Next Article in Journal
Approximate Closed-Form Solutions for a Class of 3D Dynamical Systems Involving a Hamilton–Poisson Part
Next Article in Special Issue
A Stackelberg Game-Based Model of Distribution Network-Distributed Energy Storage Systems Considering Demand Response
Previous Article in Journal
Point-Rich: Enriching Sparse Light Detection and Ranging Point Clouds for Accurate Three-Dimensional Object Detection
Previous Article in Special Issue
Six-Segment Strategy for Prosumers’ Financial Benefit Maximization in Local Peer-to-Peer Energy Trading
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Balanced Arithmetic Optimization Algorithm-Optimized Controller for Enhanced Voltage Regulation

1
Department of Computer Engineering, Batman University, 72100 Batman, Turkey
2
Institute of Postgraduate Studies, Batman University, 72100 Batman, Turkey
3
Electrical-Electronics Engineering Department, Tarsus University, 33400 Tarsus, Turkey
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(23), 4810; https://doi.org/10.3390/math11234810
Submission received: 5 November 2023 / Revised: 23 November 2023 / Accepted: 27 November 2023 / Published: 28 November 2023
(This article belongs to the Special Issue Mathematical Modeling and Optimization of Energy Systems)

Abstract

:
This work introduces an innovative approach that unites a PIDND2N2 controller and the balanced arithmetic optimization algorithm (b-AOA) to enhance the stability of an automatic voltage regulator (AVR) system. The PIDND2N2 controller, tailored for precision, stability, and responsiveness, mitigates the limitations of conventional methods. The b-AOA optimizer is obtained through the integration of pattern search and elite opposition-based learning strategies into the arithmetic optimization algorithm. This integration optimizes the controller parameters and the AVR system’s response, harmonizing exploration and exploitation. Extensive assessments, including evaluations on 23 classical benchmark functions, demonstrate the efficacy of the b-AOA. It consistently achieves accurate solutions, exhibits robustness in addressing a wide range of optimization problems, and stands out as a promising choice for various applications. In terms of the AVR system, comparative analyses highlight the superiority of the proposed approach in transient response characteristics, with the shortest rise and settling times and zero overshoot. Additionally, the b-AOA approach excels in frequency response, ensuring robust stability and a broader bandwidth. Furthermore, the proposed approach is compared with various state-of-the-art control methods for the AVR system, showcasing an impressive performance. These results underscore the significance of this work, setting a new benchmark for AVR control by advancing stability, responsiveness, and reliability in power systems.

1. Introduction

In the realm of power systems, the automatic voltage regulator (AVR) stands as a linchpin, ensuring that connected electrical equipment functions within prescribed voltage bounds [1]. The consequences of inadequate voltage regulation can be profound, from equipment damage and operational failures to costly downtime and extensive repairs [2,3,4,5]. Consequently, the AVR plays a pivotal role in power systems reliant on generators or alternators for electricity generation [6,7]. While existing control methodologies have achieved some success, they remain encumbered by limitations [8], including challenges related to robustness, overshoots, rise times, settling times, and persistent steady-state errors.
The motivation driving our study is rooted in a collective commitment to surpass these limitations and contribute to the evolution of more robust and efficient power systems. Our primary objective is to introduce an advanced control scheme capable of effectively addressing these challenges. To accomplish this, we have innovatively integrated a novel optimizer, rooted in the arithmetic optimization algorithm (AOA) [9], meticulously fine-tuned to enhance the parameters of our proposed control scheme and, by extension, its overall performance and adaptability.
In the existing landscape of AVR control, controllers have emerged as indispensable assets for the vigilant monitoring and regulation of the AVR itself [10]. These controllers serve as hubs, facilitating real-time adjustments to maintain voltage stability, enabling remote monitoring, fault detection, and automatic shutdown during emergencies, and enhancing the overall system dependability. A range of controllers, from the standard proportional–integral–derivative (PID) to more advanced variants like the PID Acceleration (PIDA), fractional-order PID (FOPID), and PID with a second-order derivative (PIDD2), offer diverse attributes to meet the specific requirements of AVR control [11,12,13,14,15,16,17,18].
However, the choice of controller alone is insufficient to address the complex challenges faced by AVR systems. The choice of a cost function is equally crucial, as it significantly impacts performance [19]. Researchers employ various cost functions, such as the integral of time-weighted squared error, integral of squared error, integral of absolute error, and the dynamic response performance criteria-based Zwe-Lee Gaing (ZLG) cost function [20,21,22]. In this context, our work introduces a novel approach that unites both the controller and the optimizer to form a comprehensive solution for enhancing AVR stability. The core innovation is the balanced arithmetic optimization algorithm (b-AOA). It marries the powerful pattern search (PS) strategy [23], renowned for its exploitation capabilities, with the elite opposition-based learning (EOBL) strategy [24], elevating exploration. This marriage optimizes the controller parameters and the AVR system’s response, harmonizing exploration and exploitation to attain a level of stability previously out of reach.
The efficacy of the b-AOA is first verified through comprehensive assessments against 23 classical unimodal, multimodal, and fixed-dimensional multimodal benchmark functions. These evaluations compare the effectiveness of the proposed b-AOA to other optimization algorithms, including the original AOA [9], sine cosine algorithm [25], weighted mean of vectors algorithm [26], and marine predators algorithm [27]. The results from the benchmark functions underscore the remarkable performance of the b-AOA. It consistently achieves mean errors close to zero, demonstrating its capability to find accurate solutions. Furthermore, its robustness and consistency make it a strong candidate for addressing a wide range of optimization problems.
In the case of the AVR system, we firstly introduce a PIDND2N2 controller designed for enhanced precision, stability, and responsiveness in voltage regulation. This configuration mitigates the limitations associated with conventional methods, promising a superior control performance. Secondly, the b-AOA optimizer fine-tunes the parameters of our proposed control scheme, improving its overall performance and adaptability. Using the ZLG cost function [28], we target the minimization of dynamic response performance criteria, such as maximum overshoot, steady-state error, settling time, and rise time, thereby ensuring that the AVR system meets the most stringent performance requirements. Our work seeks to transcend theoretical innovation, anchoring itself in the practical applicability of power systems, where stability and reliability are non-negotiable. Through extensive simulations and rigorous experimentation, we aim to demonstrate the superiority of the b-AOA-based AVR system in comparison to existing control and optimization techniques. Our focus on stability, speed of response, robustness, and efficiency aligns with the motivations presented, making our work a substantial contribution to the field of power system control.
To validate the superiority of the proposed b-AOA approach, we conducted extensive comparative analyses, evaluating its performance against well-established control methodologies, such as the sine cosine algorithm-based PID controller [29], whale optimization algorithm-based PIDA controller [30], slime mould algorithm-based FOPID controller [31], and particle swarm optimization-based PIDD2 controller [32]. The results unequivocally demonstrate that the b-AOA-based approach outshines its counterparts. It exhibits unmatched transient response characteristics, with the shortest rise time (0.033485 s) and settling time (0.050752 s) while eliminating overshoot. In contrast, other methods exhibit less favorable response characteristics. In terms of frequency response, the b-AOA approach consistently excels, showcasing robust stability, favorable gain margins, and a broader bandwidth.
To further assess the effectiveness of the proposed approach, we compared it with several other established controller approaches reported in the literature. These included several recently reported control methods for the AVR system. These methods include a variety of controllers, each tuned using different optimization algorithms such as the marine predators algorithm-based FOPID [33], hybrid atom search particle swarm optimization-based PID [34], equilibrium optimizer-based TIλDND2N2-based controller [35], reptile search algorithm-based FOPIDD2 [11], improved Runge–Kutta algorithm-based PIDND2N2 [36], symbiotic organism search algorithm-based PID-F [37], whale optimization algorithm-based 2DOF FOPI [38], Lévy flight-based reptile search algorithm with local search ability-based PID [39], chaotic black widow algorithm-based FOPID [20], genetic algorithm-based fuzzy PID [40], sine cosine algorithm-based FOPID with fractional-order filter [41], hybrid simulated annealing–Manta ray foraging optimization algorithm-based PIDD2 [42], slime mould algorithm-based PID [43], gradient-based optimization-based FOPID [44], and nonlinear sine cosine algorithm-based sigmoid PID [45]. We evaluate their transient response performance to assess the effectiveness of the proposed approach. The results demonstrate the efficacy of the b-AOA-based PIDND2N2 controller in comparison to various state-of-the-art methods as it stands out with an impressive performance, suggesting the exceptional stability and responsiveness of the b-AOA-tuned controller.
In summary, our work presents a superior solution to address the challenges in AVR control, contributing to the advancement of power systems while establishing a new benchmark for stability, responsiveness, and reliability in this critical domain. The unique integration of the b-AOA with the PIDND2N2 controller signifies a significant leap forward in achieving optimal voltage regulation and stability in power systems.

2. Overview of Arithmetic Optimization Algorithm

The arithmetic optimization algorithm (AOA) draws inspiration from arithmetic principles [9] to construct a versatile metaheuristic optimization technique. It initiates the optimization process by generating a set of randomized solutions represented as follows.
X = x 1,1 x 1 , j x 1 , n 1 x 1 , n x 2,1 x 2 , j x 2 , n x N 1,1 x N 1 , j x N 1 , n x N , 1 x N , j x N , n 1 x N , n
Following this, the algorithm employs a function known as “Math Optimizer Accelerated” (MopA) to execute exploration and exploitation tasks. The MopA function is defined as:
M o p A t = M i n + t × M a x M i n t m a x
where t represents the current iteration, t m a x denotes the maximum number of iterations, and M i n and M a x represent the minimum and maximum values of the accelerated function. The exploration phase of the algorithm is carried out when r 1 > M o p A , where r 1 is a randomly generated number. During exploration, the multiplication ( M u l t ) and division ( D i v ) operators are employed, defined as follows:
x i , j t + 1 = b e s t x j × M o p P × U B j L B j × μ + L B j ,   for   r 2 > 0.5 b e s t x j ÷ M o p P + ϵ × U B j L B j × μ + L B j ,   for   r 2 < 0.5
where x i , j t represents the j t h position of solution I at the current iteration, x i t + 1 denotes the solution of i in the next iteration, b e s t x j signifies the best solution’s j t h position obtained so far, ϵ is a small integer, μ is a control parameter that adjusts the search process, and U B j and L B j , respectively, represent the upper and lower bounds of the j t h position. The “Math optimizer probability” function, denoted by MopP, is computed as follows, with α reflecting the exploitation accuracy through iterations.
M o p P t = 1 ( t ) 1 / α ( t m a x ) 1 / α
The term r 2 is another random number utilized for position updates. The M u l t operator is employed for r 2 > 0.5 , while the D i v operator is used otherwise. Conversely, the exploitation phase occurs when r 1 < M o p A . In this stage, the addition ( A d d ) and subtraction ( S u b ) operators are utilized, defined as:
x i , j t + 1 = b e s t x j + M o p P × U B j L B j × μ + L B j ,   for   r 3 > 0.5 b e s t x j M o p P × U B j L B j × μ + L B j ,   for   r 3 < 0.5
Here, r 3 is a random number determining whether the A d d or S u b operation is applied. A d d operates when r 3 > 0.5 , while S u b is used for r 3 < 0.5 . Figure 1 presents a comprehensive flowchart of the AOA, depicting its intricate process.

3. Balanced Arithmetic Optimization Algorithm

The balanced AOA (b-AOA) is an evolution of the pattern search (PS) [46] and the opposition-based learning (OBL) [47] schemes, designed to enhance both exploration and exploitation capabilities in the context of metaheuristic optimization. This section provides a step-by-step explanation of the b-AOA’s development and its core components.
The PS scheme serves as the foundation for the b-AOA. It is a derivative-free algorithm known for its exploitation capabilities [48]. PS starts with an initial point ( X 0 ) defined by the user and proceeds by generating a mesh around this point, gradually updating the mesh as new points with lower objective function values are discovered. The process involves the following key steps:
  • Exploration Stage: If a new point with a lower objective function value ( f X 1 < f X 0 ) is found (successful poll), it becomes the source point. The mesh size is then expanded by a factor of 2, creating new points for exploration.
  • Exploitation Stage: When no new points with lower values are discovered, the mesh size is reduced by multiplying it by 0.5 (reduction factor). This contraction stage continues until the termination condition is met.
The detailed flowchart of the PS scheme is illustrated in Figure 2.
The OBL scheme, introduced as a machine-learning technique, enhances the performance of metaheuristic algorithms by considering both the current individuals and their opposites [47]. A special type of OBL mechanism known as elite OBL (EOBL) [49] focuses on the elite (best) individuals in combination with the current individuals. EOBL generates opposite solutions for the elite individuals, which are then evaluated for their fitness. The mathematical representation of the EOBL strategy is as follows:
  • Elite candidate solution: X = x 1 ,   x 2 , x k with k decision variables.
  • Elite opposition-based solution: X o = x 1 o , x 2 o , x k o where x i o = δ d a i + d b i x i and δ is a parameter within the range (0, 1) controlling the opposition magnitude.
  • Dynamic boundaries: d a i = m i n x i , d b i = m a x x i .
  • To ensure that opposite decision variables stay within the boundaries [ L b i ,   U b i ], the following rule is applied: x i o = r a n d L b i , U b i , if x i o < L b i or x i o > U b i .
The working principle of the OBL mechanism is depicted in Figure 3.
The b-AOA integrates the EOBL and the PS schemes to achieve a balanced approach with improved exploration and exploitation capabilities. Figure 4 provides an overview of the b-AOA’s operation. As depicted in the figure:
  • The algorithm commences with the original AOA and generates the best solution.
  • The EOBL scheme is introduced to produce N best solutions.
  • The PS scheme takes over to enhance exploitation, running a total of 5 times with 100 × D iterations, where D represents the problem’s dimension size.
The parameters for the b-AOA, derived from extensive simulations, include:
  • PS scheme parameters: initial mesh size = 1, mesh expansion factor = 2, mesh contraction factor = 0.5, and all tolerances = 10−6.
  • AOA parameters: sensitive parameter α = 5 , control parameter μ = 0.4975 , M i n = 0.2 , M a x = 1 .

4. Adopted Test Functions

4.1. Unimodal Benchmark Functions

In this section, we introduce a set of unimodal benchmark functions that have been adopted for our study. These benchmark functions are commonly used in the field of optimization to evaluate the performance of optimization algorithms and to compare their effectiveness. The following unimodal benchmark functions have been selected for our analysis: Sphere, Schwefel 2.2, Schwefel 1.2, Schwefel 2.21, Rosenbrock, Step, and Quartic. The mathematical equations defining these benchmark functions are, respectively, provided in the following equations.
F u n c 1 ( x ) = i = 1 D i m x i 2
F u n c 2 ( x ) = i = 1 D i m x i + i = 1 D i m x i
F u n c 3 ( x ) = i = 1 D i m j = 1 i x j 2
F u n c 4 ( x ) = max i x i , 1 i D i m
F u n c 5 ( x ) = i = 1 D i m 1 100 ( x i + 1 x i 2 ) 2 + ( x i 1 ) 2
F u n c 6 ( x ) = i = 1 D i m ( x i + 0.5 ) 2
F u n c 7 ( x ) = i = 1 D i m i x i 4 + r a n d o m [ 0 , 1 )
In addition to the mathematical expressions for these benchmark functions, we have compiled the essential properties and information related to each function in Table 1. This table provides details on the dimensionality of each function, the evaluation interval, and the global minimum values. These properties are crucial for understanding the characteristics of each benchmark function and for conducting a comprehensive analysis of the optimization algorithms’ performance.

4.2. Multimodal Benchmark Functions

In this section, we introduce a collection of multimodal benchmark functions that have been selected for our study. Multimodal functions are essential for assessing the capability of optimization algorithms to handle complex, non-convex search spaces, where multiple local optima exist. The following multimodal benchmark functions have been chosen for our study: Schwefel, Rastrigin, Ackley, Griewank, Penalized, and Penalized2. The mathematical equations representing these benchmark functions are, respectively, provided in the following equations.
F u n c 8 ( x ) = i = 1 D i m x i s i n | x i |
F u n c 9 ( x ) = i = 1 D i m 10 + x i 2 10 cos 2 π x i
F u n c 10 ( x ) = 20 + e 20 e x p 0.2 1 D i m i = 1 D i m x i 2 exp 1 D i m i = 1 D i m cos 2 π x i
F u n c 11 ( x ) = i = 1 D i m x i 2 4000 i = 1 D i m cos x i i + 1
F u n c 12 ( x ) = π D i m 10 sin 2 ( π y 1 ) + i = 1 D i m 1 ( y i 1 ) 2 1 + 10 sin 2 ( π y i + 1 ) + ( y D i m 1 ) 2 + i = 1 D i m u ( x i , 10 , 100 , 4 )
F u n c 13 ( x ) = 0.1 sin 2 ( 3 π x 1 ) + i = 1 D i m 1 ( x i 1 ) 2 1 + sin 2 ( 3 π x i + 1 ) + ( x D i m 1 ) 2 [ 1 + sin 2 ( 2 π x D i m ) ] + i = 1 D i m u ( x i , 5 , 100 , 4 )
Moreover, to facilitate a comprehensive understanding of these benchmark functions, we have compiled the vital properties and information for each function in Table 2. This table presents information on the dimensionality of each function, the evaluation interval, and the global minimum values. These properties are pivotal for comprehending the characteristics of each benchmark function and for the subsequent analysis of optimization algorithms in handling multimodal search spaces.

4.3. Fixed-Dimensional Multimodal Test Functions

This section introduces a collection of fixed-dimensional multimodal test functions, which are indispensable for evaluating the performance of optimization algorithms in solving problems with known characteristics. These benchmark functions are specifically selected for their fixed-dimensional nature, making them suitable for comparing and benchmarking various optimization techniques. The following fixed-dimensional multimodal test functions have been included in our study: Foxholes, Kowalik, Six-Hump Camel, Branin, Goldstein–Price, Hartman 3, Hartman 6, Shekel 5, Shekel 7, and Shekel 10. The mathematical equations describing these fixed-dimensional multimodal test functions can, respectively, be found in the following equations.
F u n c 14 ( x ) = 1 1 500 + j = 1 25 1 j + i = 1 2 x i a i j 6
F u n c 15 ( x ) = i = 1 11 a i x 1 b i 2 + b i x 2 b i 2 + b i x 3 + x 4 2
F u n c 16 ( x ) = 4 x 1 2 2.1 x 1 4 + 1 3 x 1 6 + x 1 x 2 4 x 2 2 + 4 x 2 4
F u n c 17 ( x ) = 10 1 1 8 π cos x 1 + 10 + x 2 5.1 4 π 2 x 1 2 + 5 π x 1 6 2
F u n c 18 ( x ) = 30 + 2 x 1 3 x 2 2 × 18 32 x 1 + 12 x 1 2 + 48 x 2 36 x 1 x 2 + 27 x 2 2 × 1 + x 1 + x 2 + 1 2 19 14 x 1 + 3 x 1 2 14 x 2 + 6 x 1 x 2 + 3 x 2 2
F u n c 19 ( x ) = i = 1 4 c i e x p j = 1 3 a i j x j p i j 2
F u n c 20 ( x ) = i = 1 4 c i e x p j = 1 6 a i j x j p i j 2
F u n c 21 ( x ) = i = 1 5 1 X a i ( X a i ) T + c i
F u n c 22 ( x ) = i = 1 7 1 X a i ( X a i ) T + c i
F u n c 23 ( x ) = i = 1 10 1 X a i ( X a i ) T + c i
To further enhance the understanding of these benchmark functions, the essential properties and information for each function are summarized in Table 3. This table provides key details such as the dimensionality of each function, the evaluation interval, and the global minimum values, enabling a comprehensive evaluation and comparison of optimization algorithms for fixed-dimensional search spaces.

5. Statistical Performance of the b-AOA on Test Functions

5.1. Compared Algorithms

In our study, we have evaluated the performance of several optimization algorithms against the proposed b-AOA by comparing their effectiveness in solving the benchmark functions. The algorithms considered for comparison include the following: original AOA [9], sine cosine algorithm (SCA) [25], weighted mean of vectors (INFO) algorithm [26], and marine predators algorithm (MPA) [27].
For each of these algorithms, we conducted 30 independent runs to ensure a robust and comprehensive assessment. By executing multiple independent runs, we aimed to account for the inherent variability in optimization processes and obtain reliable results.
Table 4 presents the key properties and control parameters associated with the compared algorithms. These properties include the population size, total iteration number, and values of other control parameters specific to each algorithm.

5.2. Statistical Results Obtained from Unimodal Benchmark Functions

In this section, we present the comparative statistical results obtained from the evaluation of the unimodal benchmark functions using various optimization algorithms. The analysis is based on the mean, standard deviation, best, and worst performance of each algorithm across the unimodal benchmark functions. Table 5 summarizes these results. Upon examining the data in Table 5, we can draw several important observations:
  • Sphere Function: The b-AOA demonstrates a superior performance, achieving a mean error of zero across multiple runs. In contrast, other algorithms exhibit varying degrees of error, with the AOA achieving the lowest mean error but still far from the precision of the b-AOA.
  • Schwefel 2.2 Function: Similar to the Sphere function, the b-AOA outperforms other algorithms by achieving a mean error close to zero. The other algorithms, in contrast, exhibit significant errors.
  • Schwefel 1.2 and Schwefel 2.21 Functions: In both cases, the b-AOA once again stands out with extremely low mean errors, indicating its effectiveness in solving these functions. The other algorithms show larger mean errors.
  • Rosenbrock Function: While the b-AOA exhibits a higher mean error compared to some other algorithms, it still achieves competitive results, and its worst-case performance is better than some other algorithms. It is important to note that the Rosenbrock function is known for its challenging optimization landscape.
  • Step Function: The b-AOA demonstrates exceptional performance with a mean error close to zero. The other algorithms exhibit more significant errors, making the b-AOA the most effective choice for this function.
  • Quartic Function: Once again, the b-AOA shows a strong performance, with a mean error significantly lower than other algorithms. It is evident that the b-AOA consistently performs exceptionally well across multiple unimodal benchmark functions.
In summary, the results obtained from the unimodal benchmark functions highlight the efficacy of the proposed b-AOA. It consistently achieves mean errors close to zero, demonstrating its capability to find accurate solutions. While some other algorithms perform well on specific functions, the b-AOA stands out as a robust choice across various unimodal benchmark functions, making it a promising optimization algorithm for solving such problems.

5.3. Statistical Results Obtained from Multimodal Benchmark Functions

In this section, we analyze the comparative statistical results obtained from the evaluation of the multimodal benchmark functions using various optimization algorithms. The statistical metrics considered include the mean, standard deviation, best, and worst performance for each algorithm across the multimodal benchmark functions. Table 6 summarizes these results. Upon examining the data in Table 6, several key observations can be made:
  • Schwefel Function: The b-AOA exhibits a mean error of −12,536, which is notably closer to the global minimum of this multimodal function. It also achieves the lowest standard deviation, indicating a high level of consistency in its performance. The worst-case result is still very competitive, showing the effectiveness of the b-AOA in solving the Schwefel function.
  • Rastrigin Function: Interestingly, for the Rastrigin function, all algorithms, including the b-AOA, achieve a mean error of zero. While the b-AOA does not stand out in this case, it demonstrates a comparable performance to other algorithms.
  • Ackley Function: For the Ackley function, the b-AOA achieves a mean error close to zero, indicating its effectiveness in minimizing the function. The standard deviation is also very low, demonstrating consistent results.
  • Griewank Function: Similar to the Rastrigin function, all algorithms, including the b-AOA, achieve a mean error of zero. While the b-AOA performs equally well in terms of mean error, its consistency is reflected in a lower standard deviation.
  • Penalized and Penalized2 Functions: The b-AOA outperforms other algorithms in minimizing both the Penalized and Penalized2 functions, as indicated by the lower mean error. Its consistent performance is highlighted by the low standard deviation, making it a robust choice for solving these multimodal functions.
In summary, the results obtained from the multimodal benchmark functions emphasize the efficacy of the proposed b-AOA. It not only achieves competitive mean errors but also demonstrates remarkable consistency in its performance, as reflected by the low standard deviations. This consistency is essential for solving complex multimodal functions where the optimization landscape can be highly challenging. The b-AOA’s ability to approach the global minimum and its robustness make it a strong candidate for addressing multimodal optimization problems.

5.4. Statistical Results Obtained from Fixed-Dimensional Multimodal Benchmark Functions

This section provides an analysis of the comparative statistical results obtained from the evaluation of fixed-dimensional multimodal benchmark functions using various optimization algorithms. The data in Table 7 present the mean, standard deviation, best, and worst performance of each algorithm for these functions. Key insights drawn from the data in Table 7 include:
  • Foxholes Function: The b-AOA stands out as it achieves a mean error of 0.998, which is very close to the global minimum of this function. Moreover, it demonstrates an extremely low standard deviation, indicating remarkable consistency. The best- and worst-case performance metrics further underscore its effectiveness in solving the Foxholes function.
  • Kowalik Function: The b-AOA once again excels, achieving a mean error of 0.00030749, which is impressively close to the global minimum. The standard deviation is nearly zero, highlighting its exceptional consistency. In contrast, other algorithms exhibit higher mean errors and standard deviations.
  • Six-Hump Camel Function: The b-AOA performs exceptionally well, achieving a mean error close to the global minimum and an almost negligible standard deviation. This indicates its strong capability to solve the Six-Hump Camel function effectively.
  • Branin Function: The b-AOA continues to demonstrate outstanding performance with a mean error of 0.39789, very close to the global minimum. It also exhibits an absence of standard deviation, showcasing the consistency of its results.
  • Goldstein–Price Function: The b-AOA delivers optimal performance by achieving a mean error of 3. This not only aligns with the global minimum but is also consistent without any standard deviation. This makes it a standout performer for the Goldstein–Price function.
  • Hartman 3, Hartman 6, Shekel 5, Shekel 7, and Shekel 10 Functions: Across all of these functions, the b-AOA consistently achieves a mean error close to the global minimum, with negligible standard deviations. This underscores its efficacy in solving these fixed-dimensional multimodal benchmark functions.
In summary, the results obtained from the fixed-dimensional multimodal benchmark functions highlight the efficacy of the proposed b-AOA. It consistently delivers mean errors close to the global minimum and demonstrates exceptional consistency with minimal standard deviations. This performance makes the b-AOA a robust choice for solving a wide range of fixed-dimensional multimodal functions, showing its potential as a versatile optimization algorithm.

6. Automatic Voltage Regulator System

6.1. Components of AVR System and Its Modeling

The AVR system comprises four main components: the exciter, generator, sensor, and amplifier, each of which plays a crucial role in the system’s performance. Figure 5 illustrates the schematic diagram of a typical AVR system, providing an overview of its structural components.
To model the AVR system effectively, it is essential to define the transfer functions and constraints for each of these components. The transfer function for the amplifier is given by:
G a m p = K a m p 1 + s T a m p
which is subjected to constraints of 10 K a m p 40 ve 0.02 T a m p 0.1 . The transfer function for the exciter is represented as:
G e x c = K e x c 1 + s T e x c
with constraints of 1 K e x c 10 ve 0.4 T e x c 1 . The generator’s transfer function is defined as:
G g e n = K g e n 1 + s T g e n
which has constraints of 0.7 K g e n 1 ve 1 T g e n 2 . The sensor’s transfer function is presented as:
H s e n = K s e n 1 + s T s e n
which is constrained by 0.9 K s e n 1.1 ve 0.001 T s e n 0.06 . To facilitate a fair comparison with the literature reports, specific parameter values of K a m p = 10 , T a m p = 0.1   s , K e x c = 1 , T e x c = 0.4   s , K g e n = 1 , T g e n = 1   s , K s e n = 1 ve T s e n = 0.01   s [29,30,31,32] are employed in this study. By applying these parameter values, the transfer function for an uncontrolled AVR system can be derived as follows.
T n o c o n t r o l s = G a m p s × G e x c s × G g e n s 1 + G a m p s × G e x c s × G g e n s × H s e n s = 0.1 s + 10 0.0004 s 4 + 0.0454 s 3 + 0.555 s 2 + 1.51 s + 11
An uncontrolled AVR system, with its main components, is illustrated in Figure 6.

6.2. Pole–Zero Map of an Uncontrolled AVR System

The pole–zero map of the uncontrolled AVR system is depicted in Figure 7. The system’s poles are located at −99.9712, −12.4892, and −0.5198 ± 4.6642i, while it possesses only one zero at −100. The system exhibits a very low damping ratio (11.1%) for complex poles, indicating the necessity for enhancing the performance of the uncontrolled system.

6.3. Time-Domain Response of an Uncontrolled AVR System

The unit step response of the uncontrolled AVR system is illustrated in Figure 8. The relevant system exhibits a maximum overshoot of 65.7226%, a rise time of 0.2607 s, a settling time of 6.9865 s, and a peak time of 0.7522 s. These values are considerably large for a power system, and the proposed control approach aims to enhance the performance of the AVR system.

6.4. Open-Loop Frequency Response of an Uncontrolled AVR System

Figure 9 displays the Bode plot of the uncontrolled open-loop AVR system. This system exhibits a gain margin of 4.6176 dB, a phase margin of 16.1028 degrees, and a bandwidth of 6.9454 rad/s. Just as with the time response criteria, it is evident that the frequency response criteria also require improvement through an effective control approach.

7. The Proposed Novel Design Method for AVR System

7.1. Reported Controller Types and PIDND2N2 Controller

In the context of the AVR system, several controller types have been reported and applied. These controllers play a critical role in regulating and stabilizing the system’s voltage. The transfer functions of some of the most commonly reported controllers, including PID, PIDA, FOPID, PIDD2, and PIDND2N2, are, respectively, provided in Equations (34)–(38) [50,51].
C P I D s = K p + K i s + K d s
C P I D A s = K a s 3 + K d s 2 + K p s + K i s 3 + α s 2 + β s
C F O P I D s = K p + K i s λ + K d s μ
C P I D D 2 s = K p + K i s + K d 1 s + K d 2 s 2
C P I D N D 2 N 2 s = K p + K i s + K d 1 n 1 s s + n 1 + K d 2 n 2 s 2 s + n 2 2
In the specific context of the AVR system, these controller types have been employed to regulate the voltage. For this study, the PIDND2N2 controller was selected due to its effectiveness in achieving the desired control objectives. The implementations were carried out on MATLAB/Simulink environment that is installed on a windows computer with 12th generation Intel i5-12400, 2.50 GHz processor and 16.00 GB RAM. A visual representation of the PIDND2N2 controller can be found in Figure 10.

7.2. Objective Function

In the literature, it is feasible to encounter commonly used error-based objective functions of F I A E (integral of the absolute error), F I S E (integral of the square of the error), F I T A E (integral of time-weighted absolute error), and F I T S E (integral of time-weighted square of the error). Their definitions are provided in the following equations [52].
F I A E = 0 | e ( t ) | d t
F I S E = 0 ( e ( t ) ) 2 d t
F I T A E = 0 t | e ( t ) | d t
F I T S E = 0 t ( e ( t ) ) 2 d t
In Equations (39)–(42), the term e t represents the error signal, which, for the AVR system, is defined as e t = V t V r e f where V t is the terminal voltage and V r e f is the reference voltage. Additionally, the F Z L G objective function, which utilizes time response performance criteria, is widely employed in the literature. In this study, the latter one has been preferred, which is given by the following equation [53].
F Z L G = 1 θ M P + E S + θ ( S T R T )
In Equation (43), M P denotes the maximum overshoot, E S stands for the steady-state error, S T represents the settling time, and R T signifies the rise time. The parameter θ in the equation serves as a weighting factor and is set to e 1 = 0.3679 in this study.

7.3. Integration of the Algorithm to PIDND2N2-Controlled AVR System

Table 8 lists the lower and upper boundary values used for the controller parameters when applying the proposed b-AOA to the PIDND2N2 controller. These values are utilized during the optimization process to determine the range within which parameter values should be sought.
In Figure 11, we present a block diagram that outlines the application of our proposed b-AOA in conjunction with the PIDND2N2 controller and the ZLG objective function within the AVR system. This block diagram encapsulates the essential steps involved in optimizing the controller parameters. The optimization process begins with the initialization of the b-AOA. This involves setting up the initial population of parameter sets for the PIDND2N2 controller. These initial values serve as starting points for the algorithm. The ZLG objective function is a dynamic response performance criteria-based cost function employed in our work. It encapsulates various metrics, including maximum overshoot, steady-state error, settling time, and rise time. The proposed algorithm evaluates the performance of each set of controller parameters by calculating the ZLG cost based on the system’s dynamic response. The b-AOA iteratively refines the controller parameters by considering the performance evaluated through the ZLG objective function. During each iteration, the algorithm employs a combination of PS and EOBL strategies to explore and exploit the parameter space effectively. The algorithm adjusts the PIDND2N2 controller parameters based on the outcomes of the ZLG objective function evaluation. The adjustment aims to minimize the cost function, indicating the improved dynamic response characteristics of the AVR system. The optimization process continues until reaching the total number of iterations. The final values of the PIDND2N2 controller parameters, which correspond to the lowest ZLG cost achieved during the optimization process, are identified as the optimal or best parameters for the given AVR system.

8. Simulation Results and Discussion

8.1. Statistical Performance of b-AOA and AOA Methods for AVR System

In the optimization of the AVR system, the b-AOA and AOA were executed 30 times. A population size of 30 and a maximum iteration count of 50 were chosen for minimizing the objective function. The statistical results obtained from all runs are presented in Table 9. As observed in the table, all statistical metrics for optimizing the F Z L G objective function favor the b-AOA, indicating its superior performance. These results additionally confirm the statistical stability of the b-AOA.

8.2. Obtained Best Controller Parameters and Transfer Functions of the Optimized System

In this section, we discuss the results regarding the best controller parameters and the corresponding transfer functions of the optimized system. Figure 12 provides the convergence curve, illustrating the progress of the b-AOA and the original AOA in minimizing the objective function. Notably, it shows that the b-AOA outperforms the original AOA by achieving the lowest objective function value through iterations.
Table 10 presents the optimal parameters of the PIDND2N2 controller, obtained using both the b-AOA and the original AOA.
Using those values would yield the following transfer functions of the optimized systems for the original AOA and proposed b-AOA.
T A O A P I D N D 2 N 2 s = 1.015 e 04 s 5 + 1.674 e 07 s 4 + 1.77 e 09 s 3 + 2.026 e 10 s 2 + 4.661 e 10 s + 2.486 e 10 0.0004 s 8 + 1.36 s 7 + 1531 s 6 + 6.279 e 05 s 5 + 5.621 e 07 s 4 + 2.229 e 09 s 3 + 2.157 e 10 s 2 + 4.754 e 10 s + 2.486 e 10
T b A O A P I D N D 2 N 2 s = 5.876 e 04 s 5 + 1.003 e 08 s 4 + 1.057 e 10 s 3 + 1.157 e 11 s 2 + 3.035 e 11 s + 1.255 e 11 0.0004 s 8 + 2.26 s 7 + 4322 s 6 + 2.944 e 06 s 5 + 2.929 e 08 s 4 + 1.29 e 10 s 3 + 1.22 e 11 s 2 + 3.084 e 11 s + 1.255 e 11

8.3. Stability of the Proposed Design Method

In this section, we analyze the stability of the proposed design method by evaluating the step response and open-loop frequency response of the b-AOA- and AOA-tuned PIDND2N2 controllers.
Figure 13 and Table 11 present the transient response performance metrics for the b-AOA- and AOA-tuned PIDND2N2 controllers. The step response of both controllers is observed concerning the change in the terminal voltage. As illustrated in Figure 13, the b-AOA-tuned PIDND2N2 controller exhibits a faster rise time and settling time with zero overshoot compared to the AOA-tuned PIDND2N2 controller. This implies that the b-AOA-tuned system reaches the desired state more rapidly without oscillations, demonstrating its superior stability in the time domain. The numerical results from Table 11 confirm these visual observations.
Figure 14 and Table 12 present the open-loop Bode diagrams and frequency response performance metrics for the controllers. In the frequency domain, the b-AOA-tuned PIDND2N2 controller showcases a higher phase margin, greater gain margin, and a wider bandwidth compared to the AOA-tuned PIDND2N2 controller. These results signify that the b-AOA-based controller maintains better stability and frequency response characteristics, making it superior in terms of overall system stability.

8.4. Compared Algorithms and Respective Transfer Functions

In this section, we provide a comparative analysis of well-known methods in the literature, which employ different types of controllers. The controller types used in these approaches are as follows: sine cosine algorithm (SCA)-based PID controller [29], whale optimization algorithm (WOA)-based PIDA controller [30], slime mould algorithm (SMA)-based FOPID controller [31], and particle swarm optimization (PSO)-based PIDD2 controller [32].
The parameters for the SCA-based PID controller [29] are as follows: K p = 0.9826 , K i = 0.8337 , and K d = 0.4982 . The transfer function of the closed-loop AVR system using this approach is given by the following equation.
T S C A P I D s = 0.04982 s 3 + 5.08 s 2 + 9.909 s + 8.337 0.0004 s 5 + 0.0454 s 4 + 0.555 s 3 + 6.492 s 2 + 10.83 s + 8.337
The parameters for the WOA-based PIDA controller [30] are as follows: K p = 777.401 , K i = 397.741 , K d = 500.652 , K a = 103.02 , α = 550.118 , and β = 915.041 . The transfer function of the closed-loop AVR system using this approach is given by the following equation.
T W O A P I D A s = 10.3 s 4 + 1080 s 3 + 5084 s 2 + 7814 s + 3977 0.0004 s 7 + 0.2654 s 6 + 25.9 s 5 + 348.4 s 4 + 2370 s 3 + 6938 s 2 + 8689 s + 3977
The parameters for the SMA-based FOPID controller [31] are as follows: K p = 2.2554 , K i = 1.2586 , K d = 0.6472 , λ = 1.0274 , and μ = 1.1877 . The transfer function of the closed-loop AVR system using this approach is given by the following equation.
T S M A F O P I D s = 0.06472 s 3.2151 + 6.472 s 2.2151 + 0.22554 s 2.0274 + 22.554 s 1.0274 + 0.12586 s + 12.586 0.0004 s 5.0274 + 0.0454 s 4.0274 + 0.555 s 3.0274 + 6.472 s 2.2151 + 1.51 s 2.0274 + 23.554 s 1.0274 + 12.586
The parameters for the PSO-based PIDD2 controller [32] are as follows: K p = 2.7784 , K i = 1.8521 , K d 1 = 0.9997 , and K d 2 = 0.07394 . The transfer function of the closed-loop AVR system using this approach is given by the following equation.
T P S O P I D D 2 s = 0.007394 s 4 + 0.8394 s 3 + 10.27 s 2 + 27.97 s + 18.52 0.0004 s 5 + 0.0454 s 4 + 1.294 s 3 + 11.51 s 2 + 28.78 s + 18.52
These equations define the transfer functions of the AVR systems under the influence of different control methods. The following subsections provide a comparative analysis of these methods based on various performance criteria.

8.5. Comparative Transient Response Analysis

Figure 15 displays the comparative step response of different control approaches for the AVR system. This figure visually represents the transient response of various control methods and provides insights into their performance. The step response graph shows how each method reacts to a change in the terminal voltage.
Table 13 complements the visual representation by providing numerical values for the transient response metrics of different control approaches. These metrics include the rise time, settling time, and overshoot, which are essential indicators of the system’s dynamic behavior.
Upon analyzing both the figure and the table, it becomes evident that the b-AOA-tuned PIDND2N2 controller excels in achieving a superior transient response compared to other control approaches. It exhibits the shortest rise time (0.033485 s) and settling time (0.050752 s) while completely eliminating overshoot. In contrast, the other control methods, including the AOA-tuned PIDND2N2, SCA-tuned PID, WOA-tuned PIDA, SMA-tuned FOPID, and PSO-tuned PIDD2, exhibit longer rise and settling times and, in some cases, significant overshoot. These results emphasize the superiority of the b-AOA-based control approach in providing a faster and more stable transient response, which is crucial for maintaining the AVR system’s stability and performance during dynamic voltage changes.

8.6. Comparative Frequency Response Analysis

Figure 16 provides a comparative view of the Bode diagrams for different control approaches applied to the AVR system. These diagrams illustrate the frequency response characteristics of each control method, offering insights into how they perform across a range of frequencies.
Table 14 complements the visual representation with numerical values that quantify the frequency response metrics for each control approach. These metrics include the phase margin, gain margin, and bandwidth, which are crucial indicators of the system’s stability and ability to handle varying frequencies.
Upon analyzing both the figure and the table, it is clear that the b-AOA-tuned PIDND2N2 controller stands out as the superior choice for frequency response analysis. It exhibits the highest phase margin (70.797°), indicating robust stability and the most favorable gain margin (28.888 dB) among all the methods, ensuring ample room for gain adjustments without instability. Moreover, it possesses the widest bandwidth (64.82 rad/s), signifying a faster system response to frequency variations. In contrast, the other control approaches, including the AOA-tuned PIDND2N2, SCA-tuned PID, WOA-tuned PIDA, SMA-tuned FOPID, and PSO-tuned PIDD2, generally display lower phase margins, lower gain margins, and narrower bandwidths. The b-AOA-based controller, on the other hand, excels in maintaining system stability across a broad frequency range and offers improved performance for handling dynamic frequency changes. These results underscore the superiority of the b-AOA-tuned PIDND2N2 controller in providing robust and responsive frequency characteristics, which are vital for the stable and efficient operation of the AVR system under various operating conditions.

8.7. Comparisons with the Reported Recent Works

In this section, we compare the proposed PIDND2N2 controller tuned with the b-AOA to several recently reported control methods for the AVR system. These methods include a variety of controllers, each tuned using different optimization algorithms such as marine predators algorithm (MPA)-based FOPID [33], hybrid atom search particle swarm optimization (h-ASPSO)-based PID [34], equilibrium optimizer (EO)-based TIλDND2N2-based controller [35], reptile search algorithm (RSA)-based FOPIDD2 [11], improved Runge–Kutta (iRUN) algorithm-based PIDND2N2 [36], symbiotic organism search (SOS) algorithm-based PID-F [37], whale optimization algorithm (WOA)-based 2DOF FOPI [38], Lévy flight-based RSA with local search ability (L-RSANM)-based PID [39], chaotic black widow algorithm (ChBWO)-based FOPID [20], genetic algorithm (GA)-based fuzzy PID [40], sine cosine algorithm (SCA)-based FOPID with fractional-order filter [41], hybrid simulated annealing–Manta ray foraging optimization (SA-MRFO) algorithm-based PIDD2 [42], slime mould algorithm (SMA)-based PID [43], gradient-based optimization (GBO)-based FOPID [44], and nonlinear SCA-based sigmoid PID [45].
We evaluate their transient response performance to assess the effectiveness of the proposed approach. Table 15 provides a comprehensive overview of the transient response metrics, including rise time, settling time, and overshoot, for the proposed approach and other recent methods. The results demonstrate the efficacy of the b-AOA-based PIDND2N2 controller in comparison to various state-of-the-art methods as it stands out with an impressive performance, featuring a remarkably low rise time (0.033485 s), a fast settling time (0.050752 s), and zero overshoot. This suggests the exceptional stability and responsiveness of the b-AOA-tuned controller. Therefore, the table clearly illustrates the effectiveness of the proposed b-AOA-based PIDND2N2 controller in achieving rapid responses and maintaining stable performance, as evidenced by its minimal overshoot. It consistently outperforms or rivals the other methods in the evaluation, reinforcing its superiority for the AVR system’s transient response.

9. Conclusions and Future Works

In this study, we have introduced a novel approach to enhance the control of the AVR in power systems. By uniting a PIDND2N2 controller with the novel b-AOA, we aimed to address the limitations associated with conventional methods. The introduction of the PIDND2N2 controller offers enhanced precision, stability, and responsiveness in voltage regulation. This innovative configuration mitigates the shortcomings of existing approaches, promising a superior control performance. The b-AOA optimizer is meticulously fine-tuned with the integration of PS and EOBL strategies into the original AOA in order to demonstrate an exceptional performance. The assessment on 23 benchmark functions shows that it consistently achieves accurate solutions, exhibits robustness in addressing various optimization problems, and showcases remarkable potential for a wide range of applications. Extensive comparative analyses reveal the superiority of the proposed approach in transient response characteristics. The b-AOA-based AVR control approach excels in rise time, settling time, and overshoot, outperforming other methods. It also ensures robust stability with favorable gain margins and a broader bandwidth, offering improved performance for handling dynamic frequency changes. The results of our work set a new benchmark for AVR control, advancing stability, responsiveness, and reliability in power systems.
Future research in this domain should focus on several key aspects. Firstly, further refinement of the b-AOA optimization framework, exploration of additional optimization problems, and evaluation of its applicability to diverse domains are promising directions. Inspired by recent developments in integrated energy systems [54], our subsequent work will explore the adaptation of our optimization approach to various energy systems, aiming to showcase its advantages and contribute to the broader field. Secondly, investigating the practical implementation of our proposed control scheme in real-world power systems and conducting extensive field testing will provide valuable insights into its real-world performance. Additionally, the integration of emerging technologies, such as machine learning and artificial intelligence, into AVR control systems holds potential for further enhancement. Lastly, addressing scalability and assessing the applicability of our approach in more complex power systems will be crucial for its broader adoption. The pursuit of more efficient, stable, and responsive AVR systems remains a vibrant field of research, and we anticipate potential breakthroughs on the horizon.

Author Contributions

Methodology, S.E., H.Ç. and D.I.; software, S.E. and H.Ç.; validation, D.I. and E.K.; investigation, S.E., H.Ç., D.I. and E.K.; writing—original draft, D.I. and H.Ç.; writing—review & editing, E.K.; project administration, S.E. and D.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All produced data are available within the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kushwah, B.; Batool, S.; Gill, A.; Singh, M. ANN and ANFIS Techniques for Automatic Voltage Regulation. In Proceedings of the 2023 4th International Conference for Emerging Technology (INCET), Belgaum, India, 26–28 May 2023; IEEE: New York, NY, USA, 2023; pp. 1–8. [Google Scholar]
  2. Saat, S.; Ghazali, M.R.; Ahmad, M.A.; Mustapha, N.M.Z.A.; Tumari, M.Z.M. An Implementation of Brain Emotional Learning Based Intelligent Controller for AVR System. In Proceedings of the 2023 IEEE International Conference on Automatic Control and Intelligent Systems (I2CACIS), Shah Alam, Malaysia, 17 June 2023; IEEE: New York, NY, USA, 2023; pp. 60–64. [Google Scholar]
  3. Bhookya, J.; Jatoth, R.K. Optimal FOPID/PID Controller Parameters Tuning for the AVR System Based on Sine–Cosine-Algorithm. Evol. Intell. 2019, 12, 725–733. [Google Scholar] [CrossRef]
  4. Micev, M.; Ćalasan, M.; Oliva, D. Fractional Order PID Controller Design for an AVR System Using Chaotic Yellow Saddle Goatfish Algorithm. Mathematics 2020, 8, 1182. [Google Scholar] [CrossRef]
  5. Noman, A.M.; Almutairi, S.Z.; Aly, M.; Alqahtani, M.H.; Aljumah, A.S.; Mohamed, E.A. A Marine-Predator-Algorithm-Based Optimum FOPID Controller for Enhancing the Stability and Transient Response of Automatic Voltage Regulators. Fractal Fract. 2023, 7, 690. [Google Scholar] [CrossRef]
  6. Ohanu, C.P.; Odo, K.C.; Omeje, L.U.; Sutikno, T. Output Performance Evaluation of the Automatic Voltage Regulator System on Pre-Filter Control Technique. Int. J. Power Electron. Drive Syst. 2023, 14, 789. [Google Scholar] [CrossRef]
  7. Feng, W.; Xuebing, C.; Shuyu, C.; Beng, S.C.; Zhiqiang, C.; Jet, T.K.; Vilathgamuwa, D.M. MPC Based Dynamic Voltage Regulation Using Grid-Side BESPS with the Consideration of Communication Delay. IEEE Trans. Energy Convers. 2023, 38, 838–848. [Google Scholar] [CrossRef]
  8. Micev, M.; Ćalasan, M.; Radulović, M. Optimal Tuning of the Novel Voltage Regulation Controller Considering the Real Model of the Automatic Voltage Regulation System. Heliyon 2023, 9, e18707. [Google Scholar] [CrossRef]
  9. Abualigah, L.; Diabat, A.; Mirjalili, S.; Abd Elaziz, M.; Gandomi, A.H. The Arithmetic Optimization Algorithm. Comput. Methods Appl. Mech. Eng. 2021, 376, 113609. [Google Scholar] [CrossRef]
  10. Omar, O.A.M.; Marei, M.I.; Attia, M.A. Comparative Study of AVR Control Systems Considering a Novel Optimized PID-Based Model Reference Fractional Adaptive Controller. Energies 2023, 16, 830. [Google Scholar] [CrossRef]
  11. Can, Ö.; Andiç, C.; Ekinci, S.; Izci, D. Enhancing Transient Response Performance of Automatic Voltage Regulator System by Using a Novel Control Design Strategy. Electr. Eng. 2023, 105, 1993–2005. [Google Scholar] [CrossRef]
  12. Mok, R.; Ahmad, M.A. Fast and Optimal Tuning of Fractional Order PID Controller for AVR System Based on Memorizable-Smoothed Functional Algorithm. Eng. Sci. Technol. Int. J. 2022, 35, 101264. [Google Scholar] [CrossRef]
  13. Ćalasan, M.; Micev, M.; Djurovic, Ž.; Mageed, H.M.A. Artificial Ecosystem-Based Optimization for Optimal Tuning of Robust PID Controllers in AVR Systems with Limited Value of Excitation Voltage. Int. J. Electr. Eng. Educ. 2020, 60 (Suppl. 1), 1857–1884. [Google Scholar] [CrossRef]
  14. Ekinci, S.; Can, Ö.; Izci, D. Controller Design for Automatic Voltage Regulator System Using Modified Opposition-Based Weighted Mean of Vectors Algorithm. Int. J. Model. Simul. 2023, 1–18. [Google Scholar] [CrossRef]
  15. Izci, D.; Rizk-Allah, R.M.; Snášel, V.; Ekinci, S.; Hashim, F.A.; Abualigah, L. A Novel Control Scheme for Automatic Voltage Regulator Using Novel Modified Artificial Rabbits Optimizer. E-Prime Adv. Electr. Eng. Electron. Energy 2023, 6, 100325. [Google Scholar] [CrossRef]
  16. Mohamadwasel, N.B. Rider Optimization Algorithm Implemented on the AVR Control System Using MATLAB with FOPID. IOP Conf. Ser. Mater. Sci. Eng. 2020, 928, 032017. [Google Scholar] [CrossRef]
  17. Foroutan, F.; Tootoonchian, F. Experimental Evaluation of Optimal Tuning for PID Parameters in an AVR System. Sci. Iran. 2021, 29, 3317–3325. [Google Scholar] [CrossRef]
  18. Mohd Tumari, M.Z.; Ahmad, M.A.; Suid, M.H.; Hao, M.R. An Improved Marine Predators Algorithm-Tuned Fractional-Order PID Controller for Automatic Voltage Regulator System. Fractal Fract. 2023, 7, 561. [Google Scholar] [CrossRef]
  19. Çelik, E.; Durgut, R. Performance Enhancement of Automatic Voltage Regulator by Modified Cost Function and Symbiotic Organisms Search Algorithm. Eng. Sci. Technol. Int. J. 2018, 21, 1104–1111. [Google Scholar] [CrossRef]
  20. Munagala, V.K.; Jatoth, R.K. Improved Fractional PIλDμ Controller for AVR System Using Chaotic Black Widow Algorithm. Comput. Electr. Eng. 2022, 97, 107600. [Google Scholar] [CrossRef]
  21. Alghamdi, S.; Sindi, H.F.; Rawa, M.; Alhussainy, A.A.; Calasan, M.; Micev, M.; Ali, Z.M.; Abdel Aleem, S.H.E. Optimal PID Controllers for AVR Systems Using Hybrid Simulated Annealing and Gorilla Troops Optimization. Fractal Fract. 2022, 6, 682. [Google Scholar] [CrossRef]
  22. Dakua, B.K.; Sahoo, B.; Pati, B.B. Design of PIλDµ Controller for a Fractional-Order Automatic Voltage Regulator System. IFAC-Pap. 2022, 55, 649–654. [Google Scholar] [CrossRef]
  23. Koessler, E.; Almomani, A. Hybrid Particle Swarm Optimization and Pattern Search Algorithm. Optim. Eng. 2021, 22, 1539–1555. [Google Scholar] [CrossRef]
  24. Khanduja, N.; Bhushan, B. Chaotic State of Matter Search with Elite Opposition Based Learning: A New Hybrid Metaheuristic Algorithm. Optim. Control Appl. Methods 2021, 44, 533–548. [Google Scholar] [CrossRef]
  25. Mirjalili, S. SCA: A Sine Cosine Algorithm for Solving Optimization Problems. Knowl. Based Syst. 2016, 96, 120–133. [Google Scholar] [CrossRef]
  26. Ahmadianfar, I.; Heidari, A.A.; Noshadian, S.; Chen, H.; Gandomi, A.H. INFO: An Efficient Optimization Algorithm Based on Weighted Mean of Vectors. Expert Syst. Appl. 2022, 195, 116516. [Google Scholar] [CrossRef]
  27. Faramarzi, A.; Heidarinejad, M.; Mirjalili, S.; Gandomi, A.H. Marine Predators Algorithm: A Nature-Inspired Metaheuristic. Expert Syst. Appl. 2020, 152, 113377. [Google Scholar] [CrossRef]
  28. Mohapatra, S.; Choudhury, D.; Bishi, K.; Keshari, S.; Dakua, B.K.; Kaunda, C.; Panda, A. A Comparison between the FOTID and FOPID Controller for the Close-Loop Speed Control of a DC Motor System. In Proceedings of the 2023 International Conference on Artificial Intelligence and Applications (ICAIA) Alliance Technology Conference (ATCON-1), Bengaluru, India, 21–22 April 2023; pp. 1–6. [Google Scholar]
  29. Hekimoğlu, B. Sine-Cosine Algorithm-Based Optimization for Automatic Voltage Regulator System. Trans. Inst. Meas. Control. 2019, 41, 1761–1771. [Google Scholar] [CrossRef]
  30. Mosaad, A.M.; Attia, M.A.; Abdelaziz, A.Y. Whale Optimization Algorithm to Tune PID and PIDA Controllers on AVR System. Ain Shams Eng. J. 2019, 10, 755–767. [Google Scholar] [CrossRef]
  31. Izci, D.; Ekinci, S.; Zeynelgil, H.L.; Hedley, J. Fractional Order PID Design Based on Novel Improved Slime Mould Algorithm. Electr. Power Compon. Syst. 2021, 49, 901–918. [Google Scholar] [CrossRef]
  32. Sahib, M.A. A Novel Optimal PID plus Second Order Derivative Controller for AVR System. Eng. Sci. Technol. Int. J. 2015, 18, 194–206. [Google Scholar] [CrossRef]
  33. Mohd Tumari, M.Z.; Ahmad, M.A.; Mohd Rashid, M.I. A Fractional Order PID Tuning Tool for Automatic Voltage Regulator Using Marine Predators Algorithm. Energy Rep. 2023, 9, 416–421. [Google Scholar] [CrossRef]
  34. Izci, D.; Ekinci, S.; Hussien, A.G. Effective PID Controller Design Using a Novel Hybrid Algorithm for High Order Systems. PLoS ONE 2023, 18, e0286060. [Google Scholar] [CrossRef]
  35. Tabak, A. Novel TIλDND2N2 Controller Application with Equilibrium Optimizer for Automatic Voltage Regulator. Sustainability 2023, 15, 11640. [Google Scholar] [CrossRef]
  36. Izci, D.; Ekinci, S. An Improved RUN Optimizer Based Real PID plus Second-Order Derivative Controller Design as a Novel Method to Enhance Transient Response and Robustness of an Automatic Voltage Regulator. E-Prime—Adv. Electr. Eng. Electron. Energy 2022, 2, 100071. [Google Scholar] [CrossRef]
  37. Ozgenc, B.; Ayas, M.S.; Altas, I.H. Performance Improvement of an AVR System by Symbiotic Organism Search Algorithm-Based PID-F Controller. Neural Comput. Appl. 2022, 34, 7899–7908. [Google Scholar] [CrossRef]
  38. Padiachy, V.; Mehta, U.; Azid, S.; Prasad, S.; Kumar, R. Two Degree of Freedom Fractional PI Scheme for Automatic Voltage Regulation. Eng. Sci. Technol. Int. J. 2022, 30, 101046. [Google Scholar] [CrossRef]
  39. Ekinci, S.; Izci, D.; Abu Zitar, R.; Alsoud, A.R.; Abualigah, L. Development of Lévy Flight-Based Reptile Search Algorithm with Local Search Ability for Power Systems Engineering Design Problems. Neural Comput. Appl. 2022, 34, 20263–20283. [Google Scholar] [CrossRef]
  40. Dogruer, T.; Can, M.S. Design and Robustness Analysis of Fuzzy PID Controller for Automatic Voltage Regulator System Using Genetic Algorithm. Trans. Inst. Meas. Control 2022, 44, 1862–1873. [Google Scholar] [CrossRef]
  41. Ayas, M.S.; Sahin, E. FOPID Controller with Fractional Filter for an Automatic Voltage Regulator. Comput. Electr. Eng. 2021, 90, 106895. [Google Scholar] [CrossRef]
  42. Micev, M.; Ćalasan, M.; Ali, Z.M.; Hasanien, H.M.; Abdel Aleem, S.H.E. Optimal Design of Automatic Voltage Regulation Controller Using Hybrid Simulated Annealing—Manta Ray Foraging Optimization Algorithm. Ain Shams Eng. J. 2021, 12, 641–657. [Google Scholar] [CrossRef]
  43. Izci, D.; Ekinci, S. Comparative Performance Analysis of Slime Mould Algorithm For Efficient Design of Proportional–Integral–Derivative Controller. Electrica 2021, 21, 151–159. [Google Scholar] [CrossRef]
  44. Altbawi, S.M.A.; Mokhtar, A.S.B.; Jumani, T.A.; Khan, I.; Hamadneh, N.N.; Khan, A. Optimal Design of Fractional Order PID Controller Based Automatic Voltage Regulator System Using Gradient-Based Optimization Algorithm. J. King Saud Univ.—Eng. Sci. 2021, in press. [Google Scholar] [CrossRef]
  45. Suid, M.H.; Ahmad, M.A. Optimal Tuning of Sigmoid PID Controller Using Nonlinear Sine Cosine Algorithm for the Automatic Voltage Regulator System. ISA Trans. 2022, 128, 265–286. [Google Scholar] [CrossRef] [PubMed]
  46. Lewis, R.M.; Torczon, V. Pattern Search Algorithms for Bound Constrained Minimization. SIAM J. Optim. 1999, 9, 1082–1099. [Google Scholar] [CrossRef]
  47. Tizhoosh, H.R. Opposition-Based Learning: A New Scheme for Machine Intelligence. In Proceedings of the International Conference on Computational Intelligence for Modelling, Control and Automation and International Conference on Intelligent Agents, Web Technologies and Internet Commerce (CIMCA-IAWTIC’06), Washington, DC, USA, 28–30 November 2005; Volume 1, pp. 695–701. [Google Scholar]
  48. Satria, H.; Syah, R.B.Y.; Nehdi, M.L.; Almustafa, M.K.; Adam, A.O.I. Parameters Identification of Solar PV Using Hybrid Chaotic Northern Goshawk and Pattern Search. Sustainability 2023, 15, 5027. [Google Scholar] [CrossRef]
  49. Yildiz, B.S.; Pholdee, N.; Bureerat, S.; Yildiz, A.R.; Sait, S.M. Enhanced Grasshopper Optimization Algorithm Using Elite Opposition-Based Learning for Solving Real-World Engineering Problems. Eng. Comput. 2021, 38, 4207–4219. [Google Scholar] [CrossRef]
  50. Izci, D.; Ekinci, S.; Mirjalili, S.; Abualigah, L. An Intelligent Tuning Scheme with a Master/Slave Approach for Efficient Control of the Automatic Voltage Regulator. Neural Comput. Appl. 2023, 35, 19099–19115. [Google Scholar] [CrossRef]
  51. Oladipo, S.; Sun, Y.; Wang, Z. Optimization of PID and FOPID Controllers with New Generation Metaheuristic Algorithms for Controlling AVR System: Concise Survey. In Proceedings of the 2020 12th International Conference on Computational Intelligence and Communication Networks (CICN), Erode, Tamil Nadu, India, 22–24 February 2023; pp. 280–286. [Google Scholar]
  52. Mousakazemi, S.M.H. Comparison of the Error-Integral Performance Indexes in a GA-Tuned PID Controlling System of a PWR-Type Nuclear Reactor Point-Kinetics Model. Prog. Nucl. Energy 2021, 132, 103604. [Google Scholar] [CrossRef]
  53. Shaija, P.J.; Daniel, A.E. Optimal Tuning of PI Controllers for IM Drive Using GWO and TLBO Algorithms. In Proceedings of the 2023 Fifth International Conference on Electrical, Computer and Communication Technologies (ICECCT), Erode, Tamil Nadu, India, 22–24 February 2023; pp. 1–9. [Google Scholar]
  54. Zhang, N.; Sun, Q.; Yang, L.; Li, Y. Event-Triggered Distributed Hybrid Control Scheme for the Integrated Energy System. IEEE Trans. Industr. Inform. 2022, 18, 835–846. [Google Scholar] [CrossRef]
Figure 1. Flowchart of the original arithmetic optimization algorithm.
Figure 1. Flowchart of the original arithmetic optimization algorithm.
Mathematics 11 04810 g001
Figure 2. Flowchart of pattern search method.
Figure 2. Flowchart of pattern search method.
Mathematics 11 04810 g002
Figure 3. Working principle of OBL mechanism.
Figure 3. Working principle of OBL mechanism.
Mathematics 11 04810 g003
Figure 4. Flowchart of proposed b-AOA.
Figure 4. Flowchart of proposed b-AOA.
Mathematics 11 04810 g004
Figure 5. Schematic diagram of a typical AVR system.
Figure 5. Schematic diagram of a typical AVR system.
Mathematics 11 04810 g005
Figure 6. An uncontrolled AVR system.
Figure 6. An uncontrolled AVR system.
Mathematics 11 04810 g006
Figure 7. Pole–zero map of uncontrolled AVR system.
Figure 7. Pole–zero map of uncontrolled AVR system.
Mathematics 11 04810 g007
Figure 8. Step response of the uncontrolled AVR system.
Figure 8. Step response of the uncontrolled AVR system.
Mathematics 11 04810 g008
Figure 9. Open-loop Bode plot of the uncontrolled AVR system.
Figure 9. Open-loop Bode plot of the uncontrolled AVR system.
Mathematics 11 04810 g009
Figure 10. Block diagram of PIDND2N2 controller.
Figure 10. Block diagram of PIDND2N2 controller.
Mathematics 11 04810 g010
Figure 11. The block diagram of the implementation of the proposed approach to AVR system.
Figure 11. The block diagram of the implementation of the proposed approach to AVR system.
Mathematics 11 04810 g011
Figure 12. Convergence curve for b-AOA and original AOA.
Figure 12. Convergence curve for b-AOA and original AOA.
Mathematics 11 04810 g012
Figure 13. Step response of b-AOA- and AOA-tuned PIDND2N2 controllers for the change in the terminal voltage.
Figure 13. Step response of b-AOA- and AOA-tuned PIDND2N2 controllers for the change in the terminal voltage.
Mathematics 11 04810 g013
Figure 14. Open-loop Bode diagrams for b-AOA- and AOA-tuned PIDND2N2 controllers.
Figure 14. Open-loop Bode diagrams for b-AOA- and AOA-tuned PIDND2N2 controllers.
Mathematics 11 04810 g014
Figure 15. Comparative step response of different control approaches for AVR system.
Figure 15. Comparative step response of different control approaches for AVR system.
Mathematics 11 04810 g015
Figure 16. Comparative Bode diagrams of different control approaches for AVR system.
Figure 16. Comparative Bode diagrams of different control approaches for AVR system.
Mathematics 11 04810 g016
Table 1. Properties of the adopted unimodal benchmark functions.
Table 1. Properties of the adopted unimodal benchmark functions.
Name FunctionDimensionEvaluation IntervalGlobal Minimum
Sphere F u n c 1 ( x ) 30 [ 100 ,   100 ] D i m 0
Schwefel 2.2 F u n c 2 ( x ) 30 [ 10 ,   10 ] D i m 0
Schwefel 1.2 F u n c 3 ( x ) 30 [ 100 ,   100 ] D i m 0
Schwefel 2.21 F u n c 4 ( x ) 30 [ 100 ,   100 ] D i m 0
Rosenbrock F u n c 5 ( x ) 30 [ 30 ,   30 ] D i m 0
Step F u n c 6 ( x ) 30 [ 100 ,   100 ] D i m 0
Quartic F u n c 7 ( x ) 30 [ 1.28 ,   1.28 ] D i m 0
Table 2. Properties of the adopted multimodal benchmark functions.
Table 2. Properties of the adopted multimodal benchmark functions.
Name FunctionDimensionEvaluation IntervalGlobal Minimum
Schwefel F u n c 8 ( x ) 30 [ 500 ,   500 ] D i m −1.2569 × 104
Rastrigin F u n c 9 ( x ) 30 [ 5.12 ,   5.12 ] D i m 0
Ackley F u n c 10 ( x ) 30 [ 32 ,   32 ] D i m 0
Griewank F u n c 11 ( x ) 30 [ 600 ,   600 ] D i m 0
Penalized F u n c 12 ( x ) 30 [ 50 ,   50 ] D i m 0
Penalized2 F u n c 13 ( x ) 30 [ 50 ,   50 ] D i m 0
Table 3. Properties of the adopted fixed-dimensional multimodal benchmark functions.
Table 3. Properties of the adopted fixed-dimensional multimodal benchmark functions.
NameFunctionDimensionEvaluation IntervalGlobal Minimum
Foxholes F u n c 14 ( x ) 2 [ 65.536 ,   65.536 ] D i m 0.998
Kowalik F u n c 15 ( x ) 4 [ 5 ,   5 ] D i m 3.0749 × 10−4
Six-Hump Camel F u n c 16 ( x ) 2 [ 5 ,   5 ] D i m −1.0316
Branin F u n c 17 ( x ) 2 5 ,   10 × 0 ,   15 0.39789
Goldstein–Price F u n c 18 ( x ) 2 [ 2 ,   2 ] D i m 3
Hartman 3 F u n c 19 ( x ) 3 [ 0 ,   1 ] D i m −3.8628
Hartman 6 F u n c 20 ( x ) 6 [ 0 ,   1 ] D i m −3.322
Shekel 5 F u n c 21 ( x ) 4 [ 0 ,   10 ] D i m −10.1532
Shekel 7 F u n c 22 ( x ) 4 [ 0 ,   10 ] D i m −10.4029
Shekel 10 F u n c 23 ( x ) 4 [ 0 ,   10 ] D i m −10.5364
Table 4. Properties of the compared algorithms (population size, total iteration number, values of other control parameters).
Table 4. Properties of the compared algorithms (population size, total iteration number, values of other control parameters).
Algorithm Population SizeTotal Iteration NumberValues of Other Control Parameters
b-AOA 30500 α = 5 , μ = 0.4975 , M i n = 0.2 , M a x = 1 , i n i t i a l   m e s h   s i z e = 1 ,   m e s h   e x p a n s i o n   f a c t o r = 2 ,   m e s h   c o n t r a c t i o n   f a c t o r = 0.5 ,   a l l   t o l e r a n c e s = 10 6
AOA [9]30500 α = 5 , μ = 0.4975 , M i n = 0.2 , M a x = 1
SCA [25]30500 A = 2
INFO [26]30500 c = 2 , d = 4
MPA [27]30500 F A D s = 0.2 , P = 0.5
Table 5. Comparative statistical results obtained from unimodal benchmark functions.
Table 5. Comparative statistical results obtained from unimodal benchmark functions.
FunctionAlgorithmMeanStandard DeviationBestWorst
F u n c 1 ( x ) b-AOA0000
AOA0.000296560.00114133.9226 × 10−380.0060134
SCA16.53736.4269.5633 × 10−6175.47
INFO1.0185 × 10−534.997 × 10−543.3545 × 10−552.0178 × 10−53
MPA4.0116 × 10−236.3963 × 10−233.6461 × 10−252.7727 × 10−22
F u n c 2 ( x ) b-AOA8.5996 × 10−24104.333 × 10−3202.2954 × 10−239
AOA2.8674 × 10−18609.6235 × 10−2968.6022 × 10−185
SCA0.0212410.0315670.000137670.13042
INFO1.0943 × 10−263.6605 × 10−274.7283 × 10−271.9892 × 10−26
MPA2.6444 × 10−132.8514 × 10−138.2406 × 10−151.2622 × 10−12
F u n c 3 ( x ) b-AOA0000
AOA1.60113.38161.3815 × 10−716.177
SCA8640.84939.51709.520103
INFO1.4606 × 10−501.1602 × 10−508.6654 × 10−523.9712 × 10−50
MPA9.9612 × 10−50.000223467.2658 × 10−90.001186
F u n c 4 ( x ) b-AOA9.0422 × 10−24401.2808 × 10−2532.6479 × 10−242
AOA0.154160.0948770.0146320.36318
SCA37.03313.08712.16661.964
INFO2.1028 × 10−271.4215 × 10−273.5852 × 10−287.4954 × 10−27
MPA2.7542 × 10−91.5152 × 10−93.1553 × 10−106.0257 × 10−9
F u n c 5 ( x ) b-AOA0.616151.88143.0737 × 10−96.3967
AOA28.6930.2754927.90229.18
SCA1.3673 × 1053.2682 × 105107.541175700
INFO22.5850.5171121.29823.462
MPA25.2680.4545124.48726.042
F u n c 6 ( x ) b-AOA2.4395 × 10−129.2009 × 10−131.086 × 10−125.8521 × 10−12
AOA3.75240.333313.05614.4582
SCA14.25413.5424.719155.025
INFO1.2654 × 10−83.7987 × 10−83.9266 × 10−112.07 × 10−7
MPA4.1868 × 10−82.2575 × 10−81.3296 × 10−81.2965 × 10−7
F u n c 7 ( x ) b-AOA3.629 × 10−52.8489 × 10−56.8524 × 10−70.00010771
AOA9.4896 × 10−57.1313 × 10−52.0672 × 10−60.00029718
SCA0.0991580.0905090.00858470.44986
INFO0.00159370.00126340.000172270.0049221
MPA0.00134950.000603520.000419660.0026601
Table 6. Comparative statistical results obtained from multimodal benchmark functions.
Table 6. Comparative statistical results obtained from multimodal benchmark functions.
FunctionAlgorithmMeanStandard DeviationBestWorst
F u n c 8 ( x ) b-AOA−12,536172.87−12,569−11,623
AOA−7980.7446.84−9196.5−7230.3
SCA−3848.4286.86−4371−3283.7
INFO−8630.7700.38−9763.3−7101.2
MPA−8736.9438.15−9687.9−7946.9
F u n c 9 ( x ) b-AOA0000
AOA0000
SCA29.30830.1890.13996122.46
INFO0000
MPA0000
F u n c 10 ( x ) b-AOA8.8818 × 10−1608.8818 × 10−168.8818 × 10−16
AOA8.8818 × 10−1608.8818 × 10−168.8818 × 10−16
SCA14.2088.32120.04340120.382
INFO8.8818 × 10−1608.8818 × 10−168.8818 × 10−16
MPA1.7196 × 10−121.1519 × 10−122.7045 × 10−135.8482 × 10−12
F u n c 11 ( x ) b-AOA0000
AOA194.1265.89672.408323.52
SCA0.845690.411640.235451.9083
INFO0000
MPA0000
F u n c 12 ( x ) b-AOA2.1943 × 10−131.5539 × 10−135.0331 × 10−146.0379 × 10−13
AOA0.291540.0538090.145380.43947
SCA524281.5261E+051.0947614,430
INFO1.4456 × 10−92.8117 × 10−95.3463 × 10−121.1459 × 10−8
MPA0.000142860.00050592.4157 × 10−90.0023059
F u n c 13 ( x ) b-AOA3.1668 × 10−122.4141 × 10−127.6907 × 10−139.0849 × 10−12
AOA2.44840.169152.12172.8078
SCA1.0872 × 1052.7869 × 1052.20421,305,400
INFO0.0637520.142733.2034 × 10−100.69157
MPA0.0122150.0368762.8969 × 10−80.19763
Table 7. Comparative statistical results obtained from fixed-dimensional multimodal benchmark functions.
Table 7. Comparative statistical results obtained from fixed-dimensional multimodal benchmark functions.
FunctionAlgorithmMeanStandard DeviationBestWorst
F u n c 14 ( x ) b-AOA0.9981.5701 × 10−170.9980.998
AOA8.36963.23890.99812.671
SCA1.7950.98590.9982.9821
INFO2.11112.59030.99810.763
MPA0.9981.515 × 10−160.9980.998
F u n c 15 ( x ) b-AOA0.000307491.4923E−150.000307490.00030749
AOA0.0154170.0256040.000371890.11249
SCA0.00106610.000370020.00058290.0015477
INFO0.00243520.00608630.000307490.020363
MPA0.000307494.3122 × 10−150.000307490.00030749
F u n c 16 ( x ) b-AOA−1.03161.9902 × 10−16−1.0316−1.0316
AOA−1.03166.0816 × 10−7−1.0316−1.0316
SCA−1.03163.7905 × 10−5−1.0316−1.0315
INFO−1.03166.5843 × 10−16−1.0316−1.0316
MPA−1.03164.4024 × 10−16−1.0316−1.0316
F u n c 17 ( x ) b-AOA0.3978900.397890.39789
AOA0.409870.0098640.398440.43767
SCA0.400260.00235430.397970.40949
INFO0.3978900.397890.39789
MPA0.397899.5078 × 10−150.397890.39789
F u n c 18 ( x ) b-AOA3033
AOA6.69.3351330
SCA35.4359 × 10−533.0002
INFO38.6883 × 10−1633
MPA32.1709 × 10−1533
F u n c 19 ( x ) b-AOA−3.86282.4116 × 10−15−3.8628−3.8628
AOA−3.85230.0038518−3.8593−3.842
SCA−3.85470.0024361−3.861−3.8495
INFO−3.86282.6823 × 10−15−3.8628−3.8628
MPA−3.86282.4945 × 10−15−3.8628−3.8628
F u n c 20 ( x ) b-AOA−3.3222.1608 × 10−13−3.322−3.322
AOA−3.04710.091025−3.1762−2.8234
SCA−2.87840.34163−3.1199−1.6747
INFO−3.27840.058273−3.322−3.2031
MPA−3.3221.7554 × 10−11−3.322−3.322
F u n c 21 ( x ) b-AOA−10.1537.6605 × 10−13−10.153−10.153
AOA−3.50231.1997−6.0307−1.8035
SCA−2.62022.0715−7.8686−0.49728
INFO−9.10392.4723−10.153−2.6305
MPA−10.1534.1471 × 10−11−10.153−10.153
F u n c 22 ( x ) b-AOA−10.4031.1144 × 10−12−10.403−10.403
AOA−3.56191.2118−6.8762−1.4002
SCA−3.20231.8303−5.9956−0.52105
INFO−9.04882.7774−10.403−2.7659
MPA−10.4035.9857 × 10−11−10.403−10.403
F u n c 23 ( x ) b-AOA−10.5363.2315 × 10−12−10.536−10.536
AOA−3.87331.6156−6.5892−1.5825
SCA−3.74211.7935−6.1434−0.94135
INFO−9.00393.151−10.536−2.4217
MPA−10.5362.5368 × 10−11−10.536−10.536
Table 8. Boundaries for PIDND2N2 controller parameters.
Table 8. Boundaries for PIDND2N2 controller parameters.
Bound K p K i K d 1 K d 2 n 1 n 2
Lower0.0010.0010.0010.0015050
Upper555520002000
Table 9. Statistical performance of b-AOA and original AOA for AVR system.
Table 9. Statistical performance of b-AOA and original AOA for AVR system.
Algorithm MeanStandard DeviationBestWorst
b-AOA0.00651389.3497 × 10−50.00635220.0067022
AOA0.00788630.000123950.00768250.0081212
Table 10. Optimal parameters of PIDND2N2 controller obtained via b-AOA and original AOA.
Table 10. Optimal parameters of PIDND2N2 controller obtained via b-AOA and original AOA.
Optimized by K p K i K d 1 K d 2 n 1 n 2
b-AOA4.87232.02401.80940.150491595.21971.2
AOA3.94482.11881.67570.130141544.2871.72
Table 11. Transient response performance metrics for b-AOA- and AOA-tuned PIDND2N2 controllers.
Table 11. Transient response performance metrics for b-AOA- and AOA-tuned PIDND2N2 controllers.
Design Method Rise Time (s)Settling Time (s)Overshoot (%)
b-AOA-tuned PIDND2N2 0.0334850.0507520
AOA-tuned PIDND2N2 0.0373930.0575230.043859
Table 12. Frequency response performance metrics for b-AOA- and AOA-tuned PIDND2N2 controllers.
Table 12. Frequency response performance metrics for b-AOA- and AOA-tuned PIDND2N2 controllers.
Design MethodPhase Margin (°)Gain Margin (dB)Bandwidth (rad/s)
b-AOA-tuned PIDND2N2 70.79728.88864.820
AOA-tuned PIDND2N2 69.81023.36857.819
Table 13. Comparative numerical values for transient response of different control approaches.
Table 13. Comparative numerical values for transient response of different control approaches.
Design Method Rise Time (s)Settling Time (s)Overshoot (%)
b-AOA-tuned PIDND2N2 0.0334850.0507520
AOA-tuned PIDND2N2 0.0373930.0575230.043859
SCA-tuned PID [29]0.14720.8413311.425
WOA-tuned PIDA [30]0.327720.495431.6483
SMA-tuned FOPID [31]0.0875410.497915.998
PSO-tuned PIDD2 [32]0.0929350.163470.0025797
Table 14. Comparative numerical values for frequency response of different control approaches.
Table 14. Comparative numerical values for frequency response of different control approaches.
Design MethodPhase Margin (°)Gain Margin (dB)Bandwidth (rad/s)
b-AOA-tuned PIDND2N270.79728.88864.820
AOA-tuned PIDND2N269.81023.36857.819
SCA-tuned PID [29]52.59620.30014.821
WOA-tuned PIDA [30]67.67126.1236.7076
SMA-tuned FOPID [31]49.14220.19323.914
PSO-tuned PIDD2 [32]79.638Infinite23.503
Table 15. Transient response performance of the proposed approach with respect to recently reported other efficient methods.
Table 15. Transient response performance of the proposed approach with respect to recently reported other efficient methods.
Ref.YearUsed Controller TypeTuning MethodRise Time (s)Settling Time (s)Overshoot (%)
ProposedPIDND2N2b-AOA0.0334850.0507520
[33]2023FOPIDMPA0.08330.11060.55
[34]PIDh-ASPSO0.30970.46791.2476
[35]TIλDND2N2EO0.037520.05960.4128
[11]FOPIDD2RSA0.04870.08060
[36]2022PIDND2N2iRUN0.03990.06260
[37]PID-FSOS0.2670.3710.007
[38]2DOF fractional-order PIWOA1.121.741.17
[39]PIDL-RSANM0.30760.46690.9582
[20]FOPIDChBWO0.11030.1691.1838
[40]Fuzzy PIDGA0.18570.29631.0407
[41]2021FOPID with fractional filterSCA0.12300.16700.1262
[42]PIDD2SA-MRFO0.05350.07980.7562
[43]PIDSMA0.31490.48170.6071
[44]FOPIDGBO0.08850.65311.3
[45]Sigmoid PIDNSCA0.4980.5792.2
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

Ekinci, S.; Çetin, H.; Izci, D.; Köse, E. A Novel Balanced Arithmetic Optimization Algorithm-Optimized Controller for Enhanced Voltage Regulation. Mathematics 2023, 11, 4810. https://doi.org/10.3390/math11234810

AMA Style

Ekinci S, Çetin H, Izci D, Köse E. A Novel Balanced Arithmetic Optimization Algorithm-Optimized Controller for Enhanced Voltage Regulation. Mathematics. 2023; 11(23):4810. https://doi.org/10.3390/math11234810

Chicago/Turabian Style

Ekinci, Serdar, Haluk Çetin, Davut Izci, and Ercan Köse. 2023. "A Novel Balanced Arithmetic Optimization Algorithm-Optimized Controller for Enhanced Voltage Regulation" Mathematics 11, no. 23: 4810. https://doi.org/10.3390/math11234810

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