Next Article in Journal
An Integrated Approach to Evaluating Eco-Innovation Strategies from the Perspective of Strategic Green Transformation: A Case of the Lithuanian Furniture Industry
Next Article in Special Issue
Critical Technical Issues with a Voltage-Source-Converter-Based High Voltage Direct Current Transmission System for the Onshore Integration of Offshore Wind Farms
Previous Article in Journal
Artificial Intelligence and Urban Green Space Facilities Optimization Using the LSTM Model: Evidence from China
Previous Article in Special Issue
Growth Optimizer for Parameter Identification of Solar Photovoltaic Cells and Modules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of Power System Stabilizers Using Proportional-Integral-Derivative Controller-Based Antlion Algorithm: Experimental Validation via Electronics Environment

by
Nader M. A. Ibrahim
1,*,
Hossam E. A. Talaat
2,
Abdullah M. Shaheen
3 and
Bassam A. Hemade
1
1
Electrical Power System and Machines Department, Faculty of Technology & Education, Suez University, Suez 43512, Egypt
2
Electrical Engineering Department, Future University in Egypt (FUE), Cairo 11835, Egypt
3
Department of Electrical Power Engineering, Faculty of Engineering, Suez University, Suez 43533, Egypt
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(11), 8966; https://doi.org/10.3390/su15118966
Submission received: 23 March 2023 / Revised: 23 April 2023 / Accepted: 4 May 2023 / Published: 1 June 2023
(This article belongs to the Special Issue Sustainable Future of Power System: Estimation and Optimization)

Abstract

:
A robust, optimized power system stabilizer (PSS) is crucial for oscillation damping, and thus improving electrical network stability. Additionally, real-time testing methods are required to significantly reduce the likelihood of software failure in a real-world setting at the user location. This paper presents an Antlion-based proportional integral derivative (PID) PSS to improve power system stability during real-time constraints. The Antlion optimization (ALO) is developed with real-time testing methodology, using hardware-in-the-loop (HIL) that can communicate multiple digital control schemes with real-time signals. The dynamic power system model runs on the dSPACE DS1104, and the proposed PSS runs on the field programmable gate arrays (FPGA) (NI SbRIO-9636 board). The optimized PSS performance was compared with a modified particle swarm optimization (MPSO)-based PID-PSS, through different performance indices. The test cases include other step load perturbations and several short circuit faults at various locations. Twelve different test cases have been applied, through real-time constraints, to prove the robustness of the proposed PSS. These include 5 and 10% step changes through 3 different operating conditions and single, double, and triple lines to ground short circuits through 3 different operating conditions, and at various locations of the system transmission lines. The analysis demonstrates the effectiveness of ALO and MPSO in regaining the system’s stability under the three loading conditions. The integral square of the error (ISE), integral absolute of the error (IAE), integral time square of the error (ITSE), and integral time absolute of the error (ITAE) are used as performance indices in the analysis stage. The simulation results demonstrate the effectiveness of the proposed PSS, based on the ALO algorithm. It provides a robust performance, compared to the traditional PSS. Regarding the applied indices, the proposed PSS, based on the ALO algorithm, obtains significant improvement percentages in ISE, IAE, ITSE, and ITAE with 30.919%, 23.295%, 51.073%, and 53.624%, respectively.

1. Introduction

1.1. Motivation and Incitement

The dynamic stability of an electrical power system is defined as the system’s ability to regain a new equilibrium point after being subjected to a significant disturbance [1,2]. These disturbances adversely affect the system security and power transfer capability, which requires immediate countermeasures to alleviate. Blackouts—such as those notable incidents which have occurred in the United Kingdom (1980), the system separation in the western region of the USA/Canada (1996), the USA (14 August 2003), and Italy (28 September 2003)—involved a low frequency of oscillation (LFO) in the range of 0.1 to 0.7 Hz [3].
These countermeasures, mainly performed by the controlling devices, aim to cancel, or diminish the oscillations associated with disturbances. Typically, modifying the field excitation current reduces the electromechanical changes. However, the power system stabilizer (PSS) provides an additional control signal to the excitation system to enhance its ability to dampen the oscillation and, as a result, improve the system stability [4]. In this regard, the optimal design of the PSS parameters is a critical aspect of enhancing the performance of the modern PSS controller.

1.2. Literature Review

Over the last few years, many techniques have been used to optimize the PSS parameter. In [5], the genetic algorithm (GA) has been implemented to produce a bank of conventional PSS-optimized parameters under different operating points. The adaptive neuro-fuzzy inference system has been activated to select the most suitable parameter for each situation. In [6], the non-dominated sorting genetic algorithm II (NSGAII) has optimized the lead-lag PSS. The study has considered different operating points to solve a multi-objective optimization problem for increasing the damping ratios and shifting the electromechanical modes as much as possible toward the D-shape sector. However, in this model of [6], the power system has been linearized around the steady-state operational condition.
Similarly, in [7], the firefly algorithm has been developed to optimize the PSS parameters. In [8], an advanced population-based incremental learning (PBIL) technique has been combined with an adaptive learning rate and developed to overcome the premature convergence problems in PBIL for designing PSS. In [9], the steepest descent method has been integrated with three metaheuristic algorithms of a gravitational search optimizer (bat optimization technique and PSO) to optimize the PSS parameters. Moreover, the farmland fertility algorithm, which emulates farmers’ behavior when applying various fertilizers to farmlands with varying soil conditions, has been utilized to design the PSS’s proportional integral derivative (PID) controller [10]. In [11], The PSO was modified by converting the search space boundary to work as a reflecting wall. The revised edges prevent any particle from escaping the search space by absorbing part of its velocity by the edge and randomly reflecting it to the search space. In [12], an intelligent hybrid optimization version between the atom search algorithm and simulated annealing has been presented for designing a PSS damping controller.
Robust PSS requires an accurate design and many testing conditions to ensure and validate its robustness. The researchers have investigated the PSS performance through various test cases and diverse system structures. A three-phase short circuit (S.C.) for a six-cycle, through a single-machine infinite bus (SMIB) model, has been covered in [12]. In [13], a 5% step change in mechanical power, across a multi-machine power system (MMPS), has been simulated as a disturbance condition. Different loading conditions have been conducted, with the step load increasing within the MMPS in [14]. Additionally, the effects of a three-phase fault have been analyzed in [15]. The fault lasted for 0.1 s, using a SMIB power system. A 100 ms 3-phase to ground fault, using the MMPS model, was investigated in [16]. On the other hand, offline simulation tests have been analyzed, based on small perturbations within four different system structures (2, 3, 4, and 10 machine systems) in [17], or small step load changes in [18]. In [19], different test scenarios, covering various operating conditions, have been demonstrated to ensure the PSS’s robustness.
For the sake of adequate investigation of the robustness of the optimized PSS in electrical power systems, several performance indices were adopted with different objective functions to improve the optimization process. In [19,20,21], the inverse time absolute square of the error (ITAE) has been utilized. The ITAE is used as a single objective function. The robustness of the developed method was evaluated using a damping ratio, overshot, settling time, and eigenvalues as performance indices. The ITAE-based objective function has also been suggested in [12]. In [22], the minimum-maximum damping ratio and eigenvalues were employed as objective functions where both the damping ratio and eigenvalue performance were used as indices.
On the other hand, considering the adopted simulation environment, all study cases reported in [5,6] have relied on simulation packages. The simulation package-based verification methods have great potential and flexibility, but suffer heavily from biased and unrealistic assumptions imposed by researchers in other cases. In other words, the simulation results are as promising as the researchers are adept at using these packages. However, in [23], experimental validation methodologies were adopted to overcome the insufficiency related to simulation methods. A combination of both strategies was also employed [24]. On the other hand, researchers recently used contemporary technologies and techniques to verify the robustness of their PSS models. Rapid control prototyping (RCP), hardware-in-the-loop (HIL), and software-in-the-loop (SIL) are some of the prevailing techniques [25].
In the last few decades, electrical power networks have evolved and become more complicated. Therefore, relevant verification methodologies and related test devices have also made their share of progress. Real-time technology provides:
  • High-speed processing powers;
  • Higher computation capabilities;
  • Faster operations.;
  • Enhanced performance than other methods.
Compared to simulation studies, these modern technologies can validate newly created procedures where prototyping can considerably improve the entire process. Many simulation assumptions are not applicable in the physical world. Additionally, the simulated model is a simplified representation of the physical system, which means there is no guarantee that the well-optimized model, during simulation, will behave the same way in the field. Such a challenge necessitates more inquiry to match the academic outcomes to the real-world problem, resulting in more sophisticated yet unrealistic solutions [26].
Nowadays, HIL-based verification methodologies have received significant attention from power system researchers [27]. The typical structure of the HIL-based method consists of a real-time controller under investigation and a virtual simulation of the test system [4,28]. Recently, various platforms have been established via HIL, where the most attractive structure was found depending on a field programmable gate array (FPGA), as indicated in [28,29,30,31,32,33]. HIL-based verification methodology produces more tractability and replicability than simulation outcomes. Such advantages encourage researchers to use experimental investigation to validate their developed PSS model [30,34].
On the other hand, the rapid development of electronic chip manufacturing has also led to sophisticated technologies such as RCP. It usually uses a fixed-point processor and can incorporate any digital controller under testing, saving time and effort in the development. Such a technique accelerates the transition stage from research to production [35].
Alternatively, the SIL-based verification procedure utilizes the computational capability of a supercomputer, or vector processor framework, to investigate the developed PSS and the model under investigation on the same framework. SIL does not have any external input/output connections [36]. Modern solutions for experimental verification were adopted in [37], employing the real-time digital simulator platform (RTDS). In this structure, the RCP technology assimilates the newly developed PSS, and the test system under investigation is emulated through a supercomputer with MATLAB/SIMULINK and RTDS. RTDS is an excellent choice for large-scale laboratories and large companies in the power system field.

1.3. Contribution and Paper Organization

This paper gives several contributions to the research field, as described below:
  • Proposing an ALO algorithm to optimize the gains of the PID-based PSS that improves the power system stability during real-time constraints;
  • Presenting experimental validation of the proposed ALO-based PID-PSS controller via a real-time domain, using the dSPACE electronics environment;
  • Several test cases, considering light, nominal, and heavy loadings, are performed, including different step load perturbations and several short circuit faults at various locations;
  • The high effectiveness of ALO and MPSO is demonstrated to regain the system’s stability under the three loading conditions;
  • The proposed ALO-based PID-PSS controller has a robust performance over the MPSO-based PID-PSS.
The rest of the paper is organized as follows: in Section 2, Single machine infinite bus (SMIB) model; in Section 3, Proposed HIL-Based Experimental Model-based Antlion Optimizer (ALO); in Section 4, Results and Analysis in Section 5, Conclusions.

2. SMIB Model

2.1. Block Diagram

The power system model used in this study is a SMIB. The state-space representation of the SMIB model has been summarized. The synchronous generator of the adopted SMIB model is mathematically represented using a fourth-order equation. Figure 1 shows the block diagram representation of the Heffron–Philips SMIB model [38]. The machine has been described in the direct and quadrature axis, while the field circuit is on the direct axis without damper windings. The SMIB model has been used for designing and calculating the constants of the developed PSS. The detailed technical specs of the SMIB model are listed in [38,39]. In this regard, a workbench MATLAB/SIMULINK dynamic model for PSS performance examination is shown in Figure 2 [39]. It illustrates the dynamic test model used throughout this case study [38]. Its dynamics and operating parameters are adapted to fit the model within the HIL method requirements.
The equations of the linear SMIB model without the proposed PID-PSS: -
X 1 = D M X 1 K 1 M X 2 K 2 M X 3
X 2 = ω b X 1
X 3 = K 4 T d o X 2 1 T d o K 3 X 3 + 1 T d o X 4
X 4 = K A K 5 T A X 2 K A K 6 T A X 3 1 T A X 4
where, states:
X = Δ ω Δ δ Δ E q Δ E f d
The above fifth equations are applied only for optimization. In contrast, the testing process depends on the full dynamics.

2.2. Dynamic Model via dSPACE Electronics Environment

In this section, the experimental application of the developed PSS is accomplished by merging dSPACE and NI-single board in a unique configuration. The dynamic model of the SMIB is implemented on the dSPACE board and communicationally linked with the developed PSS, implemented on the NI SbRIO-9636 board via ADC/DAC port. Both boards have FPGAs inside, and accessible communication between the two is guaranteed. Additionally, the MATLAB/SIMULINK power system model’s configuration parameter is adopted, as shown in Figure 3, to effectively perform the dSPACE parameter for real-time links. The real-time controller and the generator 1 (G1) excitation device connection are adopted using MATLAB RTI blocks. The detailed dynamic model has been discussed in [39]. The dynamic test model parameters with the whole states are described in the Appendix A. This model was designed in MATLAB Simulink and then built in the DSPACE board to be used as the test system.
As shown in Figure 3, the path of the signals in Figure 3 can be illustrated in the following order:
  • The block “gain 1” shifts the G1 speed deviation by k;
  • Then, the signal is sent to the developed PSS on the NI-sbRIO-9636 through the DAC converter (DS1104DAC_C1) of dSPACE;
  • The developed PSS receives the real-time signal from dSPACE through the analog pins of the NI-sbRIO-9636, and then compensates for the shift which retained before analyzing the signal;
  • The correction signal was sent back from the PSS after being shifted to the dSPACE through the analog output pin of the NI-SbRIO board;
  • The dSPACE analog input received the correction signal using the ADC converter, represented by the block DS1104ADC_C5.
Considering this, Table 1 reports the I/O characteristics of dSPACE ADC/DAC at different ports.

3. Proposed HIL-Based Experimental Model-based Antlion Optimizer (ALO)

3.1. HIL-Based Experimental Model via dSPACE DS1104 and NI SbRIO-9636 Board

Most implemented case studies push the controller to its limits, which cannot be achieved in the natural operation of physical networks. HIL verification methodology evaluates and validates the effectiveness of the developed controllers by running a set of virtual scenarios in a complex system before real-world implementation [41,42]. HIL simulation blends digital simulation’s speed with analog simulation’s real-time nature [41]. It provides many benefits, including shorter development times, the ability to run multiple tests on practical systems, and experience realization without causing damage to equipment [43]. Figure 4 displays the FPGA platforms of the experimental setup, which include dSPACE DS1104 and NI SbRIO-9636 boards.
Figure 4a demonstrates the essential connections between the power system dynamic model embedded on the dSPACE board, and the PSSs prototype embedded into the FPGA board. The dSPACE R&D real-time hardware platform RTI1104 control board is used widely as an obvious choice to test MATLAB/Simulink models in real-time (its specifications and system parameter modification on a Simulink model are described in Appendix A). The DS1104 I/O capabilities in MATLAB Simulink are available through the rtilib1104 library. The DS1104 has a configurable I/O port that can be programmed graphically, by inserting and connecting the blocks in a Simulink block diagram. Once the code designing is finished, according to the main goals, the MATLAB coder generates the model code in the real-time model, which is then compiled and downloaded on the board to achieve the minimum possible time for developing and implementing a digital controller.
Additionally, there are different configurations for National Instrument reconfigurable I/O boards. This paper uses the NI SbRIO-9636 board to assimilate the developed controller under investigation (its technical specifications are described in Appendix A [44]. As shown in Figure 4b, the board is comprised of an ARM architecture 32-bit microprocessor, compelling programable hardware, and a reconfigurable electronic component (FPGA).

3.2. Antlion Optimizer (ALO)

One of the most contemporary algorithms, rooted in nature, created by Mirjalili in 2015, is the Antlion optimization (ALO) technique. The ALO procedure, which imitates the natural hunting behavior of antlions, has been thoroughly surveyed in [45,46], along with a review of how it is applied to tackle optimizing issues in many domains. It shows great superiority, compared to several other algorithms, and validates several mathematical benchmark functions. The ALO method is appropriate for finding the optimal solution and achieving excellent convergence, because it is straightforward to implement, versatile, scalable, and has a sufficient level of exploring and exploiting abilities. As a result, several application fields—including global optimization, picture segmentation, load balancing, feature selection, and renewable energy integration—have been effectively implemented.
The Antlion optimizer was first proposed in 2015 by S. Mirjalili, miming the hunting behavior of antlions in nature [47]. It is mathematically modeled through two crucial phases in its lifecycle: the larval phase and adulthood. A natural lifespan can be up to 3 years, mostly in the larval stage (only 3–5 weeks are spent in adulthood) [47]. The larval stage is simulated where antlions build cone-shaped traps and then hide in the sand, waiting for prey ants. The size of the trap depends on the antlion’s hunger level and the phase which the moon is in at the time. The antlion purposefully throws sand towards the pit edge to slide the prey into the bottom of the hole. When a prey target falls into the pit, it is pulled under the soil and swallowed. After consuming the prey, antlions throw the remains outside the pit and then reconstruct it for the next hunt [48]. They have evolved and adapted this method to improve their chance of survival. Figure 5 depicts the flow chart of the ALO and highlights its mathematical model via five significant steps, as displayed in Table 2, which can be illustrated as follows:
Step 1: The random walking patterns of ants in search of space: it is necessary for the antlion and the ant to engage with one another, as part of the antlion’s hunting behavior. The ants must travel to an area in which they look for food and cover, and this is where the antlions set their traps for the ants. The stochastic ant’s behavior—while they look for food simulated by a random walking pattern—is selected as Equation (6), where t indicates the random walk’s steps; max_Iter addresses the highest numbers of iteration, cusum evaluates the cumulative sum, and r ( t ) refers to the stochastics expression specified by Equation (7). R is a random, uniformly distributed number inside the range [0, 1]. Therefore, the normalized random walking pattern is represented in Equation (8) where ai and bi show, respectively, related to every variable (i), the lower and higher values of the random walking pattern, and C i t and d i t symbolize the lowest and the highest values in every iteration (t) related to every variable (i).
Step 2: Traps, established where the fittest antlions with the best chance of capturing prey, are selected through this step. The ALO program uses the roulette wheel to find the fittest antlion during optimization.
Step 3: Ants become caught in nets where the antlion’s traps affect the random walking pattern. Equations (9) and (10) thus represent the formal relationship for this premise. In both equations, the hypersphere of randomly moving ants, around the chosen antlion, is indicated by C and D.
Step 4: Ants move toward the antlion, forcing the ant inside the trap. The sliding motion of an ant into the pit is represented by the two algebraic Equations (11) and (12), where I denotes the ratio specified in Equation (13), t indicates the current iteration, and w represents the constant that depends on the present iteration.
Step 5: This step mimics catching the targeted prey and reconstructing the trap where the antlion captures the ant. Consequently, the hunting process is complete, as described by Equation (14), based on the comparative assessment of their fitness value. In Equation (14), A n t l i o n j t and A n t i t stand for, respectively, the positions of the chosen antlion (j) and ant (i). The functions f ( A n t i t ) and f ( A n t l i o n j t ) demonstrate, respectively, the values of the goal of the ant and the antlion. The antlion then upgrades its location, or builds a new trap to capture a new target.
Step 6: Elitism enables the program to find the best solution at each stage of the refining process. In every iteration of the ALO, the best antlion is located and stored as an elite, since the top selection influences all ant movement throughout rounds. As a result, it is assumed that every ant travels through the selected antlion at random and concurrently, as in Equation (15), where A n t i t indicates the position of an ant ( i ) in iteration (t); R A t refers to the random walking pattern around the chosen antlion in this instance; and R E t shows the random walking pattern close to the elite solution. The termination condition of a flowchart, of ALO in Figure 5, is achieved when the number of iterations reaches the maximum number.

3.3. Optimization Process Using the Proposed ALO in Handling the HIL-Based Experimental Model

The two optimization methods, ALO and MPSO, were exploited to optimize the parameters of the PID-PSS for the generator named G1 on the SMIB power system. Equations (16)–(19) demonstrate J ’s objective function to optimize the PID-PSS gains. All the viewed objective functions have been applied within the two AI optimization tools, while objective J 2 provides the best performance with MPSO and J 4 gives the best performance with ALO.
J 1 = max R e a l λ i
J 2 = min D a m p i n g r a t i o
J 3 = I T S E = 0 t e 2 t d t
J 4 = I T A E = 0 t e t d t
MPSO and ALO optimizes the PID parameters using the four objective functions J’s according to Equations (20)–(22).
K P m i n K P K P m a x
K I m i n K I K I m a x
K d m i n K d K d m a x
The MPSO differs from the original PSO by modifying the search space boundary, which is used as a reflecting wake to reflect the particles into the search and prevent it from going out of the search space. These modifications of the original PSO algorithm improve the algorithm’s overall performance. The author used this method with a PD-based controller to enhance the performance of the PSS [11].

4. Results and Analysis

In this part, the performance of the developed PSS is illustrated. First, the proposed ALO is applied, and its performance is compared to one of the highly effective techniques, modified particle swarm optimization (MPSO). The MPSO technique is an effective modified version of the original PSO algorithm to improve the algorithm’s overall performance, which was first proposed in [49]. In this MPSO, the search space boundary is used as a reflecting wake to exhibit the particles in the search and prevent it from going out of the search space. This effective version has been employed with a PD-based controller to enhance the performance of the PSS [50]. In this study, the upper and lower boundaries of the PID gains are set from −50 to 50 for the two optimization processes. The proposed ALO is applied to tune the PID controller, and Table 3 shows the ALO parameters. Table 4 exhibits the proposed PID-PSS parameters, compared to MPSO.
The ALO is applied by the mentioned parameters in Table 3, on the m-file containing linearized SMIB with PID-PSS, according to the suggested objective function to compute the PSS gains shown in Table 4. Next, the performance is tested experimentally, as mentioned previously. Various case studies adopted optimization algorithms, as summarized in Table 5. Different loading conditions are considered. These conditions are named according to the percentage of loading from light, nominal, and heavy loading conditions. Additionally, twelve distinct scenarios are addressed and evaluated, considering different step changes, single, double, and three-phase short circuits (S. C.) at various transmission line (TL) locations. All the graphs in the following sections were plotted from real-time test results.

4.1. Light Loading

In this case, the light loading condition is considered, and both controllers, previously optimized in the preceding section, are applied. Table 5 shows their impacts on the system, compared to the initial case. The eigenvalues, damping ratio, damped and undamped frequencies, and the synchronizing and damping torque coefficients are tabulated in Table 6. The second and third eigenvalues are positive (0.0142 ± 5.3557i), indicating that the system is initially unstable. Using the MPSO and proposed ALO-based PID-PSS controllers, the system achieves acceptable performance where the second and third eigenvalues are transformed to the negative side of (0.7802 ± 4.1830i) and (−1.1504 ± 4.5691i), respectively.
Moreover, the proposed ALO-based PID-PSS boosts the damping ratio and the synchronizing and damping torque more than the modified PSO-based PID-PSS. The natural frequency of the ALO-based PSS was found to be higher than the MPSO–PSS. On the other hand, MPSO-based PSS have a higher damped frequency, compared with ALO-based PSS. Consequently, under light loading conditions, the developed ALO-based PID-PSS is superior and more robust than the modified PSO-based PID-PSS.
The experimental validations are described for each scenario at this loading condition as follows:

4.1.1. Simulation of Case Study No. (1) during Light Loading

In this case, a 5% load step change is considered, where Figure 6 shows its impacts on the controlled machine speed deviation, rotor angle deviation, active power, and voltage at bus 1 (B1). As shown, the proposed ALO-based PID-PSS derives superior stability performance than the modified PSO-based PID-PSS for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1. The developed PSS improves the system stability with the lowest settling time and area under the curve. Furthermore, the adopted optimization algorithm supports the damping of the LFOs faster than the MPSO.
In accompanying those figures, the numerical analysis of the considered objectives is summarized in Table 7. From this table, regarding the ISE performance measures, the proposed ALO-based PID-PSS has a much lower value than the modified PSO-based PID-PSS, with the improvement of 23.64%, 8.55%, 7.87%, and 7.86% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.
Similar findings are obtained, considering the IAE performance measures, the proposed ALO-based PID-PSS has a much lower value than the modified PSO-based PID-PSS with the improvement of 20.28%, 8.21%, 7.87%, and 7.86% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively. Additionally, considering the ITSE performance measures, the proposed ALO-based PID-PSS has a much lower value than the modified PSO-based PID-PSS, with the improvement of 31.79%, 15.56%, 15.11%, and 15.11% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively. Considering the ITAE performance measures, the proposed ALO-based PID-PSS has a much lower value than the modified PSO-based PID-PSS with the improvement of 25.87%, 15.34%, 15.11%, and 15.11% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.
The developed PSS has the smallest ISE, IAE, ISAE, and ITAE, compared with the MPSO. On average, the proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an improvement of 25.4%, 11.91%, 11.49%, and 11.48% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.1.2. Case Study No. (2) during Light Loading

Similar to case No. 1, a 10% load step change has been considered. A significant load change directly affects the speed deviation, rotor angle deviation, active power, and voltage on bus B1 of the controlled machine. Table 8 summarizes the effects of step load change, using step information performance indices on the PSS to show the robustness of the adopted optimization tools.
As shown, the proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 27.04%, 12.08%, 9.12%, and 11.48% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.1.3. Case Study No. (3) during Light Loading

In this case, a 3-phase S.C. is considered at the sending end of the TL at a time of 5 s. The fault lasts for five cycles. Figure 7 reveals the performance of the PSS to this stern test. The proposed ALO-based PID-PSS reduces the oscillation and makes the system regain the same equilibrium point faster than the MPSO-based PSS.
Table 9 summarizes the performance indices related to the three-phase S.C. at the beginning of the TL. As shown, the proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 15.79%, 4.49%, 11.43%, and 11.49% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.1.4. Case Study No. (4) during Light Loading

In this case, a single-phase S.C. to the ground occurred in the middle of the TL. Table 10 summarizes performance indices of the developed PSS based on adopted optimization tools to show the robustness of the developed method. As shown, the proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 18.81%, 6.53%, 11.48%, and 11.49% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.1.5. Case Study No. (5) during Light Loading

In this case, a three-phase S.C. to ground fault occurred at the end of the TL. Table 11 summarizes the system response to this fault using performance measures. As illustrated, the performance indices emphasize the robustness of the proposed PSS, compared with the MPSO-based PSS since it has a smaller ISE, IAE, ITSE, and ITAE than the other assessed PSS. The proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 15.12%, 7.22%, 11.46%, and 11.49% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.2. Nominal Loading

The nominal loading situation is considered in this instance, and both previously optimized controllers are used. Table 12 compares their effects on the system to the original model and displays the results. The fact that the second and third eigenvalues are positive (0.1646 ± 5.5081i), as shown, implies that the system is inherently unstable. The system obtains adequate performance when the second and third eigenvalues are transformed to the negative side of (−25.0930 ± 38.1149i) and (−24.4835 ± 30.2602i), respectively, using the MPSO and proposed ALO-based PID-PSS controllers.

4.2.1. Case Studies No. (6) and (7) during Nominal Loading

The first 2 cases in this loading condition consider 5% and 10% load changes, respectively. The impacts of the proposed ALO-based PID-PSS and the modified PSO-based PID-PSS are assessed on the controlled machine using speed deviation, rotor angle deviation, active power, and voltage at bus 1 (B1). Figure 8 displays the valuable improvement percentages of the proposed ALO controller versus MPSO. As shown, the proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS with an average improvement of 33.57%, 11.69%, 11.49%, and 11.48% at 5% step load change, and 17.76%, 6.51%, 11.48%, and 11.49% at 10% step load change for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.2.2. Case Study No. (8) during Nominal Loading

In this case, a double-line-to-ground fault at the sending end of the TL is considered. The performance indices are calculated and summarized in Table 13. The performance indices clearly show the superiority of the developed PSS, compared with the conventional PSS. This conclusion is based on the performance indices that are lower in the case of the proposed PSS. The proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 17.79%, 1.74%, 11.48%, and 11.33% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.2.3. Case Study No. (9) during Nominal Loading

Likewise, the double-line-to-ground fault is addressed in this case study. The fault is initiated at 5 s in the middle of the TL. Figure 9 shows the speed deviation, rotor angle deviation, active power, and B1 positive sequence voltage. As shown, the four signals emphasize the significant impact of the developed PSS to dampen the LFO better than the modified PSO-based PSS. In accompanying those figures, the numerical analysis of the considered objectives is summarized in Table 14. From this table, the proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 19.72%, 3.74%, 11.48%, and 11.49% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.3. Heavy Loading

The third and final step toward validating the robustness of the developed PSS covers the heavy loading condition. Both previously adjusted controllers are used. Table 15 compares their effects on the system to the original instance and displays the results. The fact that the second and third eigenvalues are positive (0.1843 ± 5.5145i), as shown, implies that the system is inherently unstable. The system obtains adequate performance when the second and third eigenvalues are transformed to the negative side of (−24.9295 ± 38.8812i) and (−24.3121 ± 30.9152i), respectively, using the MPSO and proposed ALO-based PID-PSS controllers.

4.3.1. Case Study No. (10) during Heavy Loading

The load step change disturbance is applied in this case, where 5% of the load is suddenly added after 4.1 s. Figure 10 shows the controlled machine speed deviation, rotor angle deviation, active power, and the bus B1 system response voltage in this case study. This figure reveals that the system’s response to the proposed PSS has a shorter settling time, oscillation range, area under the curve, and higher rise time than the other PSS. In this regard, Table 16 summarizes the system response to this fault using the performance measures.
As illustrated, the performance indices emphasize the robustness of the proposed PSS, compared with the MPSO-based PSS, since it has a smaller ISE, IAE, ITSE, and ITAE than the other assessed PSS. The proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 15.12%, 7.22%, 11.46%, and 11.49% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.3.2. Case Study No. (11) during Heavy Loading

In this case, a three-line-to-ground fault has been applied to evaluate the robustness of the proposed PSS. The fault has been involved at the sending end of the TL. The performance indices of this case study are reported in Table 17. As can be seen, the proposed PSS enhances the overall system stability more than the modified PSO-based PID-PSS. The proposed ALO-based PID-PSS has much lower objective metrics than the modified PSO-based PID-PSS, with an average improvement of 12.28%, 2.81%, 11.47%, and 11.49% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively.

4.3.3. Case Study No. (12) during Heavy Loading

To fully assess the resilience of the optimized PSS, a three-line-to-ground fault is planned to occur in the center of the TL. Five cycles of the test are conducted. The deviations of the governed machine speed, the rotary angle, the active power, and the voltage are shown in Figure 11 for bus B1. This finding confirms the results of the other case studies in this work. In contrast to the modified PSO-based PID PSS, Figure 11 demonstrates that the system responses, using the suggested PSS, have superior performance.
To emphasize the amount of participation of the optimized controller, Figure 12 displays the performance metrics of both MPSO-based PID-PSS and the proposed ALO-based PID-PSS controllers. As shown in Figure 12, with an average increase of 10.73%, 8.86%, 11.48%, and 11.49% for stabilizing the machine speed deviation, rotor angle deviation, active power, and voltage at B1, respectively, the suggested ALO-based PID-PSS has significantly lower objective metrics than the modified PSO-based PID-PSS.

5. Conclusions

This paper proposed an advanced, ALO-based PID-PSS to improve the power system stability. In this study, the adoption of powerful optimization tools, in combination with hardware-in-the-loop (HIL), provides real-time experimental validation. The developed PSS, via the ALO algorithm, has complied with a reconfigurable input/output single-board (SbRIO-9636), made by the National Instrument considering software package, MATLAB/SIMULINK. A total of 12 case studies were simulated under different loading conditions and disturbances to ensure the robustness and effectiveness of the developed ALO-based PID-PSS over the MPSO-based PID-PSS. Under the various loading conditions, the proposed ALO-based PID-PSS successfully demonstrated a high ability to regain power system stability.
Several disturbances were investigated, which were symmetrical and unsymmetrical faults, different fault locations, and step load changes. The performance indices of ALO-based PID-PSS and MPSO-based PID-PSS were compared to assess the developed PSS’s superiority under all test cases. The comparison results showed that the developed PSS improved the rise time, settling time, oscillation range, and the area under the curve for all cases by 22.031%, 12.894%, 19.859%, and 267.849%, respectively. The average scores of ISE, IAE, ITSE, and ITAE were 30.919%, 23.295%, 51.073%, and 53.624%, respectively.
The current study tested ALO’s ability to produce PSS which is able improve the power system stability through real-time constraints. The valuable part of this paper is that the system is configured with low-cost electronics boards, which is a new development in HIL tests, since existing HIL tests depend on costly real-time digital simulator systems or costly boards and high-speed, high-cost computers. However, one of the disadvantages of ALO is that it may become caught in premature convergence. The ALO technique in this paper was developed to handle the problem of PSS design, via a PID-based controller, to improve the power system stability of a single machine infinite bus model. The assessment of the ALO algorithm is compared to one of the improved and efficient versions of PSO technique. The applications show the significant performance of the ALO in solving the considered problems. Additionally, experimental validations, through the DSPACE electronics environment, using several scenarios of fault types and locations, provides great validation of the ALO in solving the considered problems.
The authors currently study how the current system can test the system with a higher number of generators, which may be a significant challenge with the same boards—especially the DSPACE board—since it runs the system in real time. This issue needs more investigation and work, which may depend on the cost of the system components.

Author Contributions

Conceptualization, N.M.A.I.; Methodology, N.M.A.I.; Software, A.M.S.; Validation, N.M.A.I. and A.M.S.; Formal analysis, B.A.H.; Data curation, B.A.H.; Writing—original draft, N.M.A.I.; Writing—review & editing, A.M.S. and B.A.H.; Supervision, H.E.A.T.; Project administration, H.E.A.T. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by [Future University in Egypt].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare that there is no conflict of interest.

Glossary

List of symbols:
x i :The state variables.
i :The number of states.
X i :The derivative of the state variables.
D :Damping coefficient.
M :Inertia coefficient.
K 1 : K 6 : Constants of the linearized model.
ω b :The rated speed in elec. rad/s = 2πf0 = 377 for 60 Hz system
T d o T q o :(d and q)-axis open circuit transient time constant, respectively.
U :The vector of inputs to the system.
T m :The mechanical torque, p.u.
V r e f :The reference voltage, p.u.
λ i :The electromechanical mode eigenvalue.
ξ :The damping ratio, per unit speed change MVA/MW.
K A :Voltage regulator gain.
T A :Voltage regulator time constants.
X :The state vector
Δ :Linearized incremental quantity.
Δ ω :Machine speed deviation, p.u.
Δ δ :Rotor angle or power angle deviation, rad.
Δ E q :The q-axis component of the transient voltage behind transient reactance.
Δ E f d :Generator field or exciter output voltage deviation, p.u.
X ( t ) :The random walks of ants in search of space.
c u s u m : The cumulative sum.
t : The step of random walk.
r t : A stochastic function.
r a n d : a random number generated with uniform distribution in the interval of [0, 1].
X i ( t ) : The normalized random walk of prey in the search space.
a i : The minimum of random walk of i-th variable.
b i : The maximum of random walk in i-th variable.
c i t : The minimum of i-th variable at t-th iteration.
d i t : The maximum of i-th variable at t-th iteration.
A n t l i o n j t : The position of the selected j-th antlion at t-th iteration.
c t : The minimum of all variables at t-th iteration.
d t : The vector including the maximum of all variables at t-th teration.
c j t : The minimum of all variables for i-th ant.
d j t : The maximum of all variables for i-th ant.
I : A ratio.
t : The current iteration.
T : A constant defined based on the current iteration (w = 2 when t > 0.1 T, w = 3 when t > 0.5 T, w = 4 when t > 0.75 T, w = 5 when t > 0.9 T, and w = 6 when t > 0.95 T).
A n t i t : The position of i-th ant at t-th iteration.
J i : The J’s objective function.
K S : The steady-state synchronizing torque coefficient, per unit torque/rad.
K D : The damping torque coefficient, per unit torque/per unit speed change.
ω n : The undamped natural frequency, rad/s.
ω d : The damped frequency, rad/s.
List of abbreviations:
PSS:Power system stabilizer.
PID:Proportional integral derivative.
ALO:The Antlion optimization.
HIL:Hardware-in-the-loop.
FPGA:Field programmable gate arrays.
NI:National Instruments.
MPSO:Modified particle swarm optimization.
ISE:The integral square of the error.
IAE:The integral absolute of the error.
ITSE:The integral time square of the error.
ITAE:The integral time absolute of the error.
GA:Genetic algorithm.
NSGAII:Non-dominated sorting genetic algorithm II.
PBIL:Population-based incremental learning.
SMIB:Single-machine infinite bus.
MMPS:Multi-machine power system.
RCP:Rapid control prototyping.
SIL:Software in the loop.
A/D:Analog/digital.

Appendix A

Table A1 describes the dynamic test system parameters.
Table A1. The power system model parameters [39].
Table A1. The power system model parameters [39].
DeviceParameterValue
Hydro Generator
(Salient pole)
F50 Hz
D0
M7.4000 S
T d o 4.42586 S
X d 1.305 P.U.
X d 0.296 P.U.
X q 0.474 P.U.
K A 200
T A 0.001 S
T-LL0.8737 × 10−3 H/Km
R0.01755 Ω /Km
Table A2 reports the specs and system parameter modification on a Simulink model to generate the necessary system code to be implemented in real time.
Table A2. dSPACE RTI1104 configuration parameter requirements [40].
Table A2. dSPACE RTI1104 configuration parameter requirements [40].
GainsPropertyValue
SolverStart Time0.0
Stop TimeInf.
Fixed Step-SizeFixed-Step
Solver TypeUnconstrained
Periodic Sample Time Constraint
Higher Priority Value Indicates Higher Task PriorityOff
OptimizationBlock ReductionOff
Signal Storage ReuseOff
Code GenerationSystem Target FileRTI 1104
Table A3 reports the technical specifications of the NI SbRIO-9636 device I/O, as in the product manual.
Table A3. NI SbRIO-9636 device I/O technical specification [51].
Table A3. NI SbRIO-9636 device I/O technical specification [51].
ComponentsParameter
Storage 512 MB
DRAM256 MB
Processing Speed400 MHz
AI16ch-16bit
AO4ch-16bit
RS-2322
RS-4851
CAN1
USB1
FPGA typeXilinx Spartan-6 LX45
Number of flip-flops54,576
Number of 6-input LUTs27,288
Input Range−10 V to 10 V

References

  1. Bhukya, J.; Mahajan, V. Mathematical modelling and stability analysis of PSS for damping LFOs of wind power system. IET Renew. Power Gener. 2019, 13, 103–115. [Google Scholar] [CrossRef]
  2. El Ela, A.A.A.; El-Sehiemy, R.A.; Shaheen, A.M.; Diab, A.E.G. Optimal design of PID controller based sampe-jaya algorithm for load frequency control of linear and nonlinear multi-area thermal power systems. Int. J. Eng. Res. Afr. 2020, 50, 79–93. [Google Scholar] [CrossRef]
  3. Prasertwong, K.; Mithulananthan, N.; Thakur, D. Understanding low-frequency oscillation in power systems. Int. J. Electr. Eng. Educ. 2010, 47, 248–262. [Google Scholar] [CrossRef]
  4. Reddy, J.; Kishore, M.J. Real time implementation of H∞ loop shaping robust PSS for multimachine power system using dSPACE. Int. J. Electr. Power Energy Syst. 2011, 33, 10. [Google Scholar] [CrossRef]
  5. Aghazade, A.; Kazemi, A. Simultaneous coordination of power system stabilizers and STATCOM in a multi-machine power system for enhancing dynamic performance. In Proceedings of the 4th International Power Engineering and Optimization Conference (PEOCO 2010), Program and Abstracts, Shah Alam, Malaysia, 23–24 June 2010. [Google Scholar] [CrossRef]
  6. Guesmi, T.; Farah, A.; Abdallah, H.H.; Ouali, A. Robust design of multimachine power system stabilizers based on improved non-dominated sorting genetic algorithms. Electr. Eng. 2018, 100, 1351–1363. [Google Scholar] [CrossRef]
  7. Ameli, A.; Farrokhifard, M.; Ahmadifar, A.; Safari, A.; Shayanfar, H.A. Optimal tuning of Power System Stabilizers in a multi-machine system using firefly algorithm. In Proceedings of the 12th International Conference on Environment and Electrical Engineering (EEEIC 2013), Wroclaw, Poland, 5–8 May 2013. [Google Scholar] [CrossRef]
  8. Dombo, D.A.; Folly, K. Multi-machine power system stabilizer design based on population based incremental learning. In Proceedings of the 2015 IEEE Symposium Series on Computational Intelligence (SSCI 2015), Cape Town, South Africa, 7–10 December 2015. [Google Scholar] [CrossRef]
  9. Peres, W.; Júnior, I.C.S.; Filho, J.A.P. Gradient based hybrid metaheuristics for robust tuning of power system stabilizers. Int. J. Electr. Power Energy Syst. 2018, 95, 47–72. [Google Scholar] [CrossRef]
  10. Sabo, A.; Wahab, N.I.A.; Othman, M.L.; Jaffar, M.Z.A.M. Mitigation of oscillations in smib using a novel farmland fertility optimization based PIDPSS. In Proceedings of the 2020 2nd International Conference on Smart Power and Internet Energy Systems (SPIES 2020), Bangkok, Thailand, 15–18 September 2020. [Google Scholar] [CrossRef]
  11. Ibrahim, N.M.A.; Elnaghi, B.E.; Ibrahem, H.A.; Talaat, H.E.A. Modified particle swarm optimization based on lead-lag power system stabilizer for improve stability in multi-machine power system. Int. J. Electr. Eng. Inform. 2019, 11, 161–182. [Google Scholar] [CrossRef]
  12. Izci, D. A novel improved atom search optimization algorithm for designing power system stabilizer. Evol. Intell. 2022, 15, 2089–2103. [Google Scholar] [CrossRef]
  13. Alotaibi, I.M.; Ibrir, S.; Abido, M.A.; Khalid, M. Nonlinear Power System Stabilizer Design for Small Signal Stability Enhancement. Arab. J. Sci. Eng. 2022, 47, 13893–13905. [Google Scholar] [CrossRef]
  14. Bayu, E.S.; Khan, B.; Ali, Z.M.; Alaas, Z.M.; Mahela, O.P. Mitigation of Low-Frequency Oscillation in Power Systems through Optimal Design of Power System Stabilizer Employing ALO. Energies 2022, 15, 3809. [Google Scholar] [CrossRef]
  15. Yokus, H.; Ozturk, A. A robust crow search algorithm-based power system stabilizer for the SMIB system. Neural Comput. Appl. 2022, 34, 9161–9173. [Google Scholar] [CrossRef]
  16. Ansari, J.; Abbasi, A.R.; Heydari, M.H.; Avazzadeh, Z. Simultaneous design of fuzzy PSS and fuzzy STATCOM controllers for power system stability enhancement. Alex. Eng. J. 2022, 61, 2841–2850. [Google Scholar] [CrossRef]
  17. Rodrigues, F.; Molina, Y.; Silva, C.; Ñaupari, Z. Simultaneous tuning of the AVR and PSS parameters using particle swarm optimization with oscillating exponential decay. Int. J. Electr. Power Energy Syst. 2021, 133, 107215. [Google Scholar] [CrossRef]
  18. Chaib, L.; Choucha, A.; Arif, S. Optimal design and tuning of novel fractional order PID power system stabilizer using a new metaheuristic Bat algorithm. Ain Shams Eng. J. 2017, 8, 113–125. [Google Scholar] [CrossRef]
  19. Hemmati, R. Power system stabilizer design based on optimal model reference adaptive system. Ain Shams Eng. J. 2016, 9, 311–318. [Google Scholar] [CrossRef]
  20. Abid, S.; El-Rifaie, A.M.; Elshahed, M.; Ginidi, A.R.; Shaheen, A.M.; Moustafa, G.; Tolba, M.A. Development of Slime Mold Optimizer with Application for Tuning Cascaded PD-PI Controller to Enhance Frequency Stability in Power Systems. Mathematics 2023, 11, 1796. [Google Scholar] [CrossRef]
  21. El-Sehiemy, R.; Shaheen, A.; Ginidi, A.; Al-Gahtani, S.F. Proportional-Integral-Derivative Controller Based-Artificial Rabbits Algorithm for Load Frequency Control in Multi-Area Power Systems. Fractal Fract. 2023, 7, 97. [Google Scholar] [CrossRef]
  22. Huang, J.; Liu, J.; Zhang, C.; Kuang, Y.; Weng, S. Application of Improved Quasi-Affine Transformation Evolutionary Algorithm in Power System Stabilizer Optimization. Electronics 2022, 11, 2785. [Google Scholar] [CrossRef]
  23. Radovici, A.; Culic, I. Embedded Systems Software Development. In Getting Started with Secure Embedded Systems; Springer: Berlin/Heidelberg, Germany, 2022. [Google Scholar]
  24. Irfan, A.; Khan, M.G.; Amin, A.A.; Mohsin, S.A.; Adnan, M.; Zulfiqar, A. Model-Based Design, HIL Testing, and Rapid Control Prototyping of a Low-Cost POC Quadcopter with Stability Analysis and Control. Complexity 2022, 2022, 1492170. [Google Scholar] [CrossRef]
  25. Menghal, P.M.; Laxmi, A.J. Real time simulation: A novel approach in engineering education. In Proceedings of the 2011 3rd International Conference on Electronics Computer Technology (ICECT 2011), Kanyakumari, India, 8–10 April 2011. [Google Scholar] [CrossRef]
  26. Estrada, L.; Vázquez, N.; Vaquero, J.; de Castro, Á.; Arau, J. Real-time hardware in the loop simulation methodology for power converters using labview FPGA. Energies 2020, 13, 373. [Google Scholar] [CrossRef]
  27. Picerno, M.; Lee, S.-Y.; Pasternak, M.; Siddareddy, R.; Franken, T.; Mauss, F.; Andert, J. Real-Time Emission Prediction with Detailed Chemistry under Transient Conditions for Hardware-in-the-Loop Simulations. Energies 2022, 15, 261. [Google Scholar] [CrossRef]
  28. Almas, M.S.; Vanfretti, L.; Vanfretti, L. Experimental performance assessment of a generator’s excitation control system using real-time hardware-in-the-loop simulation. In Proceedings of the 40th Annual Conference of the IEEE Industrial Electronics Society (IECON 2014), Dallas, TX, USA, 29 October–1 November 2014. [Google Scholar] [CrossRef]
  29. Tian, H.; Guo, S.; Zhao, P.; Gong, M.; Shen, C. Design and implementation of a real-time multi-beam sonar system based on fpga and dsp. Sensors 2021, 21, 1425. [Google Scholar] [CrossRef] [PubMed]
  30. Saralegui, R.; Sanchez, A.; de Castro, A. Modeling of deadtime events in power converters with half-bridge modules for a highly accurate hardware-in-the-loop fixed point implementation in fpga. Appl. Sci. 2021, 11, 6490. [Google Scholar] [CrossRef]
  31. Hermassi, M.; Krim, S.; Kraiem, Y.; Hajjaji, M.A.; Alshammari, B.M.; Alsaif, H.; Alshammari, A.S.; Guesmi, T. Design of Vector Control Strategies Based on Fuzzy Gain Scheduling PID Controllers for a Grid-Connected Wind Energy Conversion System: Hardware FPGA-in-the-Loop Verification. Electronics 2023, 12, 1419. [Google Scholar] [CrossRef]
  32. El Zerk, A.; Ouassaid, M. Real-Time Fuzzy Logic Based Energy Management System for Microgrid Using Hardware in the Loop. Energies 2023, 16, 2244. [Google Scholar] [CrossRef]
  33. Essa, M.E.S.M.; Lotfy, J.V.W.; Abd-Elwahed, M.E.K.; Rabie, K.; ElHalawany, B.M.; Elsisi, M. Low-Cost Hardware in the Loop for Intelligent Neural Predictive Control of Hybrid Electric Vehicle. Electronics 2023, 12, 971. [Google Scholar] [CrossRef]
  34. Samano-Ortega, V.; Padilla-Medina, A.; Bravo-Sanchez, M.; Rodriguez-Segura, E.; Jimenez-Garibay, A.; Martinez-Nolasco, J. Hardware in the loop platform for testing photovoltaic system control. Appl. Sci. 2020, 10, 8690. [Google Scholar] [CrossRef]
  35. Hasan, N.; Ibraheem, I.; Farooq, S. Real time simulation of automatic generation control for interconnected power system. Int. J. Electr. Eng. Inform. 2012, 4, 40–51. [Google Scholar] [CrossRef]
  36. Almas, M.S.; Vanfretti, L. Implementation of conventional and phasor based power system stabilizing controls for real-time simulation. In Proceedings of the 40th Annual Conference of the IEEE Industrial Electronics Society (IECON 2014), Dallas, TX, USA, 29 October–1 November 2014. [Google Scholar] [CrossRef]
  37. Scheibe, C.; Kuri, A.; Graf, L.; Venugopal, R.; Mehlmann, G. Real Time Co-Simulation of Electromechanical and Electromagnetic Power System Models. In Proceedings of the 5th International Conference on Smart Energy Systems and Technologies (SEST 2022), Porto, Portugal, 9–11 September 2022. [Google Scholar] [CrossRef]
  38. Prabha, O.P.M.; Kundur, S. Power System Stability And Control by Prabha Kundur; McGraw-Hill, Inc.: New York, NY, USA, 2022. [Google Scholar]
  39. Jovcic, D.; Pillai, G.N. Analytical modeling of TCSC dynamics. IEEE Trans. Power Deliv. 2005, 20, 1097–1104. [Google Scholar] [CrossRef]
  40. DS1104 R&D Controller Board. Available online: www.dspace.com/en/pub/home/products/hw/singbord/ds1104.cfm (accessed on 5 February 2022).
  41. Piyaratna, S.; Duong, N.; Carr, J.; Bird, D.; Kennedy, S.; Udina, A.; Jenkinson, P. Digital RF processing system for Hardware-in-the-Loop simulation. In Proceedings of the 2013 International Conference on Radar—Beyond Orthodoxy: New Paradigms in Radar (RADAR 2013), Adelaide, SA, Australia, 9–12 September 2013. [Google Scholar] [CrossRef]
  42. Bacic, M. On hardware-in-the-loop simulation. In Proceedings of the 44th IEEE Conference on Decision and Control, and the European Control Conference (CDC-ECC’05), Seville, Spain, 15 December 2005. [Google Scholar] [CrossRef]
  43. Latha, A.H.; Shubhanga, K.N. Hardware-in-the-Loop testing of PSS for synchronous generator. In Proceedings of the IEEE International Conference on Green Computing, Communication and Electrical Engineering (ICGCCEE 2014), Coimbatore, India, 6–8 March 2014. [Google Scholar] [CrossRef]
  44. Lee, E.A.; Seshia, S.A.; Jensen, J.C. Teaching embedded systems the Berkeley way. In Proceedings of the 2012 Workshop on Embedded and Cyber-Physical Systems Education (WESE 2012), Tampere, Finland, 12 October 2013. [Google Scholar] [CrossRef]
  45. Pathak, V.K.; Gangwar, S.; Singh, R.; Srivastava, A.K.; Dikshit, M. A comprehensive survey on the ant lion optimiser, variants and applications. J. Exp. Theor. Artif. Intell. 2022, 28, 1397–1416. [Google Scholar] [CrossRef]
  46. Assiri, A.S.; Hussien, A.G.; Amin, M. Ant lion optimization: Variants, hybrids, and applications. IEEE Access 2020, 8, 77746–77764. [Google Scholar] [CrossRef]
  47. Mirjalili, S. The Ant Lion Optimizer. Adv. Eng. Softw. 2015, 83, 80–98. [Google Scholar] [CrossRef]
  48. Raju, M.; Saikia, L.C.; Sinha, N. Automatic generation control of a multi-area system using ant lion optimizer algorithm based PID plus second order derivative controller. Int. J. Electr. Power Energy Syst. 2016, 80, 52–63. [Google Scholar] [CrossRef]
  49. Robinson, J.; Rahmat-Samii, Y. Particle swarm optimization in electromagnetics. IEEE Trans. Antennas Propag. 2004, 52, 397–407. [Google Scholar] [CrossRef]
  50. Ibrahim, N.M.; Attia, H.E.; Talaat, H.E.; Alaboudy, A.H.K. Modified Particle Swarm Optimization Based Proportional-Derivative Power System Stabilizer. Int. J. Intell. Syst. Appl. 2015, 7, 62–76. [Google Scholar] [CrossRef]
  51. NI sbRIO-960x/962x/963x OEM Instructions and Specifications. Available online: www.ni.com/pdf/manuals/realtest.pdf (accessed on 7 February 2022).
Figure 1. Heffron–Philips SMIB System.
Figure 1. Heffron–Philips SMIB System.
Sustainability 15 08966 g001
Figure 2. Dynamic workbench test Simulink model.
Figure 2. Dynamic workbench test Simulink model.
Sustainability 15 08966 g002
Figure 3. Generator excitation system preparation to connect with real-time.
Figure 3. Generator excitation system preparation to connect with real-time.
Sustainability 15 08966 g003
Figure 4. Experimental setup using dSPACE and NI SbRIO-9636 board; (a) FPGA platforms; (b) NI SbRIO-9636 board [41].
Figure 4. Experimental setup using dSPACE and NI SbRIO-9636 board; (a) FPGA platforms; (b) NI SbRIO-9636 board [41].
Sustainability 15 08966 g004
Figure 5. ALO flow chart.
Figure 5. ALO flow chart.
Sustainability 15 08966 g005
Figure 6. The system reaction to a 5% load step change during light loading.
Figure 6. The system reaction to a 5% load step change during light loading.
Sustainability 15 08966 g006aSustainability 15 08966 g006b
Figure 7. The system reaction to a three-phase S.C. at the beginning of the T-L during light loading.
Figure 7. The system reaction to a three-phase S.C. at the beginning of the T-L during light loading.
Sustainability 15 08966 g007aSustainability 15 08966 g007b
Figure 8. Improvement percentages of the proposed ALO controller versus MPSO with load changes during nominal loading. (a) 5% Step Change (Case Study No (6)); (b) 10% Step Change (Case Study No (7)).
Figure 8. Improvement percentages of the proposed ALO controller versus MPSO with load changes during nominal loading. (a) 5% Step Change (Case Study No (6)); (b) 10% Step Change (Case Study No (7)).
Sustainability 15 08966 g008
Figure 9. Stability impacts of a double-phase S.C. on the middle of the T-L during nominal loading.
Figure 9. Stability impacts of a double-phase S.C. on the middle of the T-L during nominal loading.
Sustainability 15 08966 g009
Figure 10. The effect of a 5% load step change during heavy loading.
Figure 10. The effect of a 5% load step change during heavy loading.
Sustainability 15 08966 g010
Figure 11. The effect of three-phase S.C. at the middle of the TL during heavy loading, they are listed as: (a) speed deviation; (b) Rotor angle deviation; (c) active power; (d) enlarged area (A) from subfigure (c); (e) Deviations in positive sequence line voltage at B1.
Figure 11. The effect of three-phase S.C. at the middle of the TL during heavy loading, they are listed as: (a) speed deviation; (b) Rotor angle deviation; (c) active power; (d) enlarged area (A) from subfigure (c); (e) Deviations in positive sequence line voltage at B1.
Sustainability 15 08966 g011
Figure 12. The performance metrics of both MPSO-based PID-PSS and the proposed ALO-based PID-PSS controllers.
Figure 12. The performance metrics of both MPSO-based PID-PSS and the proposed ALO-based PID-PSS controllers.
Sustainability 15 08966 g012
Table 1. The parameter of dSPACE DS1104 controls ports [40].
Table 1. The parameter of dSPACE DS1104 controls ports [40].
ParametersPorts
8 Parallel DAC Channels
One A/D Converter
(ADC1) Multiplexed to Four Channels.
Four Parallel A/D
Converters with One
Channel Each
ChannelsDACH1:DACH8ADCH1:ADCH4ADC2:ADC5
Resolution16-bit16-bit12-bit
Conversion time-2 µs800 ns
Settling timeMax. 10 µs--
I/O voltage ranges±10 V±10 V±10 V
Offset error±1 mV±5 mV±5 mV
Gain error±0.1%±0.25%±0.5%
Signal-to-noise ratio (SNR)>80 dB
(At 10 kHz)
>80 dB
(At 10 kHz)
>65 dB
Simulink I/O−1: +1 (Double)−1: +1 (Double)−1: +1 (Double)
Offset drift130 µV/K40 µV/K40 µV/K
Gain drift25 ppm/K25 ppm/K25 ppm/K
Table 2. The ALO algorithm mathematical model.
Table 2. The ALO algorithm mathematical model.
Equation and Description
The random walking of ants in search of space
X ( t ) = 0 , c u s u m 2 r ( t 1 ) 1 , c u s u m 2 r ( t m a x _ I t e r ) 1 (6)
Definition of Stochastic function
r ( t ) = 1   i f   R > 0.5 1   i f   R 0.5 (7)
The normalized random walking of prey in search of space
X i ( t ) = ( X i t a i ) × ( d i t C i t ) ( b i a i ) + C i t (8)
The relationship between the random walks of the antlion and the traps
C i t = A n t l i o n j t + C t (9)
d i t = A n t l i o n j t + d t (10)
The prey ants sliding into the mouth of the antlion
C t = C t I (11)
d t = d t I (12)
I = 10 w t m a x _ i t e r (13)
Capturing the prey and rebuilding the trap in the same position or a new position
A n t l i o n j t = A n t i t   i f   f ( A n t i t ) > f ( A n t l i o n j t ) (14)
Elitism (the fittest antlion of each iteration)
A n t i t = R A t + R E t 2 (15)
Table 3. The parameters of the ALO.
Table 3. The parameters of the ALO.
ParametersValues
Search agents40
Maximum no. of iteration500
Lower bound−50
Upper bound50
Best scoreElite antlion fitness
Best positionElite antlion position
Table 4. PID-PSS parameters of the developed ALO and MPSO.
Table 4. PID-PSS parameters of the developed ALO and MPSO.
GainsMPSO-Based PIDProposed ALO-Based PID *
K P 28.95842.7423
K I 16.983118.9831
K D 12.56038.5603
* The proposed PSS.
Table 5. HIL testing map.
Table 5. HIL testing map.
Test TypesOperating Conditions
Light LoadingNominal LoadingHeavy Loading
Step change 5%Case No. 1Case No. 6Case No. 10
Step change 10%Case No. 2Case No. 7-
Short circuit fault at the beginning of the lineCase No. 3Case No. 8Case No. 11
Short circuit fault at the middle of the lineCase No. 4Case No. 9Case No. 12
Short circuit fault at the end of the lineCase No. 5--
Table 6. Eigenvalues and stability parameters of the SMIB with different PSSs during light loading, compared to the initial condition.
Table 6. Eigenvalues and stability parameters of the SMIB with different PSSs during light loading, compared to the initial condition.
IndicesInitial Operating ConditionPSSs
MPSO-Based PID-PSSALO-Based PID-PSS
λ 1 −963.6785−964.2796−964.0872
λ 2,3 0.0142 ± 5.3557i *−27.2376 ± 21.4542i−26.9996 ± 16.0385i
λ 4,5 −36.7372−0.7802 ± 4.1830i *−1.1504 ± 4.5691i *
λ 6 000
λ 7,8 0−20.00−20.00
ξ in (per unit speed change * MVA/MW)−0.00270.18340.2442
ω n in (rad/s)5.35574.25514.7117
ω d in (rad/s)5.35574.18304.5691
K S in (per unit torque/rad)0.67560.42650.5229
K D in (per unit torque/per unit speed change)−0.210211.547017.0259
* An oscillatory mode.
Table 7. Performance indices at 5% load step change of case study No. (1) during light loading.
Table 7. Performance indices at 5% load step change of case study No. (1) during light loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS5.6141 × 10−87.5682 × 10−44.0514 × 10−77.9489 × 10−3
Proposed ALO-based PID-PSS4.2867 × 10−86.0334 × 10−42.7636 × 10−75.8924 × 10−3
Improvement %23.6420.2831.7925.87
Average improvement %25.4
Δδ
in (electrical radian)
MPSO-based PID-PSS99.227273.36382722.64782002.0799
Proposed ALO-based PID-PSS90.744467.34412298.98901695.0494
Improvement %8.558.2115.5615.34
Average improvement %11.91
Active Power
in (per unit)
MPSO-based PID-PSS4.88401.6280132.5225441.7392
Proposed ALO-based PID-PSS4.49991.4999112.5009375.0011
Improvement %7.877.8715.1115.11
Average improvement %11.49
Positive Voltage B1
in (per unit)
MPSO-based PID-PSS55.183354.72301496.61541484.4879
Proposed ALO-based PID-PSS50.845150.42051270.52491260.2205
Improvement %7.867.8615.1115.11
Average improvement %11.48
Table 8. Performance indices at 10% load step change of case study No. (2) during light loading.
Table 8. Performance indices at 10% load step change of case study No. (2) during light loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS1.8618 × 10−71.2599 × 10−31.2429 × 10−61.2518 × 10−2
Proposed ALO-based PID-PSS1.4160 × 10−79.7225 × 10−48.3481 × 10−78.9427 × 10−3
Improvement %23.9422.8332.8328.56
Average improvement %27.04
Δδ
in (electrical radian)
MPSO-based PID-PSS99.6772 73.5034 2751.3496 2012.3675
Proposed ALO-based PID-PSS 90.8837 67.3758 2318.1264 1701.8895
Improvement %8.828.3415.7515.43
Average improvement %12.08
Active Power
in (per unit)
MPSO-based PID-PSS 4.8842 16.2799 132.5242 441.7405
Proposed ALO-based PID-PSS 4.5002 14.9997 125.0260 375.0021
Improvement %7.8627.8645.65815.108
Average improvement %9.12
Positive Voltage B1
in (per unit)
MPSO-based PID-PSS 55.1986 54.7304 1496.4318 1484.3961
Proposed ALO-based PID-PSS 50.8609 50.4281 1270.3843 1260.1499
Improvement %7.8587.86115.10615.107
Average improvement %11.48
Table 9. Performance indices at three-phase S.C. at the sending end of the TL of case study No. (3) during light loading.
Table 9. Performance indices at three-phase S.C. at the sending end of the TL of case study No. (3) during light loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS1.04 × 10−51.05 × 10−28.71 × 10−51.18 × 10−1
Proposed ALO-based PID-PSS8.97 × 10−69.21 × 10−36.95 × 10−59.85 × 10−2
Improvement %13.6212.4220.2516.88
Average improvement %15.79
Δδ
in (electrical radian)
MPSO-based PID-PSS132.193182.13483055.5092090.912
Proposed ALO-based PID-PSS136.328680.01672841.1411850.655
Improvement %−3.132.587.0211.49
Average improvement %4.49
Active power
in (per unit)
MPSO-based PID-PSS4.966316.3013132.947441.8534
Proposed ALO-based PID-PSS4.583115.0222112.9271375.116
Improvement %7.7167.84715.05915.104
Average improvement %11.43
Positive voltage B1
in (per unit)
MPSO-based PID-PSS54.831154.49371493.3861482.835
Proposed ALO-based PID-PSS50.518450.20431267.9631258.669
Improvement %7.8657.87115.09515.117
Average improvement %11.49
Table 10. Performance indices at single-phase S.C. to ground at the middle of the line of case study No. (4) during light loading.
Table 10. Performance indices at single-phase S.C. to ground at the middle of the line of case study No. (4) during light loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS4.94 × 10−67.28 × 10−34.22 × 10−58.25 × 10−2
Proposed ALO-based PID-PSS4.06 × 10−66.23 × 10−33.20 × 10−56.70 × 10−2
Improvement %17.8214.4324.2418.74
Average improvement %18.81
Δδ
in (electrical radian)
MPSO-based PID-PSS122.399780.0972960.3322072.018
Proposed ALO-based PID-PSS122.402276.85832678.3781812.301
Improvement %0.004.049.5212.53
Average improvement %6.53
Active power
in (per unit)
MPSO-based PID-PSS4.886516.2812132.5237441.7258
Proposed ALO-based PID-PSS4.502915.0019112.5045374.9918
Improvement %7.8507.85815.10615.108
Average improvement %11.48
Positive voltage B1
in (per unit)
MPSO-based PID-PSS54.908454.58241494.371483.347
Proposed ALO-based PID-PSS50.592750.29131268.541259.168
Improvement %7.8607.86215.11215.113
Average improvement %11.49
Table 11. Performance indices at three-phase S.C. to ground fault at the end of the line of case study No. (5) during light loading.
Table 11. Performance indices at three-phase S.C. to ground fault at the end of the line of case study No. (5) during light loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS4.8497 × 10−67.0375 × 10−33.9695 × 10−57.8472 × 10−2
Proposed ALO-based PID-PSS4.2507 × 10−66.1757 × 10−33.2162 × 10−56.5195 × 10−2
Improvement %12.3512.2518.9816.92
Average improvement %15.12
Δδ
in (electrical radian)
MPSO-based PID-PSS117.9186 78.9027 2902.6561 2054.6824
Proposed ALO-based PID-PSS116.762375.30832600.72111788.6808
Improvement %0.984.5610.4012.95
Average improvement %7.22
Active Power
in (per unit)
MPSO-based PID-PSS 4.9262 16.2894 132.7317 441.7767
Proposed ALO-based PID-PSS4.542515.0098112.7113375.0399
Improvement %7.7897.85515.08315.106
Average improvement %11.46
Positive Voltage B1
in (per unit)
MPSO-based PID-PSS 54.9541 54.5859 1495.1461 1483.6358
Proposed ALO-based PID-PSS50.630850.29091269.07891259.3816
Improvement %7.8677.86815.12015.115
Average improvement %11.49
Table 12. Eigenvalues and stability parameters of the SMIB with different PSSs during Nominal loading.
Table 12. Eigenvalues and stability parameters of the SMIB with different PSSs during Nominal loading.
IndicesNormal Operating ConditionPSSs
MPSO-Based PID-PSSALO-Based PID-PSS
λ 1 −966.47668−968.1181−967.5935
λ 2,3 0.1646 ± 5.5081i *−25.0930 ± 38.1149i−24.4835 ± 30.2602i
λ 4,5 −34.2399−1.0417 ± 3.1403i *−1.9135 ± 3.3986i *
λ 6 0−20.00−20.00
λ 7,8 000
ξ in (per unit speed change * MVA/MW)−0.02900.31480.4906
ω n in (rad/s)5.51233.30863.9003
ω d in (rad/s)5.51003.14033.3986
K S in (per unit torque/rad)0.71570.25780.3583
K D in (per unit torque/per unit speed change)−2.368015.417228.3198
* An oscillatory mode.
Table 13. Performance indices at double-line to ground fault at the sending end of the line of case study (8) during nominal loading.
Table 13. Performance indices at double-line to ground fault at the sending end of the line of case study (8) during nominal loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS3.0812 × 10−51.4898 × 10−22.2519 × 10−41.3995 × 10−1
Proposed ALO-based PID-PSS2.6185 × 10−51.2706 × 10−21.7327 × 10−41.1421 × 10−1
Average improvement %17.79
Δδ
in (electrical radian)
MPSO-based PID-PSS 73.4539 59.6633 1612.1012 1505.4544
Proposed ALO-based PID-PSS 79.4138 59.5458 1541.2948 1347.6912
Average improvement %1.74
Active Power
in (per unit)
MPSO-based PID-PSS 49.0806 51.5533 1329.4346 1398.8343
Proposed ALO-based PID-PSS 45.2330 47.5007 1128.6774 1187.5008
Average improvement %11.48
Positive Voltage B1
in (per unit)
MPSO-based PID-PSS 53.7099 53.9568 1460.3439 1466.2232
Proposed ALO-based PID-PSS 49.8328 49.7099 1239.6082 1244.6283
Average improvement %11.33
Table 14. Performance indices at double-phase S.C. on the middle of the line of case study No. (9) during nominal loading.
Table 14. Performance indices at double-phase S.C. on the middle of the line of case study No. (9) during nominal loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS2.10 × 10−51.23 × 10−21.54 × 10−41.15 × 10−1
Proposed ALO-based PID-PSS1.74 × 10−51.03 × 10−21.15 × 10−49.19 × 10−2
Average improvement %19.72
Δδ
in (electrical radian)
MPSO-based PID-PSS68.760258.59911569.7191495.881
Proposed ALO-based PID-PSS72.11557.62631463.1411325.248
Average improvement %3.74
Active Power
in (per unit)
MPSO-based PID-PSS49.030651.55371329.1631398.829
Proposed ALO-based PID-PSS45.181847.50081128.4021187.496
Average improvement %11.48
Positive Voltage B1
in (per unit)
MPSO-based PID-PSS53.711253.97351460.3471466.309
Proposed ALO-based PID-PSS49.482949.72571239.6071244.711
Average improvement %11.49
Table 15. Eigenvalues and stability parameters of the SMIB with different PSSs during heavy loading.
Table 15. Eigenvalues and stability parameters of the SMIB with different PSSs during heavy loading.
IndicesHeavy Loading ConditionPSSs
MPSO-Based PID-PSSALO-Based PID-PSS
λ 1 −966.7365−968.4387−967.8947
λ 2,3 0.1843 ± 5.5145i *−24.9295 ± 38.8812i−24.3121 ± 30.9152i
λ 4,5 −34.0196−1.0449 ± 3.0970i *−1.9342 ± 3.3408i *
λ 6 0−20.00−20.00
λ 7,8 000
ξ in (per unit speed change MVA/MW)−0.03340.31970.5010
ω n in (rad/s)5.51763.26853.8603
ω d in (rad/s)5.51453.09703.3408
K S in (per unit torque/rad)0.71710.25160.3510
K D in (per unit torque/per unit speed change)−2.727615.464528.6262
* An oscillatory mode.
Table 16. Performance indices at three-phase S.C.-to-ground fault at the end of the line of case study no. (10) during light loading.
Table 16. Performance indices at three-phase S.C.-to-ground fault at the end of the line of case study no. (10) during light loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS4.8497 × 10−67.0375 × 10−33.9695 × 10−57.8472 × 10−2
Proposed ALO-based PID-PSS4.2507 × 10−66.1757 × 10−33.2162 × 10−56.5195 × 10−2
Improvement %12.3512.2518.9816.92
Average improvement %15.12
Δδ
in (electrical radian)
MPSO-based PID-PSS117.9186 78.9027 2902.6561 2054.6824
Proposed ALO-based PID-PSS116.762375.30832600.72111788.6808
Improvement %0.984.5610.4012.95
Average improvement %7.22
Active Power
in (per unit)
MPSO-based PID-PSS 4.9262 16.2894 132.7317 441.7767
Proposed ALO-based PID-PSS4.542515.0098112.7113375.0399
Improvement %7.7897.85515.08315.106
Average improvement %11.46
Positive Voltage B1
in (per unit)
MPSO-based PID-PSS 54.9541 54.5859 1495.1461 1483.6358
Proposed ALO-based PID-PSS50.630850.29091269.07891259.3816
Improvement %7.8677.86815.12015.115
Average improvement %11.49
Table 17. Performance indices at three-phase S.C.-to-ground fault at the end of the line of case study no. (11) during light loading.
Table 17. Performance indices at three-phase S.C.-to-ground fault at the end of the line of case study no. (11) during light loading.
SignalPSSISEIAEITSEITAE
Δ ω
in (per unit)
MPSO-based PID-PSS3.61 × 10−51.39 × 10−22.38 × 10−41.27 × 10−1
Proposed ALO-based PID-PSS3.35 × 10−51.22 × 10−22.05 × 10−41.06 × 10−1
Average improvement %12.28
Δδ
in (electrical radian)
MPSO-based PID-PSS64.109756.47231466.2151443.575
Proposed ALO-based PID-PSS68.327755.86181382.3891284.358
Average improvement %2.81
Active Power
in (per unit)
MPSO-based PID-PSS54.478154.26521473.6081472.48
Proposed ALO-based PID-PSS50.215849.99891251.1611250.018
Average improvement %11.47
Positive Voltage B1
in (per unit)
MPSO-based PID-PSS53.623853.88941457.1861464.516
Proposed ALO-based PID-PSS49.405149.64651236.9631243.178
Average improvement %11.49
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

Ibrahim, N.M.A.; Talaat, H.E.A.; Shaheen, A.M.; Hemade, B.A. Optimization of Power System Stabilizers Using Proportional-Integral-Derivative Controller-Based Antlion Algorithm: Experimental Validation via Electronics Environment. Sustainability 2023, 15, 8966. https://doi.org/10.3390/su15118966

AMA Style

Ibrahim NMA, Talaat HEA, Shaheen AM, Hemade BA. Optimization of Power System Stabilizers Using Proportional-Integral-Derivative Controller-Based Antlion Algorithm: Experimental Validation via Electronics Environment. Sustainability. 2023; 15(11):8966. https://doi.org/10.3390/su15118966

Chicago/Turabian Style

Ibrahim, Nader M. A., Hossam E. A. Talaat, Abdullah M. Shaheen, and Bassam A. Hemade. 2023. "Optimization of Power System Stabilizers Using Proportional-Integral-Derivative Controller-Based Antlion Algorithm: Experimental Validation via Electronics Environment" Sustainability 15, no. 11: 8966. https://doi.org/10.3390/su15118966

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