Next Article in Journal
Applied Machine Learning to Study the Movement of Air Masses in the Wind Farm Area
Previous Article in Journal
Dynamic Radiant Barrier for Modulating Heat Transfer and Reducing Building Energy Usage
Previous Article in Special Issue
A Grid-Wide Comprehensive Evaluation Method of Power Quality Based on Complex Network Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Load Frequency Control of Interconnected Power System Using Hybrid PSO-AHA Optimizer

1
School of Electrical Engineering, Yanshan University, Qinhuangdao 066004, China
2
Key Laboratory of Power Electronics for Energy Conservation and Drive Control of Hebei Province, Yanshan University, Qinhuangdao 066004, China
3
School of Mechanical Engineering, Yanshan University, Qinhuangdao 066004, China
4
Department of Mechatronics, Faculty of Engineering, Ain Shams University, Cairo 11566, Egypt
5
Department of Industrial Engineering, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
*
Author to whom correspondence should be addressed.
Energies 2024, 17(16), 3962; https://doi.org/10.3390/en17163962 (registering DOI)
Submission received: 28 June 2024 / Revised: 19 July 2024 / Accepted: 28 July 2024 / Published: 9 August 2024
(This article belongs to the Special Issue Power Quality and Disturbances in Modern Distribution Networks)

Abstract

:
The integration of nonconventional energy sources such as solar, wind, and fuel cells into electrical power networks introduces significant challenges in maintaining frequency stability and consistent tie-line power flows. These fluctuations can adversely affect the quality and reliability of power supplied to consumers. This paper addresses this issue by proposing a Proportional–Integral–Derivative (PID) controller optimized through a hybrid Particle Swarm Optimization–Artificial Hummingbird Algorithm (PSO-AHA) approach. The PID controller is tuned using the Integral Time Absolute Error (ITAE) as a fitness function to enhance control performance. The PSO-AHA-PID controller’s effectiveness is evaluated in two networks: a two-area thermal tie-line interconnected power system (IPS) and a one-area multi-source power network incorporating thermal, solar, wind, and fuel cell sources. Comparative analyses under various operational conditions, including parameter variations and load changes, demonstrate the superior performance of the PSO-AHA-PID controller over the conventional PSO-PID controller. Statistical results indicate that in the one-area multi-source network, the PSO-AHA-PID controller achieves a 76.6% reduction in overshoot, an 88.9% reduction in undershoot, and a 97.5% reduction in settling time compared to the PSO-PID controller. In the dual-area system, the PSO-AHA-PID controller reduces the overshoot by 75.2%, reduces the undershoot by 85.7%, and improves the fall time by 71.6%. These improvements provide a robust and reliable solution for enhancing the stability of interconnected power systems in the presence of diverse and variable energy sources.

1. Introduction

Current power systems are undergoing fast changes due to the extensive application of smart grid technologies as well as a range of nonconventional sources. Loads, distribution, transmission, and generation are components of an integrated power network. A tie line links several widely distributed service areas geographically in a large, linked power network. Modern power systems are projected to face more obstacles and difficulties as they keep growing in size in addition to complicacy [1,2]. To maintain the stability of system efficiency, power generation changes proportionately with consumer load. The constancy of any power unit in a network, or any connected systems, is impacted by sudden increases in power demand [3,4].
To maintain a planned power exchange between multiple areas, energy production in each area must be regulated. LFC methods in an interlinked power system (IPS) generate and also provide electrical power constantly as well as accurately while keeping to reasonable frequency controls. The load in a power system varies continuously. The difference between generating power and consumer demand causes the system’s frequency to shift in a highly unfavorable way, causing fluctuations and the equilibrium state to float. When this equilibrium varies, the frequency is restored to a different level. This results in severe electrical network instability in addition to serious harm to costly equipment. LFC control loop’s speed governor mechanism can be used to adjust the active power needs, which will control frequency fluctuation. Control is necessary to keep the equilibrium of active capabilities. Therefore, a control system is required to address the problem of unpredictable load alterations and to maintain frequency at desired levels. The precise tuning scheme design of the controller plays a major role in providing optimal control. In order to efficiently control system frequency for single-area multiple-source as well as two-area tie-line IPS, this study proposes to optimize PID gains using optimization approaches and PSO along with PSO-AHA [5,6,7,8,9,10]. MATLAB/Simulink 2022a was used to develop suggested power networks; by varying system parameters in both networks by ±75% for various scenarios and varying the load in distinct scenarios for double-area tie-line IPS, suggested networks with PIDs were tested. PID controller gains were tuned using PSO as well as PSO-AHA approaches, with ITAE being a fitness function. In suggested networks, frequency responses, along with tie-line power variations, were successfully optimized by a PID controller utilizing PSO and PSO-AHA techniques. Signal characteristics and frequency responses show that PSO-AHA-PID is superior to PSO-PID. The sections below discuss the results of the PID controller in addition to optimization techniques that we investigated and applied to the suggested networks.

1.1. Literature Review

The author of [11] solved the interconnected thermal power networks LFC problem by using the PSO technique to optimize the PID control’s parameters. When comparing HC in addition to the GA-optimized controller’s efficiency of suggested PSO-PID, we achieved superior results. In [12,13], the authors used the PID controller and also the ACO algorithm for the LFC of a self-sustaining nuclear power plant. A method of trial and error was applied in order to contrast responses. In [14], utilizing the ARA technique, the author tuned the PID controller’s parameters for double-area non-reheat multiple-area power systems (MAPSs). In [15,16], using wind farms along with a photovoltaic model, an LFC power network was designed. In [17], the frequency was adjusted using LFC techniques, along with multi-source electricity from wind, solar, and diesel-powered generators, all of which were regarded as parts of a power network with many sources. In [18], optimal PID parameter values for the LFC renewable power network were determined by employing the nonconventional quasi-oppositional dragonfly method. Long-term significant fluctuations in tie-line power interchange along with network frequency inside LFC power networks can harm equipment, cause operating insecurity, limit the life of connected devices, or even create energy system collapse. Using a differential evolution algorithm and an optimized parameter-based controller, an analysis of LFC within a multiple-source power network was carried out [19,20].
Stochastic PSO was used to optimize the gain value of the PI controller considered for a single-area energy network LFC. It intended to employ the PSO technique via BF-tuned PID controller to solve linked power network LFC, including thermal along with PV networks [21]. Grid-connected electrical power networks’ LFC was implemented utilizing the PSO-associated MARL approach [22]. In [23], the FPA technique was applied to address LFC problems with linked power grids by optimizing PID controller parameters, and by comparing the results of the suggested FPA technique with those of GA-PID along with PSO-PID controllers, the technique’s efficiency was verified. In [24], in order to solve the single-area energy network LFC crisis, the authors investigated a PSO-tuned PID controller. Using the PSO technique in addition to four different fitness functions, the PID was considered for LFC within an isolated energy network. In [4], for LFC, a FOFPID controller was designed. In this situation, the BBO approach was used to tune controller values. However, in a steady state, the integral term performed poorly in terms of oscillation as well as frequency variations. In [25,26], the author proposed a novel optimization technique to accomplish automatic LFC. The optimization method specifically controlled the frequency variation of a multi-source power network. However, the approach also led to slow convergence, and outcomes were not ideal.
The literature review clearly demonstrates how oscillations and power system performance are caused by unpredictable load demands inside a power-generating unit. This implies that the electrical system could experience LFC/AGC issues. These issues can be solved by employing various optimization approaches in different controllers to increase gains under different circumstances as well as requirements [4,11,12,15,17,18,19,21,22,23,24,25,27,28,29,30]. A review of the literature under study is given in Table 1.
To maintain tie-line power as well as frequency at the rated value, many control strategies that make use of soft computing methodologies have been developed recently. Table 1 lists the relevant literature for the proposed work. In order to prove PSO-AHA-PID’s superiority, this paper compares the results with the PSO-PID controller in multiple cases.
PSO-AHA-PID is proposed in addition to the PSO-PID controller using the objective function ITAE proposed in this paper. The results from PSO-AHA-PID in the suggested networks are contrasted with PSO-PID in order to investigate frequency responses by considering different scenarios in the system with respect to variations in load along with system constraints in a two-area IPS. The main goal of this study is to increase the proposed system’s effectiveness and also to maintain network balance under stressful situations so that every consumer may get good quality power.

1.2. Major Contributions and Highlights

Dual-area tie-line IPS, along with a single-area multiple-source energy network consisting of PVwind, thermal, and fuel cells, are proposed systems in this article. By evaluating the effectiveness of the suggested controller in addition to optimization approaches in examined networks under system constraint alterations as well as load variations, the ITAE objective function is taken into consideration in the design and study for the suggested work. To identify the reliable supremacy of the proposed tuned controller, a thorough examination was carried out.
  • The proposed study uses a variable, sustainable power network model that accounts for fluctuation and the effects of renewable energy sources. With detailed explanations of multiple mathematical models of nonconventional energy, the LFC issue can be resolved quite rapidly.
  • This research includes a fuel cell model, fuzzy-based MPPT PV system, and PMSG-based wind energy system based upon the P & O MPPT approach in order to set up a single-area power network with multiple sources.
  • A tie-line-connected double-area power network model is created.
  • To optimize the PID controller’s parameter values, the PSO and PSO-AHA techniques are applied.
  • Upon load variations, steady-state error is reduced to zero.
  • In numerous scenarios, the effectiveness of PSO-AHA-PID for multiple-source single-area as well as double-area tie-line IPS is examined by adjusting system constraint values within the range of approximately ±75 and also by examining fluctuations in load for dual-area tie-line IPS.
  • Compared to the PSO-PID control system, it produces superior outcomes and is more dependable. The suggested control system method has a lower undershot, overshoot, and settling time than the PSO-PID controller in every scenario this study investigates.

1.3. Structure of Article

The following is the arrangement of the manuscript: Section 1 reviews the literature on the topic of the current study as well as a range of controller optimization methods. The state space functions and mathematical models that support LFC power systems are described in Section 2. A possible control strategy employing PSO-PID and PSO-AHA-PID controllers, together with data, is covered in Section 3. The achieved results are covered in Section 4. Section 5 includes recommendations for more research in addition to a summary of the study’s main results.

2. Suggested Power Framework Modeling

The recommended power structure’s mathematical modeling is presented in the section that follows.

2.1. Components of the Thermal Power Network

The primary components of a thermal power network include the governor, turbine, and load. Each of these components plays a crucial role in the network’s operation and stability.

2.1.1. Governor

The governor is a mechanical device used to monitor and regulate an engine’s speed. Its primary function is to adjust the engine’s median speed when the load fluctuates. The block diagram of the speed governor for a power system is shown in Figure 1. In this figure, the governor constant time is denoted as TG, and the speed-governing mechanism is demonstrated with an unstable fall correction Gc(s). Other parameters include the mechanical generator opening time TM, the damping ratio D, the droop gain R, the turbine constant time Tr, the variation in frequency Δf, the change in generator power output ΔP, the turbine’s mechanical power output change ΔPm, and the power load change ΔPL.

2.1.2. Turbine

A turbine is a rotating mechanical machine that produces useful work by converting energy from the flow of a liquid, such as steam vapor, water, or air, into rotation. The transfer function model of the turbine is depicted in Figure 2. The time lag between the valve’s switching position and the turbine torque is represented by Tch.

2.1.3. Load

The electrical grid is subject to various types of loads. Figure 3 illustrates the block structure of the load mathematical model. In this model, H represents the generator inertia constant, and ΔPe denotes the change in electrical power.
Governor   1 1 + s T s g
Reheater = 1 + s K r T r 1 + s T r
Steam   generator = 1 1 + s T t
While time constants for the steam turbine, reheater, as well as governor are denoted, Tsg, Tr, and Tt. Figure 4 shows a mathematical model representation for a single-area thermal electrical network.
The governor, turbine, and load components are interconnected to form the thermal power network. The governor adjusts the turbine’s mechanical power output ΔPm in response to changes in frequency Δf and load ΔPL. The turbine then converts this mechanical power into electrical power, which is supplied to the grid. The load component models the variations in electrical power demand, impacting the overall stability and performance of the power network.

2.2. Numerical Modeling of Wind Energy Network

The explanation for wind power system modeling is as follows:

2.2.1. Wind Speed Modeling

The wind resource of a wind energy system is a critical component that determines how much electricity can be generated, under ideal conditions, that wind turbine power can capture proportional to the cube of wind speed. The kinetic energy of wind can be computed using the following expression:
E = 1 2 m v 2
where m is the moving mass of air and v motion speed. At that time, wind speed is reported as
P w i n d = E Δ t
Typically, a scalar function that changes over time to represent wind speed v is V = f(t). It can alternatively divide into two distinct elements, which stand for variations in wind: a part that varies slowly, marked as V0, and a part that varies randomly, denoted as Vt. Consequently, wind speed can be expressed as below:
V ( t ) = V 0 + V t ( t )
The literature provides three methods for modeling wind speed profiles mathematically:
The first technique is a white noise filtering technique, in which a low-pass filter with subsequent transfer function is utilized to correct the impact of turbulence:
F ( s ) = 1 1 + τ × s
τ filters constant time, which depends upon the average wind speed, rotor diameter, as well as the intensity of wind turbulence.
The second way to create a wind speed profile uses meteorologist I. Van der Hoven’s established spectral density to characterize fluctuations in wind speed. As a result, wind speed variation V(t) is expressed as a harmonic sum:
V v ( t ) = A + i = 1 n α k · sin ( ω k · t )
i is the final harmonic rank retained in wind profile computation, αk amplitude of K-order modulation, ωk vibration of K-order modulation, and A average wind speed.
The third technique, the Weibull distribution, measures the average wind speed at average periods to determine the wind potential for a given site. A histogram is then used to divide acquired data into numerical values according to wind speed classifications. Given the chosen time instant and Weibull distribution, the wind profile is given as
V v ( t ) = ( 1 + ξ v ( t ) ξ v ) · V v
where ξ v means the value of disturbance and Vv is the average wind speed, given by
ξ v ( t ) = ( ln ( r a n d ( t ) ) C v ) 1 k v
where (Cv, kv) is a duplet of parameters found in analyzing the wind class histogram, and rand(t) is a function that generates random values between 0 and 1 in a uniform distribution. In general, Cv is a scaling factor larger than 5. If the histogram resembles a normal distribution, which is defined by a regular division throughout a mean value, then the shape factor kv is larger than 3.

2.2.2. Modeling of Wind Turbine

Wind energy is a primary nonconventional energy source. Through the wind power network, the capacity of the wind turbine generator (WTG) is combined with the current energy network [46]. The dynamic model for WTG is as follows:
Δ P W T G = 1 T W T G Δ P ω 1 T W T G Δ P W T G
Pw denotes wind power, TWTG is the WTG time constant, and ΔPWTG denotes the WTG output power change. The blade-equipped rotor of a turbine converts wind energy into mechanical energy. The following formulas can be used to numerically express the rotor’s extracted wind power [47]:
P r o t o r = 1 2 ρ A V 3 C p
where A stands for the sweeping area, V is the wind speed, Cp is the power coefficient, and ρ is the wind density. The relationship between the input wind speed as well as active power is shown in the following way:
P G W = ρ a 2 V ω 3 C p 2 ( T S R , β )
β is the blade angle, TSR is the tip speed ratio, α is the area density, and Vω is the wind velocity. Rotor efficiency Cp:
C p = T S R 0.022 β 2 5.6 2 e 0.17 T S R T S R = r p m π D 60 V
Turbine-generated torque is expressed by
T t = P G W ω t
ωt represents the wind turbine rotor’s angular rotational speed.

2.2.3. Gearbox Modeling

A wind turbine’s mechanical components are made up of a turbine shaft, which rotates slowly at a speed of Ωt, and a gearbox, which has a multiplication gain of G and powers the generator Ωg through a fast secondary shaft.
By using a gearbox, the turbine speed of Ωt can be multiplied by a multiplication gain G in order to adjust to the generator’s higher speed of Ωg. The elasticity, friction, and energy loss of the gearbox are deemed insignificant, making this device efficient. The following two equations represent the mathematical model of this device’s operation:
T g = T a e r G Ω t = Ω g G
where Taer is the wind turbine’s aerodynamic torque; Ωg is the speed of the generator shaft; Tg is the torque on the generator shaft; G is the multiplication gain, which is defined as G = N1/N2; and Ωt is the turbine speed shaft.
Generator inertia Jg and turbine inertia Jt make up total inertia J, which can be expressed using the following equation:
J = J t G 2 + J g
The generator friction constant fg and turbine friction constant ft add up to the total viscous friction constant fv. Following is an expression for constant fv:
f v = f t G 2 + f g
Tmec is the total mechanical torque, which determines generator speed Ωg. This torque is the sum of torques supplied to generator shaft Tg, viscid friction torque Tv, and electromagnetic torque of generator Tem.
T m e c = J × d Ω g d t
T m e c = T g T e m T v
T v = f × Ω g
Consequently, the following represents the differential equation for mechanical network dynamics:
J × d Ω g d t = T g T e m T v i s
Figure 5 displays a block diagram for the mechanical part of the wind turbine, and different power coefficient (Cp) numerical formulas are given in Table 2.
In this article, the wind generator’s generating output is PMSG. Figure 6 and Figure 7 show block schematics of PMSG-wind turbines employing P & O MPPT. In MATLAB/Simulink, 2022a a control scheme for wind turbines with PMSG has been created, and it will be covered in Section 3. The following equations represent the PMSG wind turbine mathematically [48,49]. The output of the mathematical model, which is the rotation speed, is produced by the electromechanical torque of the generator and rotor torque acting as inputs.
d w g e n d t = 1 2 H g e n P e l e c w g e n + w 0 D t g ( w g e n w r o t ) k t g Δ θ m
d w r o t d t = 1 2 H r o t P m e c h w r o t + w 0 + D t g ( w g e n w r o t ) + k t g Δ θ m
d ( Δ θ m ) d t = w b a s e ( w g w t )
wgen denotes the speed of the generator, Pelec is the electric power, w0 is the initial velocity, wrot is the turbine speed, and Pmech is the mechanized power; a stable state appears when wgen = wrot, so d(Δθ)/dt = 0, and Pelec = Pmech. Dtg, Ktg, and wbase are constants. Mass’s geometrical dissipation demonstrates the inertial constant; the inertial moment was evaluated utilizing
H r o t o r = J r o t o r w r o t o r 2 2 P n ; H g e n = J g e n w g e n 2 2 P n
The wind rotor’s moment of inertia can be computed generally using
J r o t o r = 1 8 m r R 2
The rotor’s mass and radius are indicated by the letters mr and R, respectively. The stator output voltages of the generator, d-q, are therefore stated as follows:
V d = R d I + L d d I d d t ω g e n L q I q
V q = R q I q + L q d I q d t + ω g e n ( L d I d + φ f )
L is the generator’s inductance, R is the resistance, I is the current within d-q axes, φf is the permanent magnetic flux, and ωgen is the PMSG’s rotational speed.
ω g e n = P p ω r e f
Pp stands for the number of pole pairs. The formula for electromechanical torque, or Tgen, is
T g e n = 3 2 P p ω r e f ( L q L d ) i d i q + φ f i q

2.3. Mathematical Modeling of PV Energy System

Photovoltaic (PV) energy is an additional fundamental as well as widespread nonconventional energy source. PV panels are used to convert this energy into a current, which is then sent over conductors. To study PV power used in power networks, a basic model of PV panels can be created by considering connections between various electronic components [15,50].
The PV current is defined by the following formulas.
I = I P h I d I R s h
Iph, Id, and IRsh show the photocurrent, current across the diode, and current across the shunt resistance. The PV produces a photocurrent upon the stated temperature when exposed to solar. Iph is expressed as follows:
I P h = I S C ( S 1000 ) + K t ( T T R )
S is the solar radiance, ISC is the current across the short circuit, Kt is the temperature co-efficient, T is the PV temperature, and TR is the reference temperature. Id is expressed as follows:
I d = I o exp ( V + I R s n J T ) 1
The voltage across terminal is represented by V, Rs is the resistance across series, n is the diode ideal factor, and J is Boltzmann’s constant. Io is the saturation current, expressed as follows:
I o = I d ( T T R ) 3 exp q E g n J ( T T R T R T )
Id is the reverse current across the diode, Eg is the cell bandgap energy, and q is the electron charge. IR is expressed as follows:
I R = V + I R s R h
The PV grid’s 1st-order transfer function is as follows [50,51]:
Δ P v = K p v 1 + s T p v Δ P s o l a r
Tpv is the photovoltaic temperature, and Kpv is the specified PV gain (one). The majority of the mathematical connections between two- and one-diode photovoltaic systems are identical [52]. The fundamental circuits for single- and two-diode PV systems are shown in Figure 8 and Figure 9.
This paper describes a fuzzy-based MPPT PV system that uses MATLAB/Simulink. Figure 10 and Figure 11 display this control approach’s block illustration, including the fuzzy MPPT technique. In Figure 11, Iin(k) is the current, Vin(k) is the voltage, Pin(k) is the power, Pin (k − 1) is the preceding power, Vin (k − 1) is the preceding voltage, and E is the instant as well as the preceding value of voltage as well as power.

2.4. Fuel Cell System Modeling

This is a complicated system consisting of hydrodynamic, thermal, and electrochemical variables. The proportional relationship between a gas’s partial pressure and flow through a valve can be expressed as follows:
Q H 2 p H 2 = L a n M H 2 = L H 2
QH2 is the hydrogen molar fluid, PH2 is the partial pressure of hydrogen, Lan is the anode valve constant, MH2 is the hydrogen molar mass, and LH2 is the hydrogen valve molar constant.
Q H 2 O p H 2 O = L a n M H 2 O = L H 2 O
QH2O is the water molar fluid, PH2O is the water vapor partial pressure, MH2O is the water molar mass, and LH2O is the molar constancy water valve.
d d t p H 2 = R T V a n ( Q H 2 i n Q H 2 o u t Q H 2 r )
The following represents the basic electrochemical relationship between the stack current and hydrogen flow:
Q H 2 r = N I 2 F = 2 K r I f c
N is the stack’s number, Ifc is the current of the fuel cell, F is Faraday’s constant, and Kr is the modeling constant. Using Laplace transformation upon (23) and (27), the hydrogen partial pressure is written in the s domain as follows:
p H 2 = 1 / L H 2 1 + τ H 2 s ( Q H 2 i n 2 K r I f c )
τ H 2 = V a n L H 2 R T
τH2 is the hydrogen constant time, Van is the volume across the anode, T is the temperature across the stack, and R is the universal gas constant. The output voltage is given below:
V c e l l = E + η a c t + η o h m
Nernst voltage, when given in terms of gas molarities, is as follows:
E = N E 0 + R T 2 F log p H 2 p O 2 0.5 p H 2 O
E is Nerst’s voltage, E0 is the cell open voltage, Vcell is the voltage across the cell, and ηact is the ohm overvoltage.

2.5. Double-Area Tie-Line Modeling

The electrical network proposed in this article is the dual-area thermal power network, shown in Figure 12. The primary components of each area consist of a turbine, generator, and speed-regulating device with two outputs and three inputs. The dual-area tie-line IPS block illustration is displayed in Figure 12. Consequently, both model’s components are equipped with a PID controller. Control signals (Figure 12) are represented by u1 and u2, power demand variations are represented by ΔPD1 and ΔPD2, power variation in the generator is represented by ΔPg, power variation in the turbine is represented by ΔPt, output area control errors are represented by ACE1 and ACE2, the frequency bias co-efficient is represented by B1 and B2, and fluctuations in system frequency are represented by Δf1 and Δf2.
B1 and B2 of the dual-area power network are calculated as follows:
B 1 = 1 R 1 + D 1
B 2 = 1 R 2 + D 2
A C E 1 = B 1 Δ f 1 + Δ P t i e 1
A C E 2 = B 2 Δ f 2 + Δ P t i e 2
Utilizing specified schemes along with transfer functions, suggested networks were designed; Simulink models were designed for analysis within MATLAB/Simulink for LFC. Using PSO and PSO-AHA approaches, PID controller gains were tuned. Table 3 and Table 4.

3. Suggested Control Scheme

This section explains the suggested study, a single-area energy network linked with a PMSG wind turbine based upon the P & O MPPT technique, a fuzzy-based MPPT solar PV, as well as a fuel cell, in addition to double-area tie-line IPS. PID controller gains are adjusted using PSO-AHA and also PSO approaches. MATLAB/Simulink is used to simulate suggested control research in order to investigate and regulate the frequency. Figure 12 and Figure 13 display block illustrations for dual-area as well as single-area multiple-source energy networks.

3.1. PID Controller

The primary goal of the controller is to determine the optimal set of controls that will enable the system to continuously reach the intended state with the fewest possible variations. PID controllers are extensively used in control as well as automation domains due to enhanced functionalities along with well-proven design methodologies [53]. The sum of derivatives, along with the proportionate integral of derivatives and the proportionately integral controller outputs, is the output it generates. Figure 14 shows the fundamental configuration of the PID controller. R(s) is the input signal, the error signal is represented by E(s), the control signal is represented by U(s), and the output signal is represented by Y(s) in Figure 14.
Numerically,
u ( t ) e ( t ) + e ( t ) + d d t e ( t )
The transfer function is expressed below:
U ( s ) E ( s ) = K P + K I s + K D s
KP, KI, and KD represent controller gains. The transfer function is expressed as
K P 1 + K I K P s + K D K P s K P 1 + 1 T i s + T D s T i = K P K I & T D = K D K p
The PID controller determines its control value by taking errors from the past, present, as well as future into account.
It is used in conjunction with the one-area multiple-source network in addition to dual-area tie-line IPS to maintain the system’s intended frequency and also to decrease power variations.

3.2. Objective Function

A scale called objective function can be used to optimize controller gains. In this article, the ITAE objective function is used to modify the PID controller’s gains. It has the following mathematical representation [12]:
I T A E = t = 0 t = f i n a l Δ f × t × d t
Suggested approach tunings utilizing PSO along with PSO-AHA algorithms, with ITAE being the cost function, are covered in the sections that follow.

3.3. One-Area Multi-Source Power Network

A RESs-connected single-area energy network model was created within the Simulink environment. Table 3 displays the constraint values of the single-area thermal power network; Table 4 displays the electrical properties along with the parameter values of fuel cell, solar, and wind power sources.
In Section 2, an illustration showing a wind turbine along with the PMSG-based P & O MPPT technique and also the solar PV fuzzy-based MPPT technique is provided. Fuzzy rule ranges, membership functions, as well as types for solar PV MPPT, in addition to fuzzy rules, respectively, are displayed in Table 5 and Table 6. Table 5 shows negative big NB, negative small NS, zero ZE, positive small PS, and positive big PB. To determine PID optimal gains, PSO and PSO-AHA strategy operators were used. PSO, as well as AHA operators, are shown in Table 7. The optimal gains of the PID controller are shown in Table 8.

3.4. Dual-Area Tie-Line IPS LFC

The need for power in commercial, industrial, and residential sectors is increasing. Since it reduces reliance, integrating renewable energy is a superior option. Standard operating conditions can be changed by deregulating the electrical grid. Distributed generations in dual-region energy systems have attracted a lot of attention. Most modern power networks are made up of different control zones along with different generating sources and non-identical capacities to handle an excessive load. Any variation in generating and consumer load causes the area frequency to accidentally fluctuate from the set level, which, in turn, causes power-sharing with nearby control zones. Two thermal units make up the two-area power network tie-line IPS used in this study. Tie-lines provide the transmission of electricity between connected sectors. When adverse circumstances transpire, like extreme load disruptions, the control unit monitors changes in tie-line power and makes periodic attempts to return the system to its typical state. As a result, ITAE is represented as follows and is considered a goal function:
I T A E = 0 t ( Δ f 1 + Δ f 2 + Δ P t i e ) × t × d t
where ΔPtie represents the tie-line frequency variation, and Δf1 and Δf2 denote the frequency variation in areas 1 and 2, respectively. In Section 2, a schematic illustrating this control approach, along with the mathematical model, is provided. The relevant model values for two-area IPS are given in Table 9. Table 10 displays the configuration of the PSO and AHA parameters, while Table 11 displays the PID controller’s optimum values.

3.5. PSO Algorithm

The PSO algorithm was created by Dr. Kennedy and Eberhart in 1995, with influence from collective nature observed in fish-schooling and bird-swarming processes. Every solution in the search field is referred to as a “particle” in the PSO approach. The optimizing process is determined by the fitness values of the objective function. The particle searches space for best values, both generally as well as individually, and its momentum directs its path. Initially, the swarm is put together using a set of randomly distributed particles; then, iterations are made in order to look for optimization. To achieve optimal gain values, principles of both personal as well as global are best applied to each particle. After every cycle, the best response, referred to as the local best, is determined. The term “global best” refers to the final best value, which is found after tuning each local optimal value [11].
There are N particles in population within the D-dimensional search space, and the position of the ith particle is presented as Xi = (xi1, xi2xiD) in the D-dimensional space, with the velocity Vi = (vi1, vi2viD), i = 1,2…N. An iteration process is used to update the locations and velocities of the particles while determining the ideal values for each individual and collective particle creation.
The following equation gives the velocity of particles:
V k + 1 i = ω V k i + c 1 r 1 ( P k i X k i ) + c 2 r 2 ( P k g X k i )
The positions of the particles are constantly being updated by
X k + 1 i = X k i + V k + 1 i
X k i is the position of the particle, V k i is the velocity of the particle, P k i is the best position of the individual particle, P k g is the best position for a swarm, c1 and c2 are the cognitive and social parameters, and r1 and r2 are the random numbers between zero and one. The algorithm for PSO is depicted in the flow chart that can be found in Figure 15. Due to its simple structure and effectiveness, the PSO algorithm is gaining popularity. Furthermore, in terms of performance, it outperforms gradient-based optimization techniques, which necessitate the differentiation of the optimization issue. But if a particle gets trapped in a nearby optimal state, the PSO algorithm may produce suboptimal solutions.

3.6. AHA Algorithm

AHA is the newest biologically inspired optimization approach employed in engineering applications, and it was unveiled by year’s end in 2021. The concept is derived from the intelligence exhibited by hummingbirds [54]. Similar optimization techniques are employed in the development of the exploration and exploitation phases, as well as the AHA. Three elements—food sources, optimization, and a visit table—are used to characterize the development procedure for addressing various optimization challenges. The general attributes of sources are evaluated by hummingbirds, including the nectar content and grade of particular flowers, the rate of nectar production, and the most recent visitation date of the flower. Hummingbirds may communicate with their fellow hummingbirds about the location and pace of nectar replenishment at a certain food source. The hummingbird visiting table is utilized to document the count of hummingbirds, which consume food from each source. The visit table is typically modified at the end of each iteration. Development is influenced by three distinct foraging techniques: directed, territorial, and migratory foraging. The following mathematical representations depict the three distinct foraging strategies.

3.6.1. Initialization

A population of n hummingbirds is assigned to n food sources at random [1].
x i = L o w + r × ( U p U L ) i = 1 , , n
Here, UL and UP correspond to the lower and upper limits; xi is the place that contains the ith food source, which is the answer to a certain problem; r is a random vector in the range of [0, 1]; and the food source visit table is set up as follows:
V T i , j = 0 if   i j null i = j       i = 1 , , n ; j = 1 , , n
In the given scenario, it can be observed that when i = j, the value of VTi,j is null, indicating that a hummingbird is actively consuming food from a specific source. Specifically, this signifies that the hummingbird indexed by i has recently visited the food source. On the other hand, when i ≠ j, the value of VTi,j is 0, indicating that the hummingbird indexed by j is currently engaged in the iteration.

3.6.2. Guided Foraging

Every individual hummingbird selects the nectar source that offers the most quantity of nectar. Hummingbirds exhibit three distinct modes of flying, including axial flight, diagonal flight, and omnidirectional flight. The term axial flight can be defined as follows:
D ( i ) = 1 if   i = randi ( [ 1 , d ] ) 0 else i = 1 , , d
The following is a description of a diagonal flight:
D ( i ) = 1 if   i = P ( j ) , j [ 1 , k ] , P =   randperm   ( k ) , k [ 2 , [ r 1 × ( d 2 ) ] + 1 ] 0 else       i = 1 , , d
The following is a description of an omnidirectional flight:
D ( i ) = 1 i = 1 , , d
The function randperm (k) generates a permutation of integers ranging from 1 to k, whereas the function randi [1, d] selects a number at random between 1 and d. Additionally, r1 represents a random number within the range of [0, 1]. The subsequent expression is the mathematical equation utilized for the purpose of modeling directed foraging behavior in the presence of a suitable food source:
v i ( t + 1 ) = x i , t a r ( t ) + a × D × ( x i ( t ) x i , t a r ( t ) )
a N ( 0 , 1 )
The position of the ith hummingbird’s desired food source could be represented as xi,tar(t), where a is a guiding factor. The location of the ith food source at time t, denoted as xi(t), follows a normal distribution N (0, 1) with a mean of 0 and a standard deviation of 1.
The following is the latest position update for the ith food source:
x i ( t + 1 ) = x i ( t ) f ( x i ( t ) ) f ( v i ( t + 1 ) ) v i ( t + 1 ) f ( x i ( t ) ) > f ( v i ( t + 1 ) )
The hummingbird discontinues its use of the present food source and instead consumes from the potential food source indicated by Equation (62) if the potential food source’s rate of nectar replenishment surpasses that of the current food source Equation (64).

3.6.3. Territorial Foraging

The hummingbird has a remarkable ability to efficiently navigate within its established territory, swiftly relocating to a neighboring area in search of an alternative food source that may offer more benefits compared to its present one. The above mathematical equation outlines the local foraging method employed by hummingbirds in conjunction with an appropriate food supply that aligns with their territorial foraging approach.
v i ( t + 1 ) = x i ( t ) + b × D × x i ( t )
b N ( 0 , 1 )
The variable b represents a territorial component that follows a normal distribution with a mean of 0 and a standard deviation of 1, denoted as N (0, 1), using its unique flight abilities.

3.6.4. Migration Foraging

The migration of a hummingbird from the nectar source with the lowest nectar refilling rate to a random one might be characterized as follows:
x w o r ( t + 1 ) = L o w + r × ( U p L o w )
where xwor is the food for which the population’s rate of nectar replenishment is the lowest. In the absence of any other food sources, a hummingbird engaging in both directed foraging and territorial foraging would sequentially visit each food source as its target source throughout each iteration, as indicated by the visit table. Let us consider a scenario where the selection between the territorial and directed foraging stages is characterized by a probability of success of 50%. Additionally, in the context of guided foraging, the likelihood of success in visiting every alternate source is also 50%. In this scenario, the adoption of a migratory foraging strategy becomes necessary in order to disrupt the state of equilibrium and expand the scope of the search region. As a consequence, the proposed population size parameters for the migration coefficient are as follows:
M = 2n
The algorithm for AHA is depicted in the flow chart that can be found in Figure 15.

3.7. PSO-AHA Hybrid Algorithm

The hybrid PSO-AHA optimization algorithm is used to optimize the KP, KI, and KD values for the PID controller in a single-area multi-source and two-area tie-line IPS to achieve the desired objectives, as mentioned in Figure 15. We will outline the steps involved in the AHA cycle and the PSO cycle and then explain the decision-making process for incorporating both into a hybrid PSO-AHA strategy.

3.7.1. AHA Algorithm Cycle

AHA algorithm process are given as:

3.7.2. Initialization

Specify the values for the number of agents NP, max_iter, d, nvars, Low, and Up for KP, KI, and KD. Set up the random positions of the hummingbirds (agents) within the boundaries.

3.7.3. Fitness Evaluation

Determine the best possible fitness value for each hummingbird using the objective function associated with PID controller performance.

3.7.4. Position Update Based on Foraging Strategy

Determine whether territorial foraging applies to each hummingbird (rand < 0.5). If so, adjust the position using the foraging strategy: (59) should be used to update the location if r < 1 = 3. Update the position with Equation (60) if r < 2 = 3. Otherwise, use Equation (61) to update the location. Regardless of the outcome of the territorial foraging, update the location using Equation (62).

3.7.5. Calculation of the Objective Function

Determine the objective function for the updated positions to evaluate the effectiveness of the new solutions.

3.7.6. Migration Foraging

If the number of iterations exceeds maxiter, execute migration foraging to possibly prevent local optima.

3.7.7. Record Best Position

At the end of each iteration, or when the termination criteria are met, note the best position identified and its fitness value.

3.7.8. PSO Algorithm Cycle

The PSO algorithm process is given as follows:

3.7.9. Initialize PSO Parameters

Define variables such as NP, wmin, wmax, c1, and c2. Specify the initial positions and velocities of the particles.

3.7.10. Evaluate Fitness

Compute the fitness of each particle by applying the objective function.

3.7.11. Update Individual and Global Best

Update the position of each particle according to its best-known position and the swarm’s best-known position after evaluating its fitness.

3.7.12. Update Velocity and Position

Equation (55) is used to update the velocity of each particle depending on its current velocity, the best position of the swarm, and its own position. Equation (56) should be used to update each particle’s position, taking into account the newly calculated velocity.

3.7.13. Termination Check

If the termination conditions are met (which may be based on the number of iterations, the degree of improvement, etc.), proceed to the hybrid decision to make.

3.7.14. Hybrid Decision of PSO-AHA Optimizer

The hybrid decision of the PSO-AHA Optimizer process is given as:

3.7.15. Evaluate Performance

Determine the performance of both AHA and PSO in optimizing KP, KI, and KD values at each iteration. Determine the optimal position and fitness value achieved by both algorithms.

3.7.16. Decide Next Steps

Determine whether to proceed with the AHA cycle, the PSO cycle, or a hybrid of both based on their performance. This decision might be based on an enhancement in fitness.

3.7.17. Termination Check

Determine if the termination conditions have been satisfied (according to the fitness threshold). Once the termination requirements are satisfied, finalize the results. If not, proceed with the hybrid loop.

3.7.18. Finalize Results and Generate Output

Once the termination conditions are met, identify the optimal KP, KI, and KD values derived from the hybrid algorithm. Determine the global best particle and its corresponding PID controller parameters. Output the optimal KP, KI, and KD values for the PID controller to improve the LFC in the one-area multiple-source and two-area tie-line IPS.
This hybrid PSO-AHA optimizer strategy will repeatedly refine the KP, KI, and KD settings in order to improve the PID controller capabilities within the proposed networks. The AHA enables exploration of the search space, whereas PSO concentrates on the exploitation of the best solutions identified. The hybrid technique obtains the strengths of both algorithms, conceivably resulting in a more robust and effective optimization process.

4. Simulation Results

By analyzing results using PSO-PID as well as PSO-AHA-PID and also contrasting them with the PSO-PID controller, this section shows the superiority of the suggested control approach. Based upon MATLAB 2022a results, the proposed power network was analyzed into two cases:
  • Variations in system parameters in dual-area tie-line linked power systems and one-area multi-source power networks.
  • Load changes within dual-area tie-line IPS.

4.1. Multi-Source Single-Area Power Network

Figure 16 displays the results achieved using PSO-AHA-PID and PSO-PID. The transient responses of the suggested power network substantially improved in terms of signal characteristics. Contrasted with the PSO-PID controller, the PSO-AHA algorithm successfully adjusted PID controllers’ gains, producing the best results in terms of incredibly low undershoot and quick response times.

Robustness Investigations (Case Study 1)

Continuous changes in system parameters can severely harm closed-loop system performance. Thus, to examine the effect of parametric variations on the system, experiments were carried out. To achieve this, the initial value of each system parameter was adjusted by approximately ±75%. Table 12 illustrates two potential situations of parameter uncertainty for a single-area power network.
Under parametric uncertainty, frequency responses of single-area multiple-source energy networks are displayed in Figure 17 and Figure 18. The frequency response of PSO-AHA-PID is superior to that of the PSO-PID controller, with less overshoot as well as undershoot and an exceptionally quick settling period. The frequency response signal characteristics of a one-area multiple-source power network are displayed in Table 13.

4.2. Two-Area Tie-Line IPS

The PSO-AHA-PID and PSO-PID controller results are displayed in Figure 19. Regarding actual frequency, tie-line power variations, variation of frequency in Pu, as well as alterations of frequency in Hz, the suggested dual-area tie-line IPS shows significantly improved transient responses. The PSO-AHA approach optimized gains of PID. When contrasted to PSO-PID, it provides optimal performance in terms of incredibly little undershooting and exceptionally quick response times.

4.2.1. Robustness Investigations (Case Study 1)

Variability affects a system’s stability. The LFC network required control action against constraint variations in the controlled network to achieve an acceptable level of resiliency. Thirteen cases for constraint variations were taken to analyze the resiliency of the proposed PSO-AHA-PID as well as the PSO-PID controller within double-area IPS. Each network constraint separately varied by ±75% from its initial value to start this analysis.
In scenarios 1–12, just one constraint was altered once, varying the initial value by +75%/−75%; Table 14 shows that constraints Tg, Tt, H, and R changed +75% in both areas in scenario 13. Concurrently, there was a −75% change in B and D. Figure 20, Figure 21, Figure 22, Figure 23, Figure 24, Figure 25, Figure 26, Figure 27, Figure 28, Figure 29, Figure 30, Figure 31 and Figure 32 depict the dynamic outcomes of double-area tie-line IPS with PSO-AHA-PID as well as PSO-PID under constraint-ambiguity conditions. The frequency response signal characteristics of dual-area tie-line IPS are listed in Table 15, where high and low frequencies are represented by H/L, Osh represents overshoot, Psh represents pre-shoot, and Ush represents undershoot.

4.2.2. Load Variations (Case Study 2)

Four cases were considered against fluctuations in load, as indicated in Table 16, to test the flexibility of the system.
Figure 33, Figure 34, Figure 35 and Figure 36 display the results obtained for dual-area load alterations. The frequency responses for dual-area signal characteristics in load variations are displayed in four scenarios in Table 17. The proposed study examined the effectiveness of two power system models, a two-area tie-line IPS and a single-area multi-source power network for LFC. We analyzed these systems’ responses to a range of scenarios, including changes in load and parameter values.
We assessed the system performance for the single-area multi-source power network by adjusting system parameters within ±75%. This system includes fuzzy-based MPPT solar power, PMSG wind power utilizing P&O MPPT, fuel cells, and isolated thermal power systems. These changes, as shown in Table 12, were implemented methodically to evaluate how robust the control measures were. The measured results for the frequency response characteristics undershoot, overshoot, and settling time are shown in Table 13 and Figure 16, Figure 17 and Figure 18. As shown in Table 14 and Table 16, we investigated the response of the two-area tie-line linked power system under 13 distinct scenarios of parameter adjustments and 2 scenarios of load variations in Area 1 and Area 2.
Dynamic Outcomes Under Constraint Ambiguity: Figure 20, Figure 21, Figure 22, Figure 23, Figure 24, Figure 25, Figure 26, Figure 27, Figure 28, Figure 29, Figure 30, Figure 31 and Figure 32 present the dynamic outcomes of the two-area tie-line IPS when subjected to different scenarios of parameter variations. These figures illustrate the system’s frequency response under conditions of constraint ambiguity, comparing the performance of the PSO-AHA-PID controller with that of the PSO-PID controller.
Stability and Frequency Response: The PSO-AHA-PID controller demonstrates superior stability and better frequency response characteristics compared to the PSO-PID controller across all scenarios. Specifically, the PSO-AHA-PID controller consistently shows lower overshoot, undershoot, and settling times, indicating a more robust performance.
Resilience to Parameter Variations: The figures highlight the resilience of the PSO-AHA-PID controller to significant variations in system parameters (±75%). The frequency deviations in both areas are minimized more effectively by the PSO-AHA-PID controller, showing a better ability to handle uncertainties and maintain system stability.
Frequency Response Characteristics: Table 15 provides detailed frequency response characteristics for both Area 1 and Area 2, as well as tie-line power responses under various parameter-variation scenarios.
Performance Metrics: The metrics listed in Table 15, including undershoot (Ush%), overshoot (Osh%), pre-shoot (Psh%), high/low frequencies (H/L Hz), and fall time, clearly show that the PSO-AHA-PID controller outperforms the PSO-PID controller. For instance, the PSO-AHA-PID controller has lower overshoot and undershoot percentages and faster settling times, reflecting improved dynamic performance.
Consistency Across Scenarios: The consistency of the PSO-AHA-PID controller’s performance across different scenarios is evident. It maintains more stable and desirable frequency characteristics, regardless of the specific parameter changes applied, demonstrating its robustness and adaptability to varying operating conditions.

5. Conclusions and Future Directions

5.1. Conclusions

This study presents a novel approach to Load Frequency Control (LFC) for dual-area tie-line interconnected power systems (IPS) and single-area multiple-source power networks. The innovation lies in the application of Particle Swarm Optimization (PSO) combined with the Artificial Hummingbird Algorithm (AHA) to fine-tune Proportional–Integral–Derivative (PID) controller gains, resulting in the PSO-AHA-PID controller. This hybrid approach leverages the exploratory capabilities of AHA and the exploitative strengths of PSO, ensuring robust and optimized control settings for LFC.
Key Scientific Contributions and Their Validation are given as:

5.1.1. Enhanced Frequency Stability

5.1.2. Robust Performance across Scenarios

5.1.3. Reduction in Power Variations

  • Contribution: The methodology effectively reduces tie-line power variations, a critical factor for maintaining system stability in interconnected power networks.
  • Validation: The PSO-AHA-PID controller achieves lower overshoot, undershoot, and faster settling times compared to the PSO-PID controller. For example, in the one-area multi-source network, the settling time is reduced to 0.6 s with the PSO-AHA-PID controller compared to 5.5 s with the PSO-PID controller (Table 13, Figure 17 and Figure 18).

5.1.4. Scalability and Applicability

5.2. Future Directions

Stability issues with linked power networks for nonconventional energy sources remain a significant concern. Future research will focus on the integration and maximization of different renewable energy sources. Additionally, the emergence of innovative technologies such as artificial intelligence, cloud computing, and big data holds great potential for addressing LFC-related issues. Future research will address the following areas:
  • Nonconventional Energy: Continued research into nonconventional energy sources is vital as they introduce variability and complexity into power generation. The integration of LFC systems with RESs presents challenges in stabilization, necessitating effective device matching and information transfer.
  • Flexibility: A highly variable and secure linked power network is required due to the high penetration of RESs. Enhancing the flexibility of multi-area power systems will involve developing new optimization approaches for control, storage, and market strategies.

Author Contributions

Conceptualization, W.Y.; writing—original draft preparation, W.Y., M.Z.Y. and A.T.; validation and formal analysis, H.G.M.Q., M.T. and E.G.; review—writing and editing. All authors have reviewed and approved the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This project was funded by King Saud University through Researchers Supporting Project number RSPD2024R685, King Saud University, Riyadh, Saudia Arabia.

Data Availability Statement

Data are available on request from the corresponding author.

Acknowledgments

The authors acknowledge “Researchers Supporting Project number (RSPD2024R685), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Arya, Y.; Dahiya, P.; Çelik, E.; Sharma, G.; Gözde, H.; Nasiruddin, I. AGC performance amelioration in multi-area interconnected thermal and thermal-hydro-gas power systems using a novel controller. Eng. Sci. Technol. Int. J. 2021, 24, 384–396. [Google Scholar] [CrossRef]
  2. Elkasem, A.H.A.; Khamies, M.; Hassan, M.H.; Agwa, A.M.; Kamel, S. Optimal design of TD-TI controller for LFC considering renewables penetration by an improved chaos game optimizer. Fractal Fract. 2022, 6, 220. [Google Scholar] [CrossRef]
  3. Ameli, A.; Hooshyar, A.; El-Saadany, E.F.; Youssef, A.M. Attack detection and identification for automatic generation control systems. IEEE Trans. Power Syst. 2018, 33, 4760–4774. [Google Scholar] [CrossRef]
  4. Mohammadikia, R.; Aliasghary, M. A fractional order fuzzy PID for load frequency control of four-area interconnected power system using biogeography-based optimization. Int. Trans. Electr. Energy Syst. 2019, 29, e2735. [Google Scholar] [CrossRef]
  5. Ali, T.; Malik, S.A.; Hameed, I.A.; Daraz, A.; Mujlid, H.; Azar, A.T. Load frequency control and automatic voltage regulation in a multi-area interconnected power system using nature-inspired computation-based control methodology. Sustainability 2022, 14, 12162. [Google Scholar] [CrossRef]
  6. Ali, T.; Malik, S.A.; Daraz, A.; Aslam, S.; Alkhalifah, T. Dandelion optimizer-based combined automatic voltage regulation and load frequency control in a multi-area, multi-source interconnected power system with nonlinearities. Energies 2022, 15, 8499. [Google Scholar] [CrossRef]
  7. Metwally, M.A.; Ali, M.A.; Kutb, S.A.; Bendary, F.M. A genetic algorithm for optimum design of PID controller in multi area load frequency control for Egyptian electrical grid. Int. J. Eng. Res. Technol. 2024, 5, 267–275. [Google Scholar]
  8. Hu, G.; Zhong, J.; Zhao, C.; Wei, G.; Chang, C.-T. LCAHA: A hybrid artificial hummingbird algorithm with multi-strategy for engineering applications. Comput. Methods Appl. Mech. Eng. 2023, 415, 116238. [Google Scholar]
  9. Morgan, E.F.; El-Sehiemy, R.A.; Awopone, A.K.; Megahed, T.F.; Abdelkader, S.M. Load Frequency Control of Interconnected Power System Using Artificial Hummingbird Optimization. In Proceedings of the 2022 23rd International Middle East Power Systems Conference (MEPCON), Cairo, Egypt, 13–15 December 2022. [Google Scholar]
  10. Ebeed, M.; Abdelmotaleb, M.A.; Khan, N.H.; Jamal, R.; Kamel, S.; Hussien, A.G.; Zawbaa, H.M.; Jurado, F.; Sayed, K. A Modified Artificial Hummingbird Algorithm for solving optimal power flow problem in power systems. Energy Rep. 2024, 11, 982–1005. [Google Scholar] [CrossRef]
  11. Jagatheesan, K.; Anand, B.; Samanta, S.; Dey, N.; Ashour, A.S.; Balas, V.E. Particle swarm optimisation-based parameters optimisation of PID controller for load frequency control of multi-area reheat thermal power systems. Int. J. Adv. Intell. Paradig. 2017, 9, 464–489. [Google Scholar] [CrossRef]
  12. Dhanasekaran, B.; Siddhan, S.; Kaliannan, J. Ant colony optimization technique tuned controller for frequency regulation of single area nuclear power generating system. Microprocess. Microsyst. 2020, 73, 102953. [Google Scholar] [CrossRef]
  13. Jagatheesan, K.; Boopathi, D.; Samanta, S.; Anand, B.; Dey, N. Grey wolf optimization algorithm-based PID controller for frequency stabilization of interconnected power generating system. Soft Comput. 2024, 28, 5057–5070. [Google Scholar] [CrossRef]
  14. 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]
  15. El-Ela, A.A.A.; El-Sehiemy, R.A.; Shaheen, A.M.; Diab, A.E.-G. Design of cascaded controller based on coyote optimizer for load frequency control in multi-area power systems with renewable sources. Control. Eng. Pract. 2022, 121, 105058. [Google Scholar]
  16. Tudu, A.K.; Naguru, N.; Dey, S.H.N.; Paul, S. Load frequency control of an isolated microgrid using optimized model predictive control by GA. Electr. Eng. 2024, 1–13. [Google Scholar] [CrossRef]
  17. Alayi, R.; Zishan, F.; Seyednouri, S.R.; Kumar, R.; Ahmadi, M.H.; Sharifpur, M. Optimal load frequency control of island microgrids via a PID controller in the presence of wind turbine and PV. Sustainability 2021, 13, 10728. [Google Scholar] [CrossRef]
  18. Vedik, B.; Kumar, R.; Deshmukh, R.; Verma, S.; Shiva, C.K. Renewable energy-based load frequency stabilization of interconnected power systems using quasi-oppositional dragonfly algorithm. J. Control. Autom. Electr. Syst. 2021, 32, 227–243. [Google Scholar] [CrossRef]
  19. Mohanty, B.; Panda, S.; Hota, P. Controller parameters tuning of differential evolution algorithm and its application to load frequency control of multi-source power system. Int. J. Electr. Power Energy Syst. 2014, 54, 77–85. [Google Scholar] [CrossRef]
  20. Wadi, M.; Shobole, A.; Elmasry, W.; Kucuk, I. Load frequency control in smart grids: A review of recent developments. Renew. Sustain. Energy Rev. 2024, 189, 114013. [Google Scholar] [CrossRef]
  21. Panwar, A.; Sharma, G.; Bansal, R.C. Optimal AGC design for a hybrid power system using hybrid bacteria foraging optimization algorithm. Electr. Power Compon. Syst. 2019, 47, 955–965. [Google Scholar] [CrossRef]
  22. Bharti, K.K.; Singh, V.P.; Singh, S. PSO-based: MARL approach for frequency regulation of multi-area power system. J. Electr. Eng. Technol. 2020, 15, 1529–1539. [Google Scholar] [CrossRef]
  23. Jagatheesan, K.; Anand, B.; Samanta, S.; Dey, N.; Santhi, V.; Ashour, A.S.; Balas, V.E. Application of flower pollination algorithm in load frequency control of multi-area interconnected power system with nonlinearity. Neural Comput. Appl. 2017, 28, 475–488. [Google Scholar] [CrossRef]
  24. Kumarakrishnan, V.; Vijayakumar, G.; Boopathi, D.; Jagatheesan, K.; Saravanan, S.; Anand, B. Optimized PSO technique based PID controller for load frequency control of single area power system. Solid State Technol. 2020, 63, 7979–7990. [Google Scholar]
  25. Gupta, D.K.; Jha, A.V.; Appasani, B.; Srinivasulu, A.; Bizon, N.; Thounthong, P. Load frequency control using hybrid intelligent optimization technique for multi-source power systems. Energies 2021, 14, 1581. [Google Scholar] [CrossRef]
  26. Qu, Z.; Younis, W.; Wang, Y.; Georgievitch, P.M. A Multi-Source Power System’s Load Frequency Control Utilizing Particle Swarm Optimization. Energies 2024, 17, 517. [Google Scholar] [CrossRef]
  27. Babu, N.R.; Bhagat, S.K.; Saikia, L.C.; Chiranjeevi, T.; Devarapalli, R.; Márquez, F.P.G. A comprehensive review of recent strategies on automatic generation control/load frequency control in power systems. Arch. Comput. Methods Eng. 2023, 30, 543–572. [Google Scholar] [CrossRef]
  28. Naderipour, A.; Abdul-Malek, Z.; Davoodkhani, I.F.; Kamyab, H.; Ali, R.R. Load-frequency control in an islanded microgrid PV/WT/FC/ESS using an optimal self-tuning fractional-order fuzzy controller. Environ. Sci. Pollut. Res. 2023, 30, 71677–71688. [Google Scholar] [CrossRef]
  29. Hu, C.; Bi, L.; Piao, Z.; Wen, C.; Hou, L. Coordinative optimization control of microgrid based on model predictive control. Int. J. Ambient. Comput. Intell. 2018, 9, 57–75. [Google Scholar] [CrossRef]
  30. Bhattacharya, H.; Chattopadhyay, S.; Chattopadhyay, M.; Banerjee, A. Storage and bandwidth optimized reliable distributed data allocation algorithm. Int. J. Ambient. Comput. Intell. 2019, 10, 78–95. [Google Scholar] [CrossRef]
  31. Choudhary, R.; Rai, J.; Arya, Y. Automatic generation control for single area power system using GNA tuned PID controller. In Journal of Physics: Conference Series; IOP Publishing: Bristol, UK, 2020. [Google Scholar]
  32. Raj, T.D.; Kumar, C.; Kotsampopoulos, P.; Fayek, H.H. Load frequency control in two-area multi-source power system using bald eagle-sparrow search optimization tuned PID controller. Energies 2023, 16, 2014. [Google Scholar] [CrossRef]
  33. Ali, T.; Malik, S.A.; Daraz, A.; Adeel, M.; Aslam, S.; Herodotou, H. Load frequency control and automatic voltage regulation in four-area interconnected power systems using a gradient-based optimizer. Energies 2023, 16, 2086. [Google Scholar] [CrossRef]
  34. Debnath, M.K.; Agrawal, R.; Tripathy, S.R.; Choudhury, S. Artificial neural network tuned PID controller for LFC investigation including distributed generation. Int. J. Numer. Model. Electron. Netw. Devices 2020, 33, e2740. [Google Scholar] [CrossRef]
  35. Kumarakrishnan, V.; Vijayakumar, G.; Jagatheesan, K.; Boopathi, D.; Anand, B.; Kanendra Naidu, V. PSO optimum design-PID controller for frequency management of single area multi-source power generating system. In Contemporary Issues in Communication, Cloud and Big Data Analytics: Proceedings of CCB 2020; Springer: Berlin/Heidelberg, Germany, 2022. [Google Scholar]
  36. Shukla, S.; Gupta, A.; Pandey, R. Load Frequency Control for Multi-Area Power System Using PSO-Based Technique. In Artificial Intelligence Techniques in Power Systems Operations and Analysis; Auerbach Publications: Boca Raton, FL, USA, 2023; pp. 17–35. [Google Scholar]
  37. Dhanasekaran, B.; Kaliannan, J.; Baskaran, A.; Dey, N.; Tavares, J.M.R.S. Load frequency control assessment of a PSO-PID controller for a standalone multi-source power system. Technologies 2023, 11, 22. [Google Scholar] [CrossRef]
  38. Ojha, S.K.; Maddela, C.O. Optimization Techniques of Load Frequency Control for Renewable Integrated Two-Area Power System. In International Symposium on Sustainable Energy and Technological Advancements; Springer: Berlin/Heidelberg, Germany, 2023. [Google Scholar]
  39. Can, O.; Ayas, M.S. Gorilla troops optimization-based load frequency control in PV-thermal power system. Neural Comput. Appl. 2024, 36, 4179–4193. [Google Scholar] [CrossRef]
  40. Illias, H.; Nurjannah, N.; Mokhlis, H. Optimization of Automatic Generation Control Performance for Power System of Two-Area based on Genetic Algorithm-Particle Swarm Optimization. In Proceedings of the 2024 IEEE 4th International Conference in Power Engineering Applications (ICPEA), Penang Island, Malaysia, 4–5 September 2024. [Google Scholar]
  41. Feleke, S.; Satish, R.; Salkuti, S.R.; Abdelaziz, A.Y. Load Frequency Control in Two-Area Interconnected Systems Using DE-PID and PSO-PID. In Power Quality in Microgrids: Issues, Challenges and Mitigation Techniques; Springer: Berlin/Heidelberg, Germany, 2023; pp. 391–407. [Google Scholar]
  42. Ogar, V.N.; Hussain, S.; Gamage, K.A. Load frequency control using the particle swarm optimisation algorithm and pid controller for effective monitoring of transmission line. Energies 2023, 16, 5748. [Google Scholar] [CrossRef]
  43. Mohamed, E.A.; Aly, M.; Watanabe, M.J.M. New tilt fractional-order integral derivative with fractional filter (TFOIDFF) controller with artificial hummingbird optimizer for LFC in renewable energy power grids. Mathematics 2022, 10, 3006. [Google Scholar] [CrossRef]
  44. Franklin, R.V.R.; Fathima, A.P.J.S. Frequency regulation for state-space model-based renewables integrated to multi-area microgrid systems. Sustainability 2023, 15, 2552. [Google Scholar] [CrossRef]
  45. Katipoğlu, D.; Soylu, S. Design of Optimal FOPI Controller for Two-Area Time-Delayed Load Frequency Control System with Demand Response. J. Electr. Eng. Technol. 2024, 1–13. [Google Scholar] [CrossRef]
  46. Wang, Z.; Liu, Y.; Yang, Z.; Yang, W. Load frequency control of multi-region interconnected power systems with wind power and electric vehicles based on sliding mode control. Energies 2021, 14, 2288. [Google Scholar] [CrossRef]
  47. Eisa, S.A. Modeling dynamics and control of type-3 DFIG wind turbines: Stability, Q Droop function, control limits and extreme scenarios simulation. Electr. Power Syst. Res. 2019, 166, 29–42. [Google Scholar] [CrossRef]
  48. Eisa, S.A.; Wedeward, K.; Stone, W. Wind turbines control system: Nonlinear modeling, simulation, two and three time scale approximations, and data validation. Int. J. Dyn. Control. 2018, 6, 1776–1798. [Google Scholar] [CrossRef]
  49. Eisa, S.A. Nonlinear modeling, analysis and simulation of wind turbine control system with and without pitch control as in industry. In Advanced Control Optimization Paradigms for Wind Energy Systems; Springer: Berlin/Heidelberg, Germany, 2019; pp. 1–40. [Google Scholar]
  50. Sobhy, M.A.; Abdelaziz, A.Y.; Hasanien, H.M.; Ezzat, M. Marine predators algorithm for load frequency control of modern interconnected power systems including renewable energy sources and energy storage units. Ain Shams Eng. J. 2021, 12, 3843–3857. [Google Scholar] [CrossRef]
  51. Khooban, M.H.; Gheisarnejad, M. A novel deep reinforcement learning controller based type-II fuzzy system: Frequency regulation in microgrids. IEEE Trans. Emerg. Top. Comput. Intell. 2020, 5, 689–699. [Google Scholar] [CrossRef]
  52. Ginidi, A.; Ghoneim, S.M.; Elsayed, A.; El-Sehiemy, R.; Shaheen, A.; El-Fergany, A. Gorilla troops optimizer for electrically based single and double-diode models of solar photovoltaic systems. Sustainability 2021, 13, 9459. [Google Scholar] [CrossRef]
  53. Kouba, N.E.Y.; Menaa, M.; Hasni, M.; Boudour, M. LFC enhancement concerning large wind power integration using new optimised PID controller and RFBs. IET Gener. Transm. Distrib. 2016, 10, 4065–4077. [Google Scholar] [CrossRef]
  54. Zhao, W.; Wang, L.; Mirjalili, S. Artificial hummingbird algorithm: A new bio-inspired optimizer with its engineering applications. Comput. Methods Appl. Mech. Eng. 2022, 388, 114194. [Google Scholar] [CrossRef]
Figure 1. Block diagram of speed governor.
Figure 1. Block diagram of speed governor.
Energies 17 03962 g001
Figure 2. Transfer function model of turbine.
Figure 2. Transfer function model of turbine.
Energies 17 03962 g002
Figure 3. Block structure of load mathematical model.
Figure 3. Block structure of load mathematical model.
Energies 17 03962 g003
Figure 4. One-area thermal energy network representation.
Figure 4. One-area thermal energy network representation.
Energies 17 03962 g004
Figure 5. Block diagram showing mechanical part of wind turbine.
Figure 5. Block diagram showing mechanical part of wind turbine.
Energies 17 03962 g005
Figure 6. Block representation of PMSG wind energy network.
Figure 6. Block representation of PMSG wind energy network.
Energies 17 03962 g006
Figure 7. P&O MPPT approach.
Figure 7. P&O MPPT approach.
Energies 17 03962 g007
Figure 8. Circuit illustration of single-diode PV module.
Figure 8. Circuit illustration of single-diode PV module.
Energies 17 03962 g008
Figure 9. Circuit illustration of double-diode PV module.
Figure 9. Circuit illustration of double-diode PV module.
Energies 17 03962 g009
Figure 10. Schematic representation of fuzzy-based MPPT module.
Figure 10. Schematic representation of fuzzy-based MPPT module.
Energies 17 03962 g010
Figure 11. Fuzzy-based MPPT approach for PV.
Figure 11. Fuzzy-based MPPT approach for PV.
Energies 17 03962 g011
Figure 12. Block representation of dual-area IPS.
Figure 12. Block representation of dual-area IPS.
Energies 17 03962 g012
Figure 13. Illustration of single-area multiple-source energy network.
Figure 13. Illustration of single-area multiple-source energy network.
Energies 17 03962 g013
Figure 14. Fundamental illustration of PID controller.
Figure 14. Fundamental illustration of PID controller.
Energies 17 03962 g014
Figure 15. Flow chart of hybrid PSO-AHA optimizer.
Figure 15. Flow chart of hybrid PSO-AHA optimizer.
Energies 17 03962 g015
Figure 16. The frequency response of a single-area power network.
Figure 16. The frequency response of a single-area power network.
Energies 17 03962 g016
Figure 17. One-area multiple-source frequency response to case 1.
Figure 17. One-area multiple-source frequency response to case 1.
Energies 17 03962 g017
Figure 18. One-area multiple-source frequency response to case 2.
Figure 18. One-area multiple-source frequency response to case 2.
Energies 17 03962 g018
Figure 19. Transient frequency response of dual-area tie-line IPS.
Figure 19. Transient frequency response of dual-area tie-line IPS.
Energies 17 03962 g019
Figure 20. Frequency response of dual-area case 1.
Figure 20. Frequency response of dual-area case 1.
Energies 17 03962 g020
Figure 21. Frequency response of dual-area case 2.
Figure 21. Frequency response of dual-area case 2.
Energies 17 03962 g021
Figure 22. Frequency response of dual-area case 3.
Figure 22. Frequency response of dual-area case 3.
Energies 17 03962 g022
Figure 23. Frequency response of dual-area case 4.
Figure 23. Frequency response of dual-area case 4.
Energies 17 03962 g023
Figure 24. Frequency response of dual-area case 5.
Figure 24. Frequency response of dual-area case 5.
Energies 17 03962 g024
Figure 25. Frequency response of dual-area case 6.
Figure 25. Frequency response of dual-area case 6.
Energies 17 03962 g025
Figure 26. Frequency responses of dual-area case 7.
Figure 26. Frequency responses of dual-area case 7.
Energies 17 03962 g026
Figure 27. Frequency response of dual-area case 8.
Figure 27. Frequency response of dual-area case 8.
Energies 17 03962 g027
Figure 28. Frequency response of dual-area case 9.
Figure 28. Frequency response of dual-area case 9.
Energies 17 03962 g028
Figure 29. Frequency response of dual-area case 10.
Figure 29. Frequency response of dual-area case 10.
Energies 17 03962 g029
Figure 30. Frequency response of dual-area case 11.
Figure 30. Frequency response of dual-area case 11.
Energies 17 03962 g030
Figure 31. Frequency response of dual-area case 12.
Figure 31. Frequency response of dual-area case 12.
Energies 17 03962 g031
Figure 32. Frequency response of dual-area case 13.
Figure 32. Frequency response of dual-area case 13.
Energies 17 03962 g032
Figure 33. Frequency responses under load-variation case 1.
Figure 33. Frequency responses under load-variation case 1.
Energies 17 03962 g033
Figure 34. Frequency responses under load-variation case 2.
Figure 34. Frequency responses under load-variation case 2.
Energies 17 03962 g034
Figure 35. Frequency responses under load-variation case 3.
Figure 35. Frequency responses under load-variation case 3.
Energies 17 03962 g035
Figure 36. Frequency responses under load-variation case 4.
Figure 36. Frequency responses under load-variation case 4.
Energies 17 03962 g036
Table 1. Review of studied literature.
Table 1. Review of studied literature.
Optimization Approach/Secondary ControllerEnergy SourceWorkReference
PSO/PIDThermalContrasted results GA in addition to HC[11]
ARA/PIDNon-reheat thermalContrasted performance PSO, DE, JAYA[14]
GA/PIDThermalEfficiency PI as well as PID with and without GRC[31]
BESSO/PIDHydro, gas, thermalExamined controller performance along with contrasted finding with ordinary PID[32]
DE/PIDThermal, hydro, gasExamined controller performance using I, PI, as well as PID[19]
HBFOA/PIDThermal, PVPSO as well as BFO results investigated[21]
GBO/PIDGas, thermal, hydro, PV, windResults contrasted with GBO-I-PD, GBO-TID, and also GBO-I-P[33]
ANN/PIDDistributed power sources (WTG, DEG, AE, FC)Superiority of proposed technique achieved by employing GOA[34]
PSO/PIDPV, nuclear, hydro, gas, thermalContrast PSO-PID along with ordinary PID[24]
PSO/PIDThermal, solar, windEfficiency analysis of standard I, PI, in addition to PID[35]
PSO/PIDFThermalResults compared with PIDF and PI controller[36]
PSO/PIDThermal, hydro, gasResults compared with DE-PID and GA-PID[37]
PSO-GA-ACO-CSA/PIDPV and windPerformance of LFC evaluated under different cases by utilizing PSO, GA, ACO, and CSA[38]
GWO/PIDThermal, hydro, and nuclearResults compared with GA, PSO and ACO algorithms[13]
GTO/PIThermal and PVPerformance analyzed by changing the system parameters and load variations[39]
GA/MPCWind, solarResults compared with fuzzy-MPC, PSO-MPC, and conventional MPC[16]
GA-PSO/PIDThermalResults compared with PSO-PID and GA-PID[40]
PSO-DE/PIDThermal two-areaResults compared with PSO-PID and DE-PID[41]
PSO/PIDThermalBy varying load frequency, model system dynamics were tested[42]
AHA/PIDThermalResults compared with PSO-PID and AHA-PID[9]
AHA/TFOIDFFThermal, wind, solar, and EVResults compared tilt with filter and FOPID with filter[43]
AHA/3DOF-PIDWind, solar, and thermalResults compared AHA/3DOF-PID and existing controllers[44]
AHA/FOPIThermalResults compared AHA/FOPI, GWO/FOPI, and PSO/FOPI[45]
Table 2. Different power coefficient (Cp) numerical formulas.
Table 2. Different power coefficient (Cp) numerical formulas.
Coefficient
Type, Cp
Formula
Exponential 0.22 × 116 λ i 0.4 × β 5 × e 12.5 λ i avec 1 λ i = 1 λ + 0.08 × β 0.035 β 3 + 1
0.5 × 116 λ i 0.4 × β 5 × e 21 λ i + 0.068 × λ Avec : 1 λ i = 1 λ + 0.08 × β 0.035 β 3 + 1
0.5176 × 116 λ i 0.4 × β 5 × e 21 λ i + 0.0068 × λ Avec : 1 λ i = 1 λ + 0.08 × β 0.035 β 3 + 1
0.5109 × 116 λ i 0.4 × β 5 × e 21 λ i + 0.068 × λ Avec : 1 λ i = 1 λ + 0.08 × β 0.035 β 3 + 1
0.44 × 125 λ i 6.94 × e 16.5 λ i Avec : λ i = 1 1 λ + 0.002
Sinusoidal 0.5 0.167 × ( β 2 ) × sin π × ( λ 3 ) 18.9 0.3 × ( β 2 ) 0.00184 × ( λ 3 ) × ( β 2 )
0.3 0.00167 × ( β 2 ) × sin π × ( λ + 0.1 ) 10 0.3 × ( β 2 ) 0.00184 × ( λ 3 ) × β
Polynomial 6 × 10 7 × λ 5 + 10 5 × λ 4 65 × 10 5 × λ 3 + 2 × 10 5 × λ 2 + 76 × 10 3 × λ + 0.007
79.5633 × 10 5 × λ 5 17.375 × 10 4 × λ 4 + 9.86 × 10 3 × λ 3 9.4 × 10 3 × λ 2 + 6.38 × 10 2 × λ + 0.001
Table 3. One-area thermal power network constraint values.
Table 3. One-area thermal power network constraint values.
ConstraintConstraint Value
Turbine constant time0.5 s
Governor constant time0.2 s
Generator inertia constant5 s
Governor’s speed regulation0.05 pu
Turbine power rating250 MW
Frequency60 Hz
Duty ratio0.8
Table 4. Constraint values along with electrical features of wind, fuel cell, and solar.
Table 4. Constraint values along with electrical features of wind, fuel cell, and solar.
Characteristics of Advanced Renewable Energy AREi-230W-M6-G PV Module
Strings in parallel75
Module connected in series40
Maximal output power213.15 W
Voltage across open circuit36.3 V
Voltage at maximal power29 V
Temperature co-efficient across open circuit voltage−0.36099
Every module’s cell60
Short circuit current7.84 A
Current at maximal power7.35 A
Temperature co-efficient0.102
Stated temperature25 °C, 1000 (W/m2)
Diode Ideal factor0.98117
Resistance across shunt313.3991 ohm
Resistance across series0.39383 ohm
PMSG wind turbine features
Mechanized output power800 Kw
Generator’s base power800 Kw/0.9
Maximal power (Pu)0.90
Generator’s rotational speed (Pu)1.2
PMSG’s stator phase resistance0.0485 ohms
Wind speed12 m/s
Fuel cell constraint
Function 1: 1.36908036 atomic mass units60,000 × 8.3145 × (273 + 95) × 400 × u (1)/(2 × 96,485 × (3 × 101,325) × 0.919 × 0.995)
Function 2: 5.89420573 atomic mass units60,000 × 8.3145 × (273 + 95) × 400 × u (1)/(4 × 96,485 × (3 × 101,325) × 0.5057 × 0.21)
Stack power (Watts)50,000
Fuel cell resistance (ohms)0.66404
Nerst voltage of one cell (V)1.1342
Exchange current (A)0.91636
Exchange coefficient0.26402
Temperature across system (K)338
Fuel supply pressure (bar)1.5
Air supply pressure (bar)1
Nominal stack efficiency (%)55
Number of cells900
Operating temperature (Celsius)65
Nominal air flow rate (lpm)2100
Table 5. PV MPPT fuzzy rules.
Table 5. PV MPPT fuzzy rules.
ΔVPV × (o/p)ΔVPV(i/p)
ΔPPV(i/p) NBNSZEPSPB
NBPSPBNBNBNS
NSPSPSNSNSNS
ZEZEZEZEZEZE
PSNSNSPSPSPS
PBNSNBPBPBPS
Table 6. Fuzzy rule ranges, membership function (MF), and kinds.
Table 6. Fuzzy rule ranges, membership function (MF), and kinds.
Input/OutputNo. of MFMF RangeMF Type
Input 110−12.5 to 14.5Triangular
Input 210−4.6 to 2.3Triangular
Output10−7.2 to 3.8Triangular
Table 7. PSO and AHA algorithm operators for one-area multiple-source energy network.
Table 7. PSO and AHA algorithm operators for one-area multiple-source energy network.
ParametersAssigned Value
AHA
Fitness functionITAE
The population’s size70
Maximum iteration count200
No. of variables3
Dimension size100
Lower bounds0
Upper bounds1
PSO
Fitness functionITAE
The population’s size70
Maximum iteration count200
No. of variables3
Minimum weight of inertia0.4
Maximum inertia of weight0.9
The cognitive component1.85
The social component1.85
Random numbers countU (0, 1)
Table 8. PID gains for one-area multiple-source energy network.
Table 8. PID gains for one-area multiple-source energy network.
PID GainsKPKIKD
PSO-PID178.0278129.2391113.3715
PSO-AHA-PID103.967155.780427.4175
Table 9. Dual-area tie-line IPS constraint.
Table 9. Dual-area tie-line IPS constraint.
ConstraintArea 1Area 2
R0.050.0625
D0.60.9
H54
Base power1000 MVA1000 MVA
Gt0.2 s0.3 s
Tt0.5 s0.6 s
Table 10. PSO and AHA operators for dual-area tie-line IPS.
Table 10. PSO and AHA operators for dual-area tie-line IPS.
ParametersAssigned Value
AHA
Fitness functionITAE
The population’s size55
Maximum iteration count150
No. of variables9
Dimension size100
Lower bounds0
Upper bounds1
PSO
Fitness functionITAE
The population’s size55
Maximum iteration count150
No. of variables9
Minimum weight of inertia0.45
Maximum inertia of weight0.92
The cognitive component1.85
The social component1.85
Random numbers countU (0, 1)
Table 11. PID gains for dual-area tie-line IPS.
Table 11. PID gains for dual-area tie-line IPS.
ControllerArea 1Area 2Tie-Line
PIDPIDPID
PSO-AHA-PID0.497240.39757−0.487650.245810.908370−0.837450.495639.43 × 10−2−0.06856
PSO-PID0.758210.84530.41490.27183−0.32180.55273−0.83670.76910.51874
Table 12. Constraint variations in single area.
Table 12. Constraint variations in single area.
CasesConstraintInitial ValueAlteration RangeNew Value
Case 1Tg0.2+75%0.35
H5+75%8.75
D0.8−75%0.2
R0.05−75%0.0125
Case 2Tg0.2−75%0.05
H5−75%1.25
D0.8+75%1.4
R0.05+75%0.0875
Table 13. Frequency response one-area multiple-source signal values.
Table 13. Frequency response one-area multiple-source signal values.
ControllerFrequency Responds
Undershoot
%
Overshoot
%
Settling Time secFall Time
ms
ErrorPre-Shoot %High/Low
Hz
PSO-AHA-PID0.118−0.1180.616.9620.00060.50560/59.98
PSO-PID0.5051.5765.5327.3390.32360.24160/59.65
Case 1
PSO-AHA-PID0.34113.0680.917.8860.00080.56860/59.97
PSO-PID0.5050.0984.5414.6070.44570.19460/59.67
Case 2
PSO-AHA-PID0.246−0.2461.511.3290.00050.50560/59.96
PSO-PID6.9890.4646.6113.5860.31440.53860/59.50
Table 14. Investigated cases under system constraint variations.
Table 14. Investigated cases under system constraint variations.
Case NumberConstraintInitial ValuesAlteration RangeNew Values
Area 1Area 2Area 1Area 2
1H54+75%8.757
2H54−75%1.251
3Tt0.50.6+75%0.8751.05
4Tt0.50.6−75%0.1250.15
5B20.616.9+75%75.7521.175
6B20.616.9−75%5.154.225
7D0.60.9+75%1.051.575
8D0.60.9−75%0.150.225
9Tg0.20.3+75%0.350.525
10Tg0.20.3−75%0.050.075
11R0.050.0625+75%0.08750.1093
12R0.050.0625−75%0.01250.0156
13B20.616.9−75%5.154.225
H54+75%8.757
R0.050.0625+75%0.08750.1093
D0.60.9−75%0.150.225
Tt0.50.6+75%0.8751.05
Tg0.20.3+75%0.350.525
Table 15. Frequency response of dual-area signal characteristics.
Table 15. Frequency response of dual-area signal characteristics.
Frequency Response Area 1Frequency Response Area 2Tie-Line Power (Pu) Response
ControllerUsh%Osh%Psh%H/L
Hz
Ush%Osh%Psh%H/L
Hz
Ush%Osh%Psh%Fall time ms
PSO-AHA-PID1.9130.5050.50560/59.84−1.33911.049217.10860/59.991.1430.847−0.129208.088
PSO-PID2.0220.3130.50560/59.7834.47727.5640.64160/59.710.8020.505−0.2701196
Case 1
PSO-AHA-PID0.4272.5770.51560/59.881.4833.5490.96260/59.981.0300.769−0.262316.360
PSO-PID1.9980.3160.50560/59.8313.47211.7980.56260/59.78−0.05571.5520.8621038
Case 2
PSO-AHA-PID1.8340.5101.53160/59.8412.41131.945124.33660/59.991.2690.877−0.380190.147
PSO-PID23.570−0.0350.50560/59.7740.69430.9210.65860/59.691.65821.5993.512880.243
Case 3
PSO-AHA-PID1.9180.5050.50560/59.81−0.10610.2480.87760/59.971.4030.847−0.526274.872
PSO-PID27.17227.5640.64160/59.7430.73757.9370.79460/59.66−12.085148.893115.792329.824
Case 4
PSO-AHA-PID1.9190.59518.45260/59.91−53.88298.696302.21060/59.99−42.00684.552165.37921.547
PSO-PID1.9100.4770.50560/59.841.8360.5050.50560/59.780.51144.2030.725786.243
Case 5
PSOAHA-PID1.7071.5310.51060/59.86−8.84250.413348.72660/59.980.7770.8470.455191.507
PSO-PID17.7236.9890.53860/59.7833.30338.4066.52260/59.73−2.47038.0995.862805.789
Case 6
PSO-AHA-PID1.8460.4890.50560/59.736.41528.6100.80660/59.931.8160.9430.484295.817
PSO-PID0.5051.9850.50560/59.770.5051.3590.50560/59.701.6880.556−0.1808154
Case 7
PSO-AHA-PID1.6050.5050.50560/59.851.5196.964254.11160/59.991.0120.8470.516206.415
PSO-PID1.756−0.2290.50560/59.7931.42619.8800.60260/59.7278.03663.1150.8208578
Case 8
PSO-AHA-PID1.7270.5050.50560/59.8424.0213.3570.94360/59.971.2710.847−0.785209.712
PSO-PID19.340−1.0320.51060/59.786.74271.60979.21760/59.881.7900.562−0.4351063
Case 9
PSO-AHA-PID1.8990.5050.50560/59.824.71161.479110.90560/59.991.4830.877−0.652241.410
PSO-PID24.50610.5560.55660/59.7614.99872.39963.66960/59.861.5760.602−0.3481078
Case 10
PSO-AHA-PID1.5830.57514.36860/59.900.247214.7281.21160.1/601.0110.7350.628131.202
PSO-PID15.696.980.5360/59.610.501.960.5060/59.910.48380.9090.9091034
Case 11
PSO-AHA-PID1.65710.18575.0060/59.91246.560−246.560493.55060/59.991.1830.8200.588214.165
PSO-PID12.82724.3750.62560/59.7219.93544.2030.72560/59.650.47837.47628.826925.219
Case 12
PSO-AHA-PID1.8050.59518.45260/59.87−1.33325.0311333.1560/59.981.1600.8470.740208.194
PSO-PID16.2890.4760.50560/59.7731.02527.5640.64160/59.700.7190.521−0.3471194
Case 13
PSO-AHA-PID1.7450.4990.50560/59.690.8132.4870.86260/59.941.1630.8930.615325.216
PSO-PID0.5051.0950.50560/59.680.501.53238.0160/59.601.91685.000−0.1511447
Table 16. Load alteration for dual area.
Table 16. Load alteration for dual area.
Case No.Area 1Area 2
1250 MW increment200 MW increment
2100 MW decrement300 MW decrement
3400 MW increment150 MW decrement
4200 MW decrement350 MW increment
Table 17. Frequency response for dual-area signal characteristics in load alteration.
Table 17. Frequency response for dual-area signal characteristics in load alteration.
Area 1Area 2Tie-Line Power (pu)
ControllerUsh%Osh%Psh%H/L
Hz
Ush%Osh%Psh%H/L
Hz
Ush%Osh%Psh%Fall time ms
Case 1
PSO-AHA-PID1.7750.5050.50560/59.510.8128.610169.78660/59.981.7870.847−0.147208.101
PSO-PID18.9019.3410.54960/59.1724.6485.80351.75160/59.830.5051.6820.5051219
Case 2
PSO-AHA-PID−0.39891.3460.21360.10/60−62.720211.491198.83860/59.99−7.40168.6440.84726.065
PSO-PID1.5600.5950.37561.21/60−6.502184.362110.78760.08/601.92144.2030.661703.60
Case 3
PSO-AHA-PID1.7420.5050.50560/59.22−2.24915.178268.49460/59.991.1440.847−0.145208.033
PSO-PID1.2840.610−0.20560.60/60−13.481231.38990.46260/59.95−1.56350.7580.644703.60
Case 4
PSO-AHA-PID0.3980.5050.10960.39/60−61.603209.938100.93360/60−13.04168.6440.84726.061
PSO-PID18.50011.7980.56260/58.566.757−1.7130.50560/59.560.5050.8180.3901238
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

Younis, W.; Yameen, M.Z.; Tayab, A.; Qamar, H.G.M.; Ghith, E.; Tlija, M. Enhancing Load Frequency Control of Interconnected Power System Using Hybrid PSO-AHA Optimizer. Energies 2024, 17, 3962. https://doi.org/10.3390/en17163962

AMA Style

Younis W, Yameen MZ, Tayab A, Qamar HGM, Ghith E, Tlija M. Enhancing Load Frequency Control of Interconnected Power System Using Hybrid PSO-AHA Optimizer. Energies. 2024; 17(16):3962. https://doi.org/10.3390/en17163962

Chicago/Turabian Style

Younis, Waqar, Muhammad Zubair Yameen, Abu Tayab, Hafiz Ghulam Murtza Qamar, Ehab Ghith, and Mehdi Tlija. 2024. "Enhancing Load Frequency Control of Interconnected Power System Using Hybrid PSO-AHA Optimizer" Energies 17, no. 16: 3962. https://doi.org/10.3390/en17163962

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