational conditions are used for the comparison. The high-order controller reduced the total harmonic distortion (THD) of the injected current by 1.15% for the two-level inverter in normal conditions as compared to the PI controller and 0.9% for the three-level NPC inverters. Furthermore, it reduced the THD in balanced abnormal conditions by 0.5% for the two-level inverter and 0.18% for the three-level NPC inverters. However, the dq controller has a lower THD during unbalance and short circuit conditions.

1. Introduction

Recently, there has been a rise in the demand for energy, which has resulted in various difficulties for distribution networks, including instability in the grid and power outages. Because it results in more flexibility, balance, and stability for the grids, the utilization of distributed power generating systems (DPGS) is a feasible solution for these issues. Additionally, it can enhance the management of distribution networks and lower the amount of carbon that is released [1]. Wind turbines and photovoltaic systems as well as energy storage devices such as battery banks and fuel cells as well as active filters are DPGS examples. The output voltage for these systems is typically DC, but it must be converted to AC before it can be discharged to the grid or used to power other loads [2,3].
As a result, inverters play a crucial part in grid connected DPGS because they enable the DPGS to convert the DC voltage and current to AC and then transfer it to the grid. As far as design goes, the three-phase inverter is the most basic two-level inverter. It finds widespread applicability in a diverse range of low-voltage and low-power contexts. In order for the two-level inverter to be able to endure high voltage and high current when used in high-power applications, it requires several switch connections that are connected in series and in parallel. In order to prevent overvoltage failures, all of the switches in the series must be turned on or off at the same time. In addition to this, the two-level inverter needs the utilization of big filters for the reduction in harmonics [4].
The reliability of conventional multilevel converters has been demonstrated in a variety of high-power applications. They are most commonly categorized as neutral-point clamped (NPC), cascaded H-bridge (CHB), or flying capacitor (FC) topologies [5,6]. When compared with the conventional two-level inverters, they are capable of managing more power and voltages at lower switch voltage stress than the conventional inverters. Because of this, both the size of the output filter and the switching frequency can be lowered in a noticeable way [7].
Nevertheless, connecting inverters to the grid is critical because it causes a slew of issues, including grid instability and disturbance, if no suitable controller is in place [8]. As a result, these systems should be able to compensate for grid distortions. In this case, a high-speed controller as well as a compatibility algorithm are required. Furthermore, the controller’s design is critical and significant. Active and reactive power transmission between the grid and DPGS, regulation of DC-link voltage, and grid synchronization are all critical functions of this controller [1]. The current-controlled inverter is the one that is utilized more frequently compared with the voltage-controlled inverter. The control strategy makes use of two cascade loops: an inner current loop that regulates the utility current and an outer power loop that supplies the reference current. There are two basic categories that may be used to categorize current-based controllers, namely, the non-linear control technique and the linear control technique [9].
Grid inverters have integrated non-linear controls such as sliding mode, dead-beat, hysteresis, and predictive control [10]. As one of the predictive regulators, deadbeat control mechanisms are the most commonly used control approach. When the deadbeat controller is properly tuned, it is possible to achieve near-zero tracking error in finite sample steps. When the sampling frequency is increased, this control is more susceptible to mismatches and noise [11,12]. A voltage source inverter can use the hysteresis control approach to create switching signals for inverters by comparing the output utility current to the input reference current. This control method’s advantages include ease of use, independence from the loads they are supposed to control, and an excellent transient response [13]. The ability of the predictive control method in nonlinear control systems is well-known. Predictive control can achieve precise current regulation with low total harmonic distortion; however, it is often difficult to execute in practice. The inverter voltage required to force the utility current to follow a current reference is monitored using this control technique [14,15]. The sliding mode (SM) control approach is popular for non-linear and linear loads. Its remarkable performance has made SM control one of the most extensively used algorithms. High dynamic response, stability, robustness, and ease of implementation are the main advantages of this control method. However, the SM control technique has some disadvantages that are well recognized when it is employed with variable switching frequency [16]. These constraints include control imprecision, significant power losses, and a complicated output filter design [17].
On the other hand, the proportional–integral (PI) controller and the proportional-resonant (PR) controller are both types of linear control techniques that can be used in current-controlled inverters [10]. The PI controller has two drawbacks that are well-known in the industry: The weakness of PI controller performance in the stationary frame, which results in steady state error and a phase shift in the current as compared to the reference [18]. Both drawbacks apply to the controller. This is a result of the integral action not performing effectively with sinusoidal reference [19]. Synchronous frame PI control with voltage feed-forward is a popular option [20]. However, it requires multiple conversions in frames and can be difficult to set up with a low-cost fixed-point digital signal processor [21,22]. The PR controller is a type of current controller that does not have the aforementioned disadvantages and is more suited to working with the sinusoidal references and stationary frame than other controllers do [23]. The PR controller only produces gain at one frequency, called the resonant frequency. At all other frequencies, the controller has almost no gain [21].
This paper focuses on two types of controllers, the PI controller in dq frame and the high-order controller in αβ frame. These two different current controllers have been designed and their performance has been compared under different operational conditions for controlling the two-level inverter as well as the three-level NPC inverters. The comparison between the controllers and topologies is conducted using the simulation platform MATLAB/Simulink.

2. Materials and Methods

2.1. Converter Modeling

A half-bridge is the fundamental topology of any voltage source converter because it is the topology that contains only one leg of the converter. In this section, an average model of a two-level half-bridge converter and a three-level half-bridge NPC converter is developed. This gives us the ability to explain the dynamics of the converter as a function of the modulating signal [16].

2.1.1. Two-Level Half-Bridge Converter Averaged Model

For the purpose of describing the fundamental half-bridge two-level converter seen in Figure 1, we will refer to the state of the switch as s(t). This value will be either 1 if the switch is conducting or 0 if it is blocked, and these values are complementary to one another.
s 1 ( t ) + s 2 ( t ) = 1
when s 1 is on, V t ( t ) is equal to ( V D C 2 ) , but when s 4 is on, it is equal to ( V D C 2 )   when s 2 is on, so V t ( t )   can be expressed as a function of both s 1 ( t )   and s 2 ( t ) as follows:
V t ( t ) = ( V D C 2 ) s 1 ( t ) ( V D C 2 ) s 2
Using the averaging operator [24] to calculate the average of a variable as a function of time:
x ¯ ( t ) = 1 T   t T t x ( τ )   d τ
Additionally, applying the operator to s 1 ( t )   and s 2 ( t ) , it can be concluded that
s ¯ 1 ( t ) = d
s ¯ 2 ( t ) = 1 d
where d is the duty ratio. Substituting (4) and (5) in (2)
V t ¯ ( t ) = ( V D C 2 ) ( 2 d 1 )
In the SPWM strategy, the relationship between the reference signal (m) and the duty ratio is [25]:
m = ( 2 d 1 )
From (6) and (7), it is safe to say that
( t ) = ( V D C 2 ) ( m )

2.1.2. Three-Level Half-Bridge Converter Averaged Model

In the three-level half-bridge shown in Figure 2, the AC-side voltage may be represented for a positive modulating signal m as follows:
V t ( t ) = ( V D C 2 ) s 1 ( t ) + ( 0 )   s 3 ( t )
where s 1 ( t ) + s 3 ( t ) = 1 . In the same way, the AC-side voltage for negative modulating functions is
V t ( t ) = ( V D C 2 ) s 4 ( t ) + ( 0 )   s 2 ( t )
where s 4 ( t ) + s 2 ( t ) = 1 . Equations (9) and (10) can be unified through
V t ( t ) = ( V D C 2 ) s 1 ( t )   s g n ( m ) ( V D C 2 ) s 4 ( t )   s g n ( m )
where the function s g n ( x ) is defined by
s g n ( x ) = { 1 ,   x 0 0 ,   x < 0
Applying the averaging operator to both sides of (9) and (10), over one switching cycle, we reach the conclusion
V t ¯ ( t ) = ( V D C 2 ) m ( t )   ,   m   0
V t ¯ ( t ) = ( V D C 2 ) m ( t )   ,   m   0
Equations (13) and (14) can be unified as
V t ( t ) = [ ( V D C 2 )   s g n ( m ) + ( V D C 2 )   s g n ( m ) ]   m ( t )
Since s g n ( m ) + s g n ( m ) = 1   , Equation (15) can be written as
V t ¯ ( t ) = ( V D C 2 ) m ( t )  
Therefore, it is obvious that (16) is identical to (8), which indicates that the average model can be applied for both converters [25].

2.2. AC Side Control Model of Both Converters

Considering Figure 1 and Figure 2, the dynamics of the AC current are described by
L d d t i ( t ) + R i ( t ) = V t ( t ) V s
However, based on Equations (8) and (16), V t can be controlled by the modulating signal m . Figure 3 shows a control block diagram of the system described by (17).
Equation (17) can be expanded for the three-phase system as follows:
L d i a d t = R i a + V t a V s a
L d i b d t = R i b + V t b V s b
L d i c d t = R i c + V t c V s c

2.3. Controller Modelling and Design [25]

2.3.1. High-Order Stationary Frame αβ Controller [25]

Equations (18)–(20) can be expressed in αβ frame using the Clarke transformation as follows:
L d i α d t = R i α + V t α V s α
L d i β d t = R i β + V t β V s β
Expressing V t α and V t β in terms of m α and m β
L d i α d t = R i α + V D C 2 m α V s α
L d i β d t = R i β + V D C 2 m β V s β
Based on Equations (23) and (24), the control loop shown in Figure 4 is developed.
For real/reactive power controller, the currents i α r e f and i β r e f are generated using the following equations:
i α r e f ( t ) = 2 3 [ V s α V s α 2 + V s β   2     P s r e f ( t ) + V s β V s α 2 + V s β   2     Q s r e f ( t )   ]
i β r e f ( t ) = 2 3 [ V s α V s α 2 + V s β   2     P s r e f ( t ) + V s β V s α 2 + V s β   2     Q s r e f ( t )   ]

2.3.2. High-Order Stationary Frame αβ Compensator Design

First of all, the gain crossover frequency ω   C   of the designed compensator should be determined and it must satisfy the inequality ω   C   < ω   b < 2 ω   C   , where ω   b is the bandwidth of the closed loop which should be very larger than the command signal frequency ω   o   in such a way that ω   b   9 ω   o . However, the loop gain of Figure 4 is dictated by:
l ( s ) = k ( s )   1 L s + R
Since the command is a sinusoidal signal, the compensator should include complex-conjugate poles at the rated frequency (s2 + ω   o   2 ) to ensure zero steady-state error. Figure 5 shows the frequency response of l ( j ω ) in dashed lines after adding the complex-conjugate poles, which contain a phase shift of 90 ° at low frequencies owing to the pole s = R L , which must be canceled out by the compensator k ( s ) by including a zero at the same value. Figure 5 shows the frequency response after the pole cancelation in solid lines.
However, even after the pole cancelation, the phase margin at ω   C   is zero and a lead compensator is needed to add a phase margin of 45 ° at ω   C   to improve the stability. The dashed line in Figure 6 shows the frequency response after adding the lead compensator and it is noted that the system is stable, but the gain is almost constant at low frequencies. In order to exhibit a large gain at low frequencies, a lag compensator needs to be added to the compensator, and the solid lines in Figure 6 show the final compensator frequency response.
The final compensator is:
k ( s ) = h   ( s + R L s 2 + ω   o 2 )   k l e a d ( s )   k l a g ( s )

2.3.3. Synchronous Frame dq Controller [25]

Equations (18)–(20) can be expressed in dq frame using the Park transformation as follows:
L d i d d t = L ω   o   i d R i d + V t d V s d
L d i q d t = L ω   o   i q R i q + V t q V s q
Expressing V t d and V t q in terms of m d and m q
L d i d d t = L ω   o   i d R i d + V D C 2 m d V s d
L d i q d t = L ω   o   i q R i q + V D C 2 m q V s q
Based on (31) and (32), the control loop shown in Figure 7 is developed.
For the real-/reactive-power controller, the currents i d r e f and i q r e f are generated using the following equations:
i d r e f ( t ) = 2 3   V s d P s r e f ( t )
i q r e f ( t ) = 2 3   V s d Q s r e f ( t )

2.3.4. Synchronous Frame dq Compensator Design

For simplicity, the control loop can be simplified as shown in Figure 8. However, an integral-proportional (PI) compensator can follow a DC reference in synchronous frame control, in contrast to stationary frame control where compensators are difficult to adjust and have large dynamic orders. The transfer function of the PI controller is given as follows:
k ( s ) = k p s + k i s
As a result, the loop gain of the simplified loop will be given as follows:
k ( s ) = ( k p L s ) s + k i k p s + R L
From (35), it is noted that k i k p   must be equal to R L in order to cancel the pole effect, which can be achieved by making the constants k p and k i as follows:
k p = L τ
k i = R τ
where τ is the time constant of the closed loop and should be in the range of (0.5–5) ms.

2.3.5. Phase-Locked Loop (PLL)

Unlike the stationary frame, the synchronous frame needs to be synchronized with the grid. Phase Locked-Loop is a technique for calculating the grid angle. The grid voltage is fed into the system, and the grid angle θ and the frequency ω are outputs. For the PLL diagram shown in Figure 9, the input to the Park transformation is the voltage of each of the three phases of the grid (abc), and the outputs are the voltages in the Synchronous Rotating Frame (dq). Since the q voltage should be kept constant at zero, it is measured and compared with the reference value of zero. In order to cancel the component q, a PI controller is used to accelerate the dq frame within this feedback loop [26].

2.3.6. NPC DC Voltage Equalizer

Unlike the two-level inverter, technically the NPC’s key issue is to keep the voltages of the two DC-side capacitors equal and at a predetermined level. Due to system flaws, the voltages of DC capacitors may either significantly diverge during transients or steadily drift during steady-state operation. Voltage drift can be avoided through the use of an external converter to correct the DC component [27,28]. Here, the external converter is instructed on how much current to inject into the midpoint based on the (DC) voltage difference between the two capacitors. The biggest issue with this strategy is that it necessitates extra hardware to generate current, which raises both the cost and complexity of the overall system. However, an alternative approach is the control strategy proposed in [29], which recommends altering the converter switching pattern in order to keep the DC-side voltage at the appropriate value. The control strategy is based on the idea that if the neutral point voltage changes, the control scheme responds by adding a DC offset to the modulating signal, The capacitor balance is restored as a result of the offset creating a current in the neutral point. The controller in Figure 10 is responsible for generating the DC offset. This method is a good way to solve the NPC’s biggest technical problem from an economic point of view.

3. Results and Discussion

The system shown in Figure 11 and Figure 12 with parameters in Table 1 has been studied using the two controllers having the parameters given in Table 2. The performance of the two converters controlled by both controllers is compared in terms of power quality under five cases:
  • Case 1: Balanced voltage grid (normal conditions).
  • Case 2: Balanced voltage dip and swell.
  • Case 3: Unbalanced voltage dip and swell.
  • Case 4: Symmetrical and unsymmetrical faults at PCC point.
  • Case 5: Transient analysis.

3.1. Case 1

Table 3 shows the results of all topologies and controllers under normal conditions.
The output three-phase voltages are balanced and constant as shown in Figure 13, and the injected current is shown in Figure 14, indicating that all possible topologies and controllers are performing well and delivering power as the reference values as shown in Figure 15. However, the injected current in the three-level topology has lower Total Harmonic Distortion (THD) than the two-level topology. On the other hand, the high-order αβ controller outperforms the dq controller. The three-level high order αβ controller converter is superior; however, the other converter did not exceed the standard limits [30].

3.2. Case 2

Table 4 shows the results for all topologies and controllers at 30% balanced voltage dip and 30% balanced voltage swell occurring at 0.1 s.
All topologies and controllers exceeded the THD standard constraints; the pcc voltages and currents in the case of the balanced voltage dip are shown in Figure 16 and Figure 17, respectively. The conventional two-level topology performs better than the three-level topology, and the high-order αβ controller has a lower THD in the injected current than the dq controller. As demonstrated in Figure 18, all controllers successfully followed the power reference even under the voltage dip, despite the fact that the injected power experienced a minor transient dip at the time the voltage dip occurred.
The pcc voltages and currents for balanced voltage swell are depicted in Figure 19 and Figure 20, respectively. Although the three-level topology performs better than the two-level topology, the high-order αβ controller still has a lower THD in the injected current than the dq controller. Additionally, as shown in Figure 21, all controllers maintained the injected power as the reference value after the voltage swell occurred with a short transient swell at the time the voltage swell occurred.

3.3. Case 3

Table 5 shows the result for all topologies and controllers at 30% unbalanced voltage dip and 30% unbalanced voltage swell occurring at 0.1 s.
Identical to the previous case, each topology and each controller failed to meet the THD requirements. The currents and voltages at pcc are shown in Figure 22 and Figure 23 during the unbalanced voltage drop. The conventional two-level topology is superior to the three-level structure, and the THD of the injected current is lower for a high-level αβ controller than for a dq controller. As can be seen in Figure 24, after an unbalanced voltage drop occurred, no control device was able to prevent the injected power from fluctuating around the reference value.
Figure 25 and Figure 26 show the pcc voltages and currents during an unbalanced voltage swell. While the dq controller still has a lower THD in the injected current than the high-level αβ controller, the performance of the three-level structure is superior to that of the two-level structure. Moreover, as can be seen in Figure 27, after an unbalanced voltage swell occurs, the injected power oscillates around the reference value, indicating that no controller is able to keep it stable.

3.4. Case 4

In this case, all converters were subjected to a comprehensive short circuit test. Figure 28 and Figure 29 show the power flow in the symmetrical and unsymmetrical fault cases occurring at 0.1s, respectively. It can be seen that the dq controller has a significantly superior fault performance than the higher-level αβ controller, especially in terms of symmetrical fault power fluctuation. However, under the unsymmetrical fault, both controllers failed to maintain a constant flow of power.

3.5. Case 5

In this case, the coupling between active and reactive power has been investigated for both controllers under a transient analysis. At t = 0.1 s, Pref experiences a step change from 0 to 1 MW, followed by another step change from 1 MW to -1 MW at t = 0.2 s. At t = 0.15s, Qref is exposed to a step change that goes from 0 to 600 MVAR. Figure 30a,b demonstrates that active and reactive powers are not totally decoupled from one other in αβ frame, as indicated by Equations (25) and (26). In contrast, as shown in Figure 30c,d, the active and reactive powers in the dq frame are decoupled from one another when any of them are changed, as indicated by Equations (33) and (34). This is the case even though the performance of both controllers following the reference power is satisfactory.

4. Conclusions

In this paper, the design of both dq and higher-order αβ controllers for a three-level NPC inverter and a two-level conventional grid-connected inverter is presented, and the performances are compared through simulations under different operational conditions. From the simulation results summarized in Table 6, it can be seen that under normal conditions, balanced voltage swell and balanced voltage dip, the higher-order αβ controller obtains a lower THD than the dq controller. The dq controller has a lower THD during unbalanced and short circuit conditions.
From a topological perspective, a three-level NPC inverter injected a current with lower total harmonic distortion (THD) under normal, balanced, and unbalanced voltage swell. THD is lower for both levels in other cases.
In the weak grid, the high-order αβ controller can be investigated for future research. In addition, their performance under unbalanced conditions must be investigated with properly designed additional harmonic compensation.

Author Contributions

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

Funding

The authors acknowledge the Deanship of Scientific Research (DSR) at King Abdulaziz University, Jeddah, Saudi Arabia, for funding this project, under grant no. (RG-19-135-43).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data sources employed for analysis are presented in the text.

Acknowledgments

The Authors also acknowledge the support provided by King Abdullah City for Atomic and Renewable Energy (K.A.CARE) under K.A.CARE-King Abdulaziz University Collaboration Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blaabjerg, F.; Teodorescu, R.; Liserre, M.; Timbus, A.V. Overview of control and grid synchronization for distributed power generation systems. IEEE Trans. Ind. Electron. 2006, 53, 1398–1409. [Google Scholar] [CrossRef] [Green Version]
  2. Dogga, R.; Pathak, M.K. Recent trends in solar PV inverter topologies. Sol. Energy 2019, 183, 57–73. [Google Scholar] [CrossRef]
  3. Chen, Z.; Guerrero, J.M.; Blaabjerg, F. A review of the state of the art of power electronics for wind turbines. IEEE Trans. Power Electron. 2009, 24, 1859–1875. [Google Scholar] [CrossRef]
  4. Albatran, S.; al Khalaileh, A.R.; Allabadi, A.S. Minimizing total harmonic distortion of a two-level voltage source inverter using optimal third harmonic injection. IEEE Trans. Power Electron. 2020, 35, 3287–3297. [Google Scholar] [CrossRef]
  5. Sinha, A.; Jana, K.C.; Das, M.K. An inclusive review on different multi-level inverter topologies, their modulation and control strategies for a grid connected photo-voltaic system. Sol. Energy 2018, 170, 633–657. [Google Scholar] [CrossRef]
  6. Gao, T.; Zhang, Y.; Qi, Z.; Duan, J. Variable switching frequency PWM for three-level NPC converter in DFIG wind turbines. In Proceedings of the 4th International Conference on Power and Energy Technology, ICPET 2022, Beijing, China, 28–31 July 2022; pp. 1244–1248. [Google Scholar] [CrossRef]
  7. Makalesi, D.; Mollahasanoglu, M.; Okumus, I. A review of three phase AC-DC power factor correction converters for electric vehicle fast charging. Eur. J. Sci. Technol. 2022, 32, 663–669. [Google Scholar] [CrossRef]
  8. Ndirangu, K.; Tafti, H.D.; Fletcher, J.E.; Konstantinou, G. Impact of grid voltage and grid-supporting functions on efficiency of single-phase photovoltaic inverters. IEEE J. Photovolt. 2022, 12, 421–428. [Google Scholar] [CrossRef]
  9. Parvez, M.; Elias, M.F.M.; Rahim, N.A.; Blaabjerg, F.; Abbott, D.; Al-Sarawi, S.F. Comparative study of discrete PI and PR controls for single-phase UPS inverter. IEEE Access 2020, 8, 45584–45595. [Google Scholar] [CrossRef]
  10. Parvez, M.; Elias, M.F.M.; Rahim, N.A.; Osman, N. Current control techniques for three-phase grid interconnection of renewable power generation systems: A review. Sol. Energy 2016, 135, 29–42. [Google Scholar] [CrossRef]
  11. Xia, L.; Xu, L.; Yang, Q.; Yu, F.; Zhang, S. An MPC scheme with enhanced active voltage vector region for V2G inverter. Energies 2020, 13, 1312. [Google Scholar] [CrossRef] [Green Version]
  12. Elhassan, G.; Zulkifli, S.A.; Iliya, S.Z.; Bevrani, H.; Kabir, M.; Jackson, R.; Khan, M.H.; Ahmed, M. Deadbeat current control in grid-connected inverters: A comprehensive discussion. IEEE Access 2022, 10, 3990–4014. [Google Scholar] [CrossRef]
  13. Chavali, R.V.; Dey, A.; Das, B. A hysteresis current controller PWM scheme applied to three-level NPC inverter for distributed generation interface. IEEE Trans. Power Electron. 2022, 37, 1486–1495. [Google Scholar] [CrossRef]
  14. Cortés, P.; Kazmierkowski, M.P.; Kennel, R.M.; Quevedo, D.E.; Rodriguez, J. Predictive control in power electronics and drives. IEEE Trans. Ind. Electron. 2008, 55, 4312–4324. [Google Scholar] [CrossRef]
  15. Lunardi, A.; Conde, E.; Assis, J.; Meegahapola, L.; Fernandes, D.A.; Filho, A.J.S. Repetitive predictive control for current control of grid-connected inverter under distorted voltage conditions. IEEE Access 2022, 10, 16931–16941. [Google Scholar] [CrossRef]
  16. Saad, M.O.; Ulasyar, A.; Ali, W.; Zad, H.S.; Ullah, N.; Blazek, V.; Prokop, L.; Misak, S. Design and control of novel grid tied multilevel filter-less inverter using current based sliding mode control. IEEE Access 2022, 10, 115555–115570. [Google Scholar] [CrossRef]
  17. Yan, W.; Hu, J.; Utkin, V.; Xu, L. Sliding mode pulsewidth modulation. IEEE Trans. Power Electron. 2008, 23, 619–626. [Google Scholar] [CrossRef]
  18. Bisht, R.; Bhattarai, R.; Subramaniam, S.; Kamalasadan, S. A novel synchronously rotating reference frame based adaptive control architecture for enhanced grid support functions of single-phase inverters. IEEE Trans. Ind. Appl. 2020, 56, 4288–4298. [Google Scholar] [CrossRef]
  19. Habash, A.; Todeschini, G.; Lynch, J.W.J.; Emin, Z. Harmonic mitigation as ancillary service provided by multiple photovoltaic inverters. In Proceedings of the 2021 IEEE PES Innovative Smart Grid Technologies Europe: Smart Grids: Toward a Carbon-Free Future, ISGT Europe 2021, Espoo, Finland, 18–21 October 2021. [Google Scholar] [CrossRef]
  20. Khajeh, K.G.; Farajizadeh, F.; Solatialkaran, D.; Zare, F.; Yaghoobi, J.; Mithulananthan, N. A full-feedforward technique to mitigate the grid distortion effect on parallel grid-tied inverters. IEEE Trans. Power Electron. 2022, 37, 8404–8419. [Google Scholar] [CrossRef]
  21. Teodorescu, R.; Blaabjerg, F.; Liserre, M.; Loh, P.C. Proportional-resonant controllers and filters for grid-connected voltage-source converters. IEE Proc. Electr. Power Appl. 2006, 153, 750–762. [Google Scholar] [CrossRef] [Green Version]
  22. Xie, B.; Guo, K.; Mao, M.; Zhou, L.; Liu, T.; Zhang, Q.; Hao, G. Analysis and improved design of phase compensated proportional resonant controllers for grid-connected inverters in weak grid. IEEE Trans. Energy Convers. 2020, 35, 1453–1464. [Google Scholar] [CrossRef]
  23. Lin, B.; Peng, L.; Liu, X. Selective pole placement and cancellation for proportional-resonant control design used in voltage source inverter. IEEE Trans. Power Electron. 2022, 37, 8921–8934. [Google Scholar] [CrossRef]
  24. Peng, Q.; Buticchi, G.; Tan, N.M.L.; Guenter, S.; Yang, J.; Wheeler, P. Modeling techniques and stability analysis tools for grid-connected converters. IEEE Open J. Power Electron. 2022, 3, 450–467. [Google Scholar] [CrossRef]
  25. Yazdani, A.; Reza, I. Voltage Source Converter in Power System; IEEE Press: Piscataway, NJ, USA; John Wiley: Hoboken, NJ, USA, 2010; pp. 1–473. [Google Scholar]
  26. Sevilmiş, F.; Karaca, H. Performance enhancement of DSOGI-PLL with a simple approach in grid-connected applications. Energy Rep. 2022, 8, 9–18. [Google Scholar] [CrossRef]
  27. Mishra, M.K.; Joshi, A.; Ghosh, A. Control schemes for equalization of capacitor voltages in neutral clamped shunt compensator. IEEE Trans. Power Deliv. 2003, 18, 538–544. [Google Scholar] [CrossRef] [Green Version]
  28. Lee, D.H.; Lee, S.R.; Lee, F.C. An analysis of midpoint balance for the neutral-point-clamped three-level VSI. In Proceedings of the PESC Record—IEEE Annual Power Electronics Specialists Conference, Fukuoka, Japan, 22–22 May 1998; Volume 1, pp. 193–199. [Google Scholar] [CrossRef]
  29. Newton, C.; Sumner, M. Neutral point control for multi-level inverters: Theory, design and operational limitations. In Proceedings of the Conference Record—IAS Annual Meeting (IEEE Industry Applications Society), New Orleans, LA, USA, 5–9 October 1997; Volume 2, pp. 1336–1343. [Google Scholar] [CrossRef]
  30. ANSI/IEEE Std. 519; IEEE Std 519TM-2014: IEEE Recommended Practice and Requirements for Harmonic Control. IEEE: Piscataway, NJ, USA, 2014; Volume 2014, pp. 5–9.
Figure 1. Two-level half-bridge converter.
Figure 1. Two-level half-bridge converter.
Energies 15 09313 g001
Figure 2. Three-level half-bridge converter.
Figure 2. Three-level half-bridge converter.
Energies 15 09313 g002
Figure 3. Control model of both converters.
Figure 3. Control model of both converters.
Energies 15 09313 g003
Figure 4. Current control loop in αβ frame.
Figure 4. Current control loop in αβ frame.
Energies 15 09313 g004
Figure 5. Bode plot of the compensator with and without pole cancelation.
Figure 5. Bode plot of the compensator with and without pole cancelation.
Energies 15 09313 g005
Figure 6. Bode plot of the compensator with and without lag compensation.
Figure 6. Bode plot of the compensator with and without lag compensation.
Energies 15 09313 g006
Figure 7. Current control loop in dq frame.
Figure 7. Current control loop in dq frame.
Energies 15 09313 g007
Figure 8. Simplified current control loop in dq frame.
Figure 8. Simplified current control loop in dq frame.
Energies 15 09313 g008
Figure 9. The PLL control loop.
Figure 9. The PLL control loop.
Energies 15 09313 g009
Figure 10. DC voltage equalizer controller.
Figure 10. DC voltage equalizer controller.
Energies 15 09313 g010
Figure 11. Studied system configurations with two-level inverter using (a) the αβ controller and (b) the dq controller.
Figure 11. Studied system configurations with two-level inverter using (a) the αβ controller and (b) the dq controller.
Energies 15 09313 g011
Figure 12. Studied system configurations with three-level NPC inverter using (a) the αβ controller and (b) the dq controller.
Figure 12. Studied system configurations with three-level NPC inverter using (a) the αβ controller and (b) the dq controller.
Energies 15 09313 g012
Figure 13. Three-phase pcc voltages in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 13. Three-phase pcc voltages in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g013
Figure 14. Three-phase pcc currents in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 14. Three-phase pcc currents in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g014
Figure 15. Power flow in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 15. Power flow in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g015
Figure 16. Three-phase pcc voltages in a balance dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 16. Three-phase pcc voltages in a balance dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g016
Figure 17. Three-phase pcc currents in a balance voltage dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 17. Three-phase pcc currents in a balance voltage dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g017
Figure 18. Power flow in a balanced voltage dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 18. Power flow in a balanced voltage dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g018
Figure 19. Three-phase voltages in a balanced swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 19. Three-phase voltages in a balanced swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g019
Figure 20. Three-phase pcc currents in a balanced voltage swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 20. Three-phase pcc currents in a balanced voltage swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g020
Figure 21. Power flow in a balanced voltage swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 21. Power flow in a balanced voltage swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g021
Figure 22. Three-phase pcc voltages in an unbalance dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 22. Three-phase pcc voltages in an unbalance dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g022
Figure 23. Three-phase pcc currents in an unbalance dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 23. Three-phase pcc currents in an unbalance dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g023
Figure 24. Power flow in an unbalance voltage dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 24. Power flow in an unbalance voltage dip in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g024
Figure 25. Three-phase pcc voltages in an unbalanced swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 25. Three-phase pcc voltages in an unbalanced swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g025
Figure 26. Three-phase pcc currents in an unbalanced swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 26. Three-phase pcc currents in an unbalanced swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g026
Figure 27. Power flow in an unbalanced voltage swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 27. Power flow in an unbalanced voltage swell in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g027
Figure 28. Power flow in symmetrical fault in the case of (a) 2L-αβ, (b)3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 28. Power flow in symmetrical fault in the case of (a) 2L-αβ, (b)3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g028
Figure 29. Power flow in unsymmetrical fault in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 29. Power flow in unsymmetrical fault in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g029
Figure 30. Power under step changes in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Figure 30. Power under step changes in the case of (a) 2L-αβ, (b) 3L-αβ, (c) 2L-dq and (d) 3L-dq.
Energies 15 09313 g030
Table 1. System parameters [25].
Table 1. System parameters [25].
ParameterValue
Vdc1250 V
Grid VL-L RMS480 V
Filter inductor Lf100 µH
Filter Resistor1.19 mΩ
Switch frequency fsw3420 Hz
Pref1 MW
Qref0 VAR
Table 2. Controller parameters.
Table 2. Controller parameters.
dq Controller
ParameterValue
Time constant τ 2 ms
Porpotional gain k p 0.05
Integral gain k i 0.595
αβ controller
ParameterValue
Gain h 1258
Transfer function S + 11.9 S 2 + 377 2
Lead compensator S + 966 S + 5633
Lag compensator S + 2 S + 0.05
NPC DC voltage equalizer
ParameterValue
gain k 0.0014
Table 3. Case 1 results.
Table 3. Case 1 results.
ConverterITHDVTHD
2L-αβ1.33%Less than 1%
2L-dq2.48%Less than 1%
3 L-αβ1.29%Less than 1%
3L-dq2.19%Less than 1%
Table 4. Case 2 results.
Table 4. Case 2 results.
SwellDip
ConverterITHDVTHDITHDVTHD
2L-αβ9.05%7.88%10.46%11.81%
2L-dq9.55%7.88%10.64%11.81%
3 L-αβ8.97%7.88%10.49%11.81%
3L-dq9.15%7.88%10.84%11.81%
Table 5. Case 3 results.
Table 5. Case 3 results.
SwellDip
ConverterITHDVTHDITHDVTHD
2L-αβ10.70%7.88%16.22%11.81%
2L-dq7.76%7.88%8.78%11.81%
3 L-αβ10.61%7.88%16.24%11.81%
3L-dq7.22%7.88%8.93%11.81%
Table 6. Results summary.
Table 6. Results summary.
CaseLower THD TopologyLower THD Controller
Balanced voltage Three-levelαβ controller
Balanced voltage dipTwo-levelαβ controller
Balanced voltage swellThree-levelαβ controller
Unbalanced voltage dipTwo-leveldq controller
Unbalanced voltage swellThree-leveldq controller
Symmetrical faultTwo-leveldq controller
Unsymmetrical faultTwo-leveldq controller
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Almatani, N.O.; Alhussainy, A.A.; Alghamdi, S.; Kotb, H.; AboRas, K.M.; Vellingiri, M.; Rawa, M. Assessment of a High-Order Stationary Frame Controller for Two-Level and Three-Level NPC Grid-Connected Inverters. Energies 2022, 15, 9313. https://doi.org/10.3390/en15249313

AMA Style

Almatani NO, Alhussainy AA, Alghamdi S, Kotb H, AboRas KM, Vellingiri M, Rawa M. Assessment of a High-Order Stationary Frame Controller for Two-Level and Three-Level NPC Grid-Connected Inverters. Energies. 2022; 15(24):9313. https://doi.org/10.3390/en15249313

Chicago/Turabian Style

Almatani, Nawaf O., Abdullah Ali Alhussainy, Sultan Alghamdi, Hossam Kotb, Kareem M. AboRas, Mahendiran Vellingiri, and Muhyaddin Rawa. 2022. "Assessment of a High-Order Stationary Frame Controller for Two-Level and Three-Level NPC Grid-Connected Inverters" Energies 15, no. 24: 9313. https://doi.org/10.3390/en15249313

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