Next Article in Journal
Dynamic Responses of Human Skin and Fascia to an Innovative Stimulation Device—Shear Wave Stimulation
Previous Article in Journal
Analysis of a Regression Model for Creating Surface Microgeometry after Machining Zirconia YML Used for Dental Application
Previous Article in Special Issue
A Survey on Biomimetic and Intelligent Algorithms with Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fractional-Order Boosted Hybrid Young’s Double-Slit Experimental Optimizer for Truss Topology Engineering Optimization

1
School of Art and Design, Xi’an University of Technology, Xi’an 710054, China
2
National Demonstration Center for Experimental Arts Education, Nankai University, Tianjin 300371, China
3
Department of Applied Mathematics, Xi’an University of Technology, Xi’an 710054, China
*
Author to whom correspondence should be addressed.
Biomimetics 2024, 9(8), 474; https://doi.org/10.3390/biomimetics9080474
Submission received: 25 June 2024 / Revised: 30 July 2024 / Accepted: 1 August 2024 / Published: 5 August 2024

Abstract

:
Inspired by classical experiments that uncovered the inherent properties of light waves, Young’s Double-Slit Experiment (YDSE) optimization algorithm represents a physics-driven meta-heuristic method. Its unique search mechanism and scalability have attracted much attention. However, when facing complex or high-dimensional problems, the YDSE optimizer, although striking a good balance between global and local searches, does not converge as fast as it should and is prone to fall into local optimums, thus limiting its application scope. A fractional-order boosted hybrid YDSE, called FYDSE, is proposed in this article. FYDSE employs a multi-strategy mechanism to jointly address the YDSE problems and enhance its ability to solve complex problems. First, a fractional-order strategy is introduced into the dark edge position update of FYDSE to ensure more efficient use of the search potential of a single neighborhood space while reducing the possibility of trapping in a local best. Second, piecewise chaotic mapping is constructed at the initial stage of the population to obtain better-distributed initial solutions and increase the convergence rate to the optimal position. Moreover, the low exploration space is extended by using a dynamic opposition strategy, which improves the probability of acquisition of a globally optimal solution. Finally, by introducing the vertical operator, FYDSE can better balance global exploration and local exploitation and explore new unknown areas. The numerical results show that FYDSE outperforms YDSE in 11 (91.6%) of cec2022 sets. In addition, FYDSE performs best in 8 (66.6%) among all algorithms. Compared with the 11 methods, FYDSE obtains the optimal best and average weights for the 20-bar, 24-bar, and 72-bar truss problems, which proves its efficient optimization capability for difficult optimization cases.

1. Introduction

Optimization problems are omnipresent across numerous fields within physics and engineering, and their importance cannot be overstated [1,2]. Especially in the fields of applied mechanics and engineering, finding the optimal or near-optimal solution becomes the key to enhancing the efficiency of the system, reducing the consumption of resources, and even determining the success or failure of the technology or product [3,4]. The development and application of optimization algorithms as mathematical tools play a crucial role in solving these problems [5,6]. With the rapid increase in computing power and the arrival of the significant data era, artificial intelligence technologies, such as deep learning, have developed rapidly in recent years. At the same time, they also bring new challenges and opportunities for the research and application of optimization algorithms. In deep learning, the gradient descent method and its variant algorithms, such as stochastic gradient descent (SGD), Adam, etc., have become mainstream methods in neural network training and optimization [7,8]. However, these algorithms often need to be improved to solve complex issues, such as nonconvex optimization and multimodal problems. As a class of stochastic optimization methods based on intuitive or empirical constructions, the Metaheuristic Algorithm (MHA) has shown significant advantages in solving practical optimization designs in a wide range of domains due to its lack of a strict gradient requirement, as well as its parallel search, fewer parameters, and strong search capability [9,10].
MHAs are a class of optimization tools designed for simulating natural phenomena or processes [11]. They can accurately seek optimal or near-optimal solutions in complex search spaces, an ability that is made possible by the simulation of complex behaviors, such as biological evolution, physical motions, and chemical reactions. These algorithms are capable of global search and are flexible in adapting their strategies to diverse optimization problems [12,13]. Depending on the simulation objects and mechanisms, MHAs can be categorized into four primary groups: evolution-based, physics-based, population-based, and human-based [14]. Inspired by the theory of biological evolution, evolution-based methods explore and optimize problem solution space by simulating biological evolutionary mechanisms, such as natural selection, heredity, and mutation. Among them, genetic algorithms (GA) [15] are the pioneers in this field, and they mimic the process of biological inheritance and mutation. In addition, differential evolutionary algorithms (DEs) [16], and Evolution Search (ES) [17] are well-known evolutionary algorithms. Physical phenomena and laws inspire physics-based MHAs. The Simulated Annealing Algorithm (SA) simulates the physical process of heating an object to melt and then its gradually cooling to optimize the problem solution [18]. Similarly, the Fick’s Law Algorithm (FLA) is based on Fick’s first law of diffusion [19], while the Gravitational Search Algorithm (GSA) simulates the law of gravity in the physical world [20]. On the other hand, population-based metaheuristic algorithms simulate the behavior or group intelligence of populations in nature [21]. They represent the solution of a problem as individuals in a population and find the optimal solution through interaction, competition, and cooperation among individuals. The Particle Swarm Algorithm (PSO) [22] is the earliest traditional algorithm, and it simulates the social behavior of animal groups. Other well-known algorithms include the Gray Wolf Optimization (GWO) [23], the Harris Hawk Optimization (HHO) [24], and the Black-Winged Kite Optimization Algorithm (BKA) [25]. Finally, human-based methods are mainly generated based on human behaviors, mindsets, and cognitive processes. The Political Optimization Algorithm (PO) [26], Human Evolutionary Optimization Algorithm (HEOA) [27], and Rich-Poor Optimization Algorithm (PRO) [28] are outstanding representatives in this field. These algorithms provide new ideas and methods for solving complex optimization problems by simulating human behaviors or cognitive processes.
YDSE is a method based on physical principles proposed in 2022 [29]. It has effectively addressed many realistic engineering problems due to its advantages of high flexibility, high convergence accuracy, and global search capability. Compared with some classical and new algorithms, YDSE is highly competitive and asymptotic when facing engineering optimization problems. However, YDSE cannot effectively balance the equilibrium between global and local searches, as well as having slower convergence to the optimal solution, and tends to trap in the local best, thus limiting its application scope. Some improved versions of YDSE also try to solve these problems and verify the effectiveness of the improvement through engineering optimization problems. For example, Hu et al. presented an enhanced form of YDSE. They tried to introduce four efficient mechanisms to enhance the optimization ability of YDSE and demonstrated better optimization performance in some difficult engineering optimizations [30]. A smart weighting tool for optimizing dissolved oxygen levels based on Young’s double-slit optimizer was presented by Dong et al. [31].
Although YDSE and some existing enhanced versions still demonstrate better convergence performance and results when facing engineering optimization problems, they still do not effectively balance the global and local searches and are especially prone to becoming bogged down in localized solutions when facing complex issues. This result leads to unsatisfactory optimization results obtained by the YDSE optimizer when solving difficult optimization issues. Therefore, this study proposes a fractional-order boosted hybrid YDSE optimizer to further alleviate the shortcomings of existing YDSE algorithms and attempts to solve realistic engineering optimization problems of higher complexity. In addition to the above motivations, bringing in operators and single strategies, while improving the manifestations of one aspect of the algorithm, can indirectly impair the performance of other aspects. Therefore, we jointly improve the YDSE problem by introducing multi-strategy mechanisms and minimizing this indirect damage. Second, many original and improved algorithms have been proposed that are performance-efficient and effective. However, based on the No Free Lunch (NFL) statement [32], it has been shown that existing methods are not capable of comprehensively addressing all conceivable optimization challenges. This scenario suggests that a certain category of optimization algorithms, while capable of yielding satisfactory outcomes for particular problems, may result in unacceptable results for others [33]. In addition, the fractional-order strategy is considered a strategy to efficiently address the balance between exploration and exploitation. It has been applied to improve a variety of algorithms and has demonstrated efficient experimental performance. For example, a JS algorithm based on fractional-order improvement has been proposed to address the inability of the JS algorithm to effectively balance between exploration and exploitation, and the effectiveness of the strategy has been verified through a variety of engineering optimization problems and real-world economic forecasting problems [34]. These reasons motivate this study to propose a fractional-order boosted hybrid YDSE optimizer. Solteiro Pires et al. proposed a convergence rate control method for Particle Swarm Optimization algorithms based on fractional-order calculus [35]. Luo et al. proposed a Jellyfish Search Algorithm (JSA) for the optimization and tuning of the control gains of the developed strategy to obtain a high-quality global optimum [36]. Zhang et al. proposed an improved algorithm of the Bird Flock Algorithm for solving parameter estimation of fractional-order chaotic control and synchronization [37]. These reasons motivate this study to propose a fractional-order boosted hybrid YDSE optimizer.
Therefore, this study proposes a fractional-order boosted hybrid YDSE optimizer. First, a fractional-order strategy is introduced into the dark edge position updating of FYDSE to ensure more effective exploitation of the search potential of a single neighborhood space while reducing the possibility of getting bogged down in localized solutions. Secondly, piecewise chaotic mapping is constructed in the generation of the initial solution to gain better distributed initial individuals and to enhance the speed of convergence to the most individuals. Moreover, the low exploration space of the algorithm is extended by using a dynamic inversion strategy, which improves the probability of obtaining a globally optimal solution. Finally, by introducing the vertical operator, FYDSE can better balance global exploration and local exploitation to explore new unknown areas. In addition, the presented FYDSE is compared with a series of the most efficient comparative algorithms in a comparative experiment on the CEC2022 test. The experimental outcomes and statistical analysis validate the efficient performance of FYDSE in addressing optimization issues. Further, to verify the efficient ability of the FYDSE in facing real-world engineering application problems, comparative experiments of FYDSE are conducted on three complex truss topology optimization problems. The superior adaptability and high efficiency of the FYDSE in complex engineering issues validate the ability of the FYDSE in dealing with similar applied mechanics and engineering problems.
The key contributions of this research are enumerated as follows:
(1)
In this article, a fractional-order boosted hybrid Young’s double-slit experimental optimization algorithm is proposed.
(2)
A multi-strategy mechanism is employed to jointly improve the YDSE problem to mitigate the negative impact of a single strategy on other aspects of the algorithm and to improve its ability to solve complex problems.
(3)
Experimental results of numerical and 20-bar, 24-bar, and 72-bar topology optimization validate the performance advantages of the proposed FYDSE approach.
The remainder of the article is arranged as described below. Section 2 describes the specific details of the YDSE optimizer. Section 3 describes a fractional-order boosted hybrid improved YDSE optimization method. In Section 4, the ability of FYDSE on the CEC2022 test suit is discussed. Section 5 describes the performance of YDSE on the 20-bar, 24-bar, and 72-bar topology problems. Finally, the conclusions of this research are introduced.

2. The Theory of YDSE

YDSE is an innovative physics-based meta-heuristic method presented in 2022 that provides high-quality solutions to numerical and engineering optimization problems based on the fluctuating properties of light [29]. YDSE has already demonstrated excellent performance in the original and its improved versions, effectively addressing real-world mechanical optimization issues. The workflow of YDSE consists of three core steps: The first part is to initialize the position of the population. In the second part, YDSE simulates the fluctuating behavior of light according to Huygens’ principle and guides the algorithm in finding the optimal solution by constantly updating the traveling waves and path differences. The third part is the exploration and development phase. In the exploration stage, YDSE attempts to discover new potential optimal solutions by expanding the search range. Meanwhile, in the exploitation phase, the current optimal solution is fine-tuned to obtain a more highly interpretable solution. The ability to effectively balance these two processes dramatically determines the performance of the algorithm [29].
YDSE initializes each individual by simulating the process of projecting a monochromatic light wave source through two closely spaced slits into the barrier. To simulate this process, an original light source composed of M waves is constructed in the solution space as follows [29]:
M i j = l b j + r a n d × ( u b i l b j ) ,   i = 1 ,   2 ,   ,   n ,   j = 1 ,   2 ,   ,   d
where Mi,j represents the ith wave in dimension j. In addition, n means the population size and d means the dimension of the issue. For each j dimension, ubj and lbj denote the maximum and minimum range of the issue. In addition, rand means a randomly generated value between 0 and 1.
After initialization, waves next pass through the slit barrier and diverge in many different directions. The resulting dispersion allows for constant changes in the position of the wave source and the wave center. The YDSE method does this by setting the points on the wavefronts passing through the two slits to be equal and set to n. In a simplified way, the wavefronts of n points at two slits can be calculated by the following two equations [29]:
F S i j = M i j + L e n g t h × r a n d 1 × ( S A v g M i j )
S S i j = M i j L e n g t h × r a n d 2 × ( S A v g M i j )
where FSij and SSij denote the jth dimension coordinates of the ith point on the wavefront flowing out of the first and second slits, respectively. To simulate the random scattering phenomenon that may occur after a light wave passes through the slit, we introduce two random variables, rand1 and rand2, both of which take values in the range of [−1, 1]. Length means the length from the light source to the barrier. SAvg means the mean value of the current light wave group, which is calculated as follows [29]:
S A v g = 1 n i = 1 n M i , :
where Mi,: means the ith original light source.
Points on the wavesurface propagating from the two narrow gaps produce interference patterns. Specifically, constructive and damaging interference generate light and dark streaks. Therefore, the position of each individual is updated by constructed wave fronts (FS and SS) to realize the interference behavior and path differences between light and dark streaks. The specific equations are as follows [29]:
X i , : = ( F S i , : + S S i , : 2 ) + Δ P
where ∆P means the path length from the FS point to the SS point. In addition, FSi,: and SSi,: mean the ith FS point to SS point. The introduction of ∆P is utilized to distinguish different interference orders. Specifically, for zero-order or even-order interference, bright fringes are produced in the interference region. Meanwhile, for odd-order interference, dark fringes are produced. The specific equations are given below [29]:
Δ P = { 0 ,   i f   b r i g h t   a n d   m = 2   λ 2 ( 2 m + 1 ) ,   i f   d a r k   a n d   o d d   m       n λ ,   i f   b r i g h t   a n d   e v e n   m
where m means the order of interference. In addition, λ denotes wavelength.
The bright fringe emerges due to the combination of two waves, creating a wave of greater amplitude. The amplitude of this bright fringe is then progressively updated as follows [29]:
A b ( t ) = 2 1 + 1 ( t × cosh ( π / t ) / T max ) 2
where Ab means the mean amplitude of bright fringes. In addition, cosh(·) means the cosh function. t represents the now iteration and Tmax means the maximum iterations.
If the peak of one wave meets the trough of the other when two waves are added together, they both cancel out and produce dark fringes. Similarly, the amplitude of the dark fringe is iteratively updated as follows [29]:
A d ( t ) = ϕ × tanh 1 ( t T max + 1 )
where φ represents a constant with a value of 0.38. In addition, tanh−1(·) means the inverse of the tanh function.
Both bright and dark fringes represent potential candidate solutions when exploring solutions. In particular, the center fringe location is considered to be the optimal solution. Due to the effect of destructive interference, the fitness values of the dark fringe candidate solutions are lower than those of the bright fringe candidate solutions. This interference causes the positions in the dark fringe to be inferior to objective function than those in the light fringe. Therefore, YDSE emphasizes the priority of exploring for hopeful solutions in the dark zone during the exploration phase. A dark edge update is introduced to guide the search process and ensure that the solution space can be explored more efficiently. The specific update equation is as follows [29]:
X N o d d ( t + 1 ) = X N o d d ( t ) ( r a n d × A d ( t + 1 ) × L i N o d d ( t + 1 ) × X N o d d ( t ) Z × X b e s t ( t ) )
where XN-odd means the dark fringe, and rand denotes a randomly selected number between 0 and 1. LiN-odd denotes the dark edge intensity, which is utilized to reflect the brightness change of the dark fringe and Xbest denotes the best solution. In addition, Z means a trial vector of d, as follows [29]:
Z = t 2 × r a n d 1 h
where h means a randomly created number in the range of [−1, 1]. In addition, the definition of LiN-odd is as follows [29]:
L i N o d d ( t + 1 ) = L i max ( t + 1 ) × cos 2 ( π d λ L y d ( t + 1 ) )
where yd is employed to measure the length from the center stripe to the dark stripe, as follows [29]:
y d = λ L d ( m + 1 2 )
In addition, Limax denotes the maximum center stripe intensity, as follows [29]:
L i max ( t + 1 ) = C × t T max
where C is taken as 10−20, which represents the maximum intensity.
When the order is even, YDSE pays special attention to the bright edge regions and finds promising candidate solutions in them. It assumes that these regions are where the best solutions are potentially located, which, in turn, generates constructive interference in favor of the solution process. Therefore, during the development phase, the YDSE optimizer spares no effort in exploring all possible and promising candidate solution regions in the bright edge region. Its position update equation is concretely given as below [29]:
X N e v e n ( t + 1 ) = X N e v e n ( t ) ( ( 1 g ) × A b ( t + 1 ) × L i N e v e n ( t + 1 ) × X N e v e n ( t ) + g × Y )
where XN-even means the bright fringe. In addition, Y denotes the difference between two randomly acquired stripes and LiN-even denotes the intensity of the bright stripe. g means a randomly created number in the range of [0, 1]. The definition of Y is as follows [29]:
Y = X N r 1 ( t ) X N r 2 ( t )
where XN−r1 and XN−r2 represent two random fringes. In addition, the definition of LiN-even is as follows [29]:
L i N e v e n ( t + 1 ) = L i max ( t + 1 ) × cos 2 ( π d λ L y b ( t + 1 ) )
where Limax is defined through Equation (13).
Finally, the specific formula for the position update method of the center fringe is as follows [29]:
X N z e r o ( t + 1 ) = X b e s t ( t ) + ( A b ( t + 1 ) × L i max ( t + 1 ) × X N z e r o ( t ) r a n d × Z × X r b ( t ) )
where XN-zero means the central fringe and rand means a random number between 0 and 1. In addition, Xrb represents the bright fringe selected based on a random even number of rb.

3. Proposed FYDSE Algorithm

The YDSE method suffers from slow convergence, is easily caught up in localized solutions, and is unable to effectively balance exploration and exploitation when facing complex optimization problems in engineering. In this work, a fractional-order improved optimization method for a multi-strategy YDSE is proposed. The specific modifications of the proposed FYDSE are shown below.

3.1. Fractional-Order Modified Mechanism

Fractional calculus elevates the fundamental order of calculus by extending its realm from integers to fractions. It systematically derives the solution limit by employing differential approximations of integer calculus, specifically through the differentiation and integration of fractional orders [34,35]. The introduction of the fractional-order strategy in the population position updating process of FYDSE, which extends X(t) to fractions through the fractional-order formula, guarantees a more efficient exploration of the potential within a single neighborhood space, thereby minimizing the likelihood of falling into a local optimum. The definition of fractional-order discretization based on the Grunwald–Letnikov (G-L) method [38] is summarized as follows [34]:
D β [ ω ( t ) ] = 1 T β k = 0 r ( 1 ) k Γ ( β + 1 ) ω ( t k T ) Γ ( k + 1 ) Γ ( β k + 1 )
where β is the fractional element in the common signal ω(t), Γ denotes the gamma function, and T is the truncation term. r stands for the order.
The results of the fractional derivative are closely related to the value of the current term and the previous state, with the effect of past events diminishing over time. Then, the above equation is modified as follows [34]:
D β [ ω ( t + 1 ) ] ω ( t + 1 ) β ω ( t ) + 1 2 β ( β 1 ) ω ( t 1 ) 1 6 β ( β 1 ) ( β 2 ) ω ( t 2 )     + 1 24 β ( β 1 ) ( β 2 ) ( β 3 ) ω ( t 3 )
where β is set to 4 [34] in the specific implementation.
During the exploration phase, given YDSE’s emphasis on identifying promising solutions in the dark region, a fractional-order correction strategy was implemented to enhance the updating formula for dark streaks, thus refining our search approach. Figure 1 shows the effect of fractional-order correction on the update. The dark fringe update formula based on fractional-order correction is as follows:
X N o d d ( t + 1 ) = β X N o d d ( t ) 1 2 β ( β 1 ) X N o d d ( t 1 ) 1 6 β ( β 1 ) ( β 2 ) X N o d d ( t 2 ) + 1 24 β ( β 1 ) ( β 2 ) ( β 3 ) X N o d d ( t 3 ) ( r a n d × A d ( t + 1 ) × L i N o d d ( t + 1 ) × X N o d d ( t ) Z × X b e s t ( t ) )
Notably, upon multiplying the terms of 1/120, 1/720, or greater by the remaining components of the equation, their respective results become insignificant, scarcely influencing location updates. Consequently, these higher-order terms are disregarded.

3.2. Piecewise Chaotic Map Strategy

YDSE guides the updating of candidate solutions by iterative a priori information. Therefore, a better initial value captures valid information faster and thus searches for the optimal solution location faster [39]. The initialization of the established methods is through Gaussian distribution in order to generate randomly. However, this initialization is highly contingent, and there are more invalid solutions, thus limiting the effective updating of the solution. To obtain a better distribution of the initial solution space, FYDSE introduces a piecewise chaotic mapping distribution to initialize the solution space [40]. The population initialization in the piecewise chaotic mapping is achieved through chaotic variables rather than relying on random variables, thereby facilitating a full and efficient exploration of the solutions space [41]. The updated formula for the piecewise chaotic mapping is as follows [41]:
π ( k + 1 , d ) = { π ( k ) d , π ( k ) [ 0 , d ) π ( k ) d 0.5 d , π ( k ) [ d , 0.5 ) 1 π ( k ) d 0.5 d , π ( k ) [ 0.5 , 1 d ) 1 π ( k ) d , π ( k ) [ 1 d , 1 ) π ( k ) [ 0 ,   1 ]
where d means a random value in the scope of (0, 0.5).  π ( k + 1 , d )  means the piecewise chaotic mapping function. Therefore, the initialization method is updated as follows:
M i j = l b j + π ( k , d ) × ( u b i l b j ) ,   i = 1 ,   2 ,   ,   n ,   j = 1 ,   2 ,   ,   d

3.3. Dynamic Opposition (DO)

When solving complex optimization problems, YDSE often fails to efficiently traverse the search region of low exploration, which results in the algorithm failing to capture the global optimal solution. Therefore, a dynamic opposition (DO) strategy is introduced in the YDSE optimizer [42]. It extends the search space of low exploration by using a dynamic inversion strategy, thereby enhancing the likelihood of getting close to the global solution.
Considering that frequent DO strategy updates can affect the development performance of the algorithm, a jump rate (Jr) determines DO policy in the current iteration or not. If the number created at random is less than Jr, the DO implementation strategy is applied to all individuals. Consequently, Jr governs the likelihood of executing DO [43]. This approach, the Randomized Jump Strategy (RJS), facilitates the balanced exploration and development of methods in an effective manner through iterative selection. Furthermore, implementing DO during the selection iteration leads to an enhanced utilization of the search space, thus maximizing its efficiency. The DO strategy first requires generating the opposite solution in the neighborhood of the candidate solution X as below [42]:
X ¯ i , j = l b j + u b j X i , j ,   j = 1 ,   2 ,   ,   d
where  X ¯ i , j  represents the inverse solution.
The position of the candidate solution undergoes a shift due to the influence of a randomly generated number within the range of [0, 1], thereby yielding a new position designated as Xrand [42]:
X r a n d = r a n d ( X ¯ ) ,   r a n d [ 0 ,   1 ]
where rand means a random number.
When the candidate solution X is different from the moving direction of Xrand guided by the random number, a new candidate position based on the DO policy is constructed, as follows [43]:
X ¯ d o = X + r a n d ( X r a n d X )
where  X ¯ d o  represents the new position after dynamic opposition update.
To further illustrate the specific implementation of the DO strategy, Algorithm 1 provides pseudo-code for the DO strategy.
Algorithm 1: Dynamic Opposition Strategy
1: input: Jumping rate Jr, solution size N, candidate solution location X.
2:    if rand < Jr
3:        for i = 1 to N do
4:               X ¯ i , j = l b j + u b j X i , j ,   j = 1 ,   2 ,   ,   d
5:               X r a n d = r a n d ( X ¯ ) ,   r a n d [ 0 ,   1 ]
6:               X ¯ d o = X + r a n d ( X r and X )
7:          end for
8:    end if
9: output: New location on dynamic dichotomy.

3.4. Vertical Crossover Operator

To address the tendency of the YDSE algorithm to converge prematurely into local optimal solutions while tackling intricate problems, we develop an update strategy based on the vertical operation mechanism [44,45]. Specifically, the vertical operator performs an arithmetic crossover operation between two dimensions of the candidate solutions to generate new candidate individuals. By integrating the vertical operator, the FYDSE method gains enhanced global exploration and local exploitation abilities, allowing it to delve into novel and unexplored domains [46]. Furthermore, the adoption of the updated foraging strategy mitigates the issue of stagnant updates across iterations over many times. Assuming that the vertical crossover operation is performed with the d1 and d2 dimensions of the ith candidate solution, the generated conditioning solution  V i d 1 v c  is as follows [45]:
V i d 1 v c = c X i d 1 + ( 1 c ) X i d 2 ,   i M ( 1 , n ) ,   d 1 , d 2 M ( 1 , d ) ,
where c means uniformly distributed random values ranging from 0 to 1, n means the quantity of agents, and d signifies the dimensionality of the population. In addition, Xid1 and Xid2 denote the d1st and d2nd dimensions of the ith position. M(1, n) represents a function that selects individuals from all populations.
The vertical operator strategy is normalized for individuals, considering the upper and lower bounds for decision variables. The vertical operation of each is tailored specifically to a single individual, ensuring a focused search. This single-individual approach prevents the disruption of potentially globally optimal dimensions by abruptly departing from a locally optimal yet stagnant dimension. Furthermore, introducing a competition operator fosters healthy competition between the new and original candidate solutions. Specifically, the competition operator works as follows [46]:
X new = { V v c , i f   f ( V v c ) < f ( X ) , X ,   i f   f ( V v c ) > f ( X ) ,
where Vvc represents the candidate solution after the vertical operator operation. Xnew means new position. f(·) represents the fitness function.

3.5. The Proposed Enhanced FYDSE Meticulous Steps

Given the vast search space encountered in engineering optimization problems, attaining global convergence for intricate optimization challenges poses a significant challenge. To address this, this paper introduces a fractional-order improved boosted hybrid YDSE, explicitly tailored for solving complex optimization issues. First, a fractional-order strategy is introduced into the dark fringe position updating of FYDSE to ensure a more elegant search potential in a single neighborhood space, and also to minimize the risk of drifting into local optimality. Second, during the initialization phase, a piecewise chaotic mapping is introduced to yield a superior-quality initial population, thereby enhancing the convergence speed of the algorithm. Moreover, the lower exploration space for algorithms is extended by using a dynamic inversion strategy, which improves the probability of obtaining a globally optimal solution. Finally, by introducing the vertical operator, FYDSE can enhance equilibrium global exploration and local exploitation and explore new unknown areas.
The primary processes of FYDSE are described below:
Step 1: Initialization of the relevant elements of the FYDSE: the stripe quantity n, the dimension d, the lower and upper limits of the variable lb and ub, the maximum iteration number Tmax, the fraction coefficient β and the jump rate Jr;
Step 2: Initialize wavefronts by piecewise chaotic mapping. Obtain the position of each fringe based on the wavefronts and path difference ∆L and compute the objective function value.
Step 3: Update Limax by utilizing Equation (13) and pick the optimal location Xbest.
Step 4: When m is 0, the central stripe amplitude is updated using Equation (7) and the central stripe position is also updated according to the amplitude and the best position Xbest by Equation (17).
Step 5: When m is odd, the dark streak intensity and amplitude are updated using Equations (7) and (11). Afterwards, a new location is obtained by the dark streak update formula based on the fractional-order correction.
Step 6: When m is even, Equations (8) and (16) are utilized to update the amplitude and intensity of the bright streaks and further update the location of the bright streaks.
Step 7: Examine the lower and upper bounds of the variables and evaluate the objective function of the updated solution.
Step 8: If rand < Jr, update the dynamic opposite solution  X ¯ d o  for all the stripes by Equations (23)–(25) and compute the fitness value.
Step 9: A vertical operator is utilized to intersect the two dimensions of a randomly selected location and the availability of that new candidate solution is determined by a competitive operator.
Step 10: Consider the limitations of the assessment variables, both the upper and lower boundaries, and evaluate the solution of the objective function accordingly. Output the optimal fringe position Xbest.
The figure depicted in Figure 2 outlines the proposed FYDSE flowchart; furthermore, Algorithm 2 provides the corresponding pseudo-code for a clearer understanding of its implementation.
Algorithm 2: Fractional-order boosted hybrid Young’s double-slit experiment optimizer
1: Input: Initialize the related parameters of the proposed FYDSE algorithm: number of fringes (n), upper and lower bounds (lb and ub), dimension (d), maximum number of iterations (Tmax), the fraction coefficient β and jump rate Jr;
2: Initialize a monochromatic light source consisting of n waves using Equation (22) by piecewise chaotic mapping.
3: Calculate the wavefronts (FS and SS) of the two slits.
4: Obtain the position of each fringe based on wavefronts and path difference ∆L.
5: Examine the upper and lower bounds of the variables and evaluate the objective function of the updated solution.
6: While t < TMax do
7:    Update Limax by utilizing Equation (13) and pick the optimal location Xbest
8:    For i =1 to n do
9:          Update Z with Equation (10).
10:          If m = 0 (Central fringe)
11:              The intensity and amplitude of the central fringe are updated with Equations (17) and (7).
12:              Update the position of the center stripe Xm-zero based on the amplitude and the optimal position Xbest.
13:          Else if m = even (bright fringe)
14:              The intensity and amplitude of the bright fringe are updated with Equations (14) and (7).
15:              Update the bright fringe Xm-even by Equation (14).
16:          Else if m = odd (dark fringe)
17:              The intensity and amplitude of the dark fringe are updated with Equations (9) and (8).
18:              Update the new position Xm-odd of dark fringe based on the fractional-order modification with Equation (20).
19:          End if
20:          Check the bounds of the variables.
21:    End for
22:    If rand < Jr
23:        For i = 1 to n do
24:                 X ¯ i , j = l b j + u b j X i , j ,   j = 1 ,   2 ,   ,   d
25:                 X r a n d = r a n d ( X ¯ ) ,   r a n d [ 0 ,   1 ]
26:                 X ¯ d o = X + r a n d ( X r a n d X )
27:          End for
28:    End if
29:    B = permutate(d).
30:    For i = 1 to d/2 do
31:        Update a uniformly random value p.
32:        If p< P then let no1 = B(2i−1), and no2 = B(2i).
33:            For j = 1 to n do
34:                  Construct a random value c in the range of [0, 1].
35:                  Update the position of a stripe according to the vertical operator.
36:            End for
37:        End if
38:    End for
31:    Update the best position Xbest.
32:    t = t + 1.
33: End while
34: Output: the best solution Xbest.

3.6. The Computational Complexity of FYDSE

An important metric for evaluating optimization algorithms is the time complexity. The original YDSE has a complexity of O(Tmax×M×n+Tmax×n×d), where M represents the number of evaluations of the objective function. The proposed FYDSE includes four effective strategies. First, the segmented mapping initialization has the same complexity of O(n×d) as the Gaussian-based initialization of YDSE. Second, the fractional-order improved dark streak updating strategy has no added time complexity. Since the fractional-order improved strategy needs to save the experimental results of the first β iterations, the space complexity increases by O(β×n×d). Second, the DO strategy produces dynamic inverse solutions for all candidate solutions. However, given the limitation imposed by rand < Jr, the computational complexity of the DO strategy in the crap scenario of fulfilling the equation stands at O(Tmax×n×d). Finally, the complexity of the vertical operator is O(Tmax×n/2×d). Thus, the overall complexity of FYDSE is:
O ( F Y D S E ) = O ( T max × M × n + T max × n × d ) + O ( n × d ) + O ( β × n × d )     + O ( T max × n × d ) + O ( T max × n / 2 × d )     = O ( T max × M × n + T max × 5 n / 2 × d )

4. Numerical Simulations

To comprehensively measure the capabilities of FYDSE, by contrast to some other SOTA methods, in this section a comparative experiment on the CEC2022 set of tests is described. All the methods are set with the same parameters in the experiments, i.e., the maximum iteration number is 1000, and the population size is 30. We chose the mean, worst solution, optimal solution, and standard deviation as the evaluation indexes to visualize the experimental results. The experimental results are presented and analyzed in various forms, such as iterative plots, box plots, and radar plots. All tests were conducted on a personal computer with Matlab-2019b with a 2.11 GHz quad-core Intel(R) Core(TM) i5 and 8.00 GB.

4.1. Experimental Parameter Setting and Benchmark Functions

To deeply verify the effectiveness and superiority of our proposed FYDSE algorithm, we identified and selected a set of advanced and distinctive optimization algorithms for a comprehensive analysis and comparison with the experimental findings of FYDSE as follows: (1) YDSE [29]; (2) Artificial Hummingbird Algorithm (AHA) [47]; (3) Snake Optimization (SO) [48]; (4) Dandelion Optimizer (DO) [49]; (5) Aquila Optimizer (AO) [50]; (6) Chimpanzee Optimization Algorithm (ChOA) [51]; (7) Honey Badger Algorithm (HBA) [52]; (8) Sand Cat Swarm Optimization (SCSO) [53]; (9) Salp Swarm Algorithm (SSA) [54]; (10) White Shark Optimizer (WSO) [55]; (11) COOT Algorithm (COOT) [56]. By comparing with these algorithms, we aim to comprehensively demonstrate the advantages and features of the FYDSE algorithm in terms of performance, efficiency, and adaptability, thus further proving its capability to solve sophisticated optimization issues. Table 1 provides the parameters for the various algorithms.
The CEC2022 test suite contains 12 well-designed single-objective test functions used as benchmark functions for this experiment. These functions cover a wide range of types, such as single-peak, basis, hybrid, and combined functions, which are distinctive in complexity and morphology. Some functions exhibit smooth surface features, while others are full of sharp peaks and steep valleys. Among them, for the cec01 function, as a representation of a single-peaked function, only one globally optimal solution can measure the convergence speed and accuracy within the algorithm. On the other hand, the cec02-cec05 functions represent multimodal functions that possess numerous local optima, among which a single global optimum is designed to test whether the algorithm can successfully avoid traps of local minima and thus find the globally optimal solution. In addition, the hybrid functions cec06-cec08 are used to model the properties of real-world complex problems and allow for the evaluation of the algorithm’s effectiveness and ability to solve problems with hybrid properties. The CEC2022 test suite also introduces several combinatorial functions, such as the cec09-cec12 functions, which incorporate characteristics from a diverse set of optimization problems, presenting more intricate choices for the algorithms to overcome. In addition, the suite contains several constrained optimization problems. These problems search for optimal solutions under specific constraints and require the algorithms to adhere to these constraints strictly during optimization. The value domains of all test functions are set between [−100, 100].

4.2. Analysis of Exploration and Exploitation Behaviors

In the conventional YDSE framework, the refinement of dark fringes emphasized probing unknown areas to foster diversity. In contrast, the center and bright fringe prioritize cultivating optimal solutions within promising intervals. On the other hand, FYDSE improves the dark streak updating process through a fractional-order correction strategy, so that the positional updating of dark fringes focuses more on the search potential of a single neighborhood space, which mitigates the risk of converging to a local optimum and aids in discovering an optimal solution. Meanwhile, performing dynamic opposition for all fringe positions facilitates the exploration of the inverse positions of candidate solutions in the domain of the search, increasing the possibilities for searching undeveloped areas. Furthermore, incorporating the vertical operator enhances the exploration of potentially rewarding regions by augmenting the exploration of stripe positions within the dimensional intersection.
To ascertain the efficacy of the implemented strategy in balancing the exploration-exploitation equilibrium, we delve into the group diversity of FYDSE in the context of CEC2022. Population diversity infers group traits within a condensed environment through variations among individual dimensions [57]. Specifically, an increase in the difference between dimensions of the stripe location determines the dispersal or distribution in the search area of the population. Conversely, individuals will cluster into a convergence region. The introduction of the vertical operator improves the difference between the dimensions of the stripe positions.
First, we define the diversity (Divj) of the jth dimension and further compute the average diversity (Div) of the dimensions. Further, the proportion of exploration and exploitation is calculated based on the average diversity. The specific formula is as follows [57]:
D i v j = 1 n i = 1 n M e d i a n ( X : , j ) X i , j
D i v = 1 d j = 1 d D i v j
where n represents the number of all light and dark stripes and d represents the dimension. The median of the jth dimension across all (n) stripes is denoted as Median(X:,j).
Meanwhile, the definitions of the exploration rate Explora(%) and the exploitation rate Exploita(%) are given [57]:
E x p l o r a   ( % ) = D i v max ( D i v ) × 100
E x p l o i t a   ( % ) = | D i v max ( D i v ) | max ( D i v ) × 100
where max(·) denotes the max function.
This experiment validates the diversity results of FYDSE at the CEC2022 test function with 1000 iterations of FYDSE. Figure 3 shows the iterative plots of the exploration and exploitation rates of FYDSE at CEC2022 for all 12 test functions. From the analysis of the graphs, it can be concluded that for cec02, cec05, and cec07, FYDSE exhibits a high rate of exploration during extended iterations, while maintaining a high utilization rate during the middle and later stages of iterations. This result is attributed to the fact that the piecewise mapping initialization and dynamic opposition strategies enhance the search potential of YDSE. Meanwhile, the fractional-order modification strategy enhances the exploration performance by improving the update of dark streaks and effectively utilizing the existing prior knowledge. When dealing with cec01, cec03, cec09, and cec11, most iterations tend to be in a more intensive exploitation phase, with only a brief exploration phase in the initial stages. This is primarily due to the vertical operator enhancement of the exploitation stage. Additionally, the lower exploration rate during shorter iteration times indicates that the modified strategy aids the algorithm in swiftly locating the best solution.

4.3. Results Analysis with the Latest Methods in CEC2022

Table 2 provides the experimental outcomes of the FYDSE with the other latest comparison algorithms on the CEC2022 test dataset. To comprehensively evaluate the performance of these algorithms, we have chosen five key evaluation metrics, including the mean, the best value, the worst value, the standard deviation, and the ranking, to consider these algorithms comprehensively. These metrics provide us with a multi-dimensional view of the algorithm performance, which helps us to understand more clearly the performance differences and advantages and disadvantages between the FYDSE algorithm and other algorithms. As is evident from Table 2, the FYDSE algorithm exhibits the superior average performance among the eight test functions (cec02, cec06, cec07, cec08, cec09, cec10, cec11, cec12) under consideration, significantly outperforming the other algorithms. Not only that, but FYDSE also demonstrates remarkable competitiveness in other test functions. Comparatively, the original algorithm YDSE provides the best average results only on the specific test function cec05. The HBA algorithm, on the other hand, achieves the best average performance on two test functions, cec01 and cec03. In addition, the WSO algorithm gives the best average results for the cec04 test function. And for the cec01 function, both the HBA and SSA algorithms find the optimal solution. When exploring the cec01 single-peak function, both the HBA and SSA algorithms can converge to the global optimal solution. Meanwhile, our proposed FYDSE algorithm can also reach the sub-optimal solution level, further verifying that the FYDSE algorithm is not only fast-converging but also highly accurate. When confronted with these multimodal functions from cec02 to cec05, FYDSE achieves the best mean result in the cec02 function. Although the best mean value is not obtained on several other functions, the test results of FYDSE show the most minor standard deviation, demonstrating the stable exploration and development capability of FYDSE when facing multimodal functions. It is especially worth mentioning that the FYDSE algorithm can also find the global optimal solution when solving combinatorial and hybrid function problems. This achievement highlights the global solid search capability of FYDSE when solving complex, multivariate optimization problems and demonstrates its capability to achieve a balance in exploration and exploitation. In particular, when handling higher-dimensional, multimodal optimization tasks, the outstanding performance of FYDSE allows it to address the intricacies of real-world optimization problems with more excellent proficiency.
Table 3 provides a comprehensive overview of the statistical outcomes obtained from the Wilcoxon rank sum test conducted on FYDSE and alternative algorithms for comparison, with a predefined significance level of 0.05. Here, we describe the performance difference between the algorithms based on the size of the p-value. Specifically, when the p-value is less than 0.05, the difference between FYDSE and the comparison algorithm is significant. In contrast, when the p-value is greater than or equal to 0.05, it implies that the difference in performance between the two is insignificant. In addition, the symbols “-”, “=”, and “+” in Table 3 provide us with intuitive performance comparison information. Among them, “-” indicates that other meta-heuristic algorithms are not as practical as FYDSE in the corresponding test function. “=” indicates that FYDSE and the comparison algorithm have comparable performance in this test function, and both of them have the same effect. “+” indicates that the other algorithms obtained better results than FYDSE.
Upon examining Table 3, we can discern the distribution of the outcomes for all algorithms in the Wilcoxon test. Specifically, the results of the AHA, SO, HBA, and SSA algorithms are each 0/1/11. The statistical results imply that FYDSE does not have statistically worse results than AHA and SO in this set of tests and demonstrates better yield outcomes in a majority of the test functions. On the other hand, the results of the DO, SCSO, AO, and ChOA algorithms are 0/0/12. The statistical results imply that these algorithms neither show better performance than FYDSE nor comparable performance to FYDSE in the tests, but instead show worse performance 12 times. In addition, the Wilcoxon test results for both WSO and COOT are 0/2/10, indicating that they outperform the proposed FYDSE algorithm for no test function in the statistical test, and that their performance is comparable for the two test functions. Finally, from the statistical results of the original YDSE, it can be found that only one test function exists for YDSE that is statistically superior to FYDSE. However, in the other test function, FYDSE obtains superior experimental results. These detailed test results illustrate the performance differences between the algorithms.
Figure 4 showcases the convergence performance of the FYDSE in comparison to other approaches when evaluated on the cec2022 test function set. The x-axis represents the number of iterations, whereas the y-axis represents the fitness values, where some function results are expressed in logarithmic form with a base of ten. All algorithms start from the same initial point (i.e., iteration zero) to ensure a fair comparison. Observing Figure 4, for the cec01-cec03 functions, FYDSE exhibits a high convergence rate during the initial iteration phase, followed by a gradual localization of the optimal position and validation of the previous findings by updating the position. For the cec04-cec06 functions, FYDSE subsequently obtains the best solution at a satisfactory speed and executes a precise search optimization in proximity to the optimal solution. This performance proves its reliability in avoiding local optima. For the cec07-cec08 function, FYDSE can steer the search over a large area during the initial iteration phase to find potentially high-quality regions in the search space. FYDSE shifts to a localized search as the algorithm proceeds, updating to the optimal location over a smaller area. For the cec09-cec12 functions, FYDSE can swiftly transition between the initial search phase and later stages of development, converging to a solution that is close to optimal at an early iteration. Afterward, it continues pinpointing the optimal position and validating previous observations by updating the results. In summary, FYDSE performs well in all four types of test functions and maintains a significant advantage in most of them. Moreover, these compelling results also demonstrate that the FYDSE algorithm effectively strikes a harmonious balance between exploratory and developmental search strategies.
Figure 5 shows the distribution of the best mean values of FYDSE and other comparison algorithms in multiple test functions through box-and-line diagrams. From the figure, we can see that in most test scenarios, the distribution of the best mean values of FYDSE is more concentrated and compact, fully demonstrating the excellent performance of FYDSE. In addition, this distribution characteristic further proves that FYDSE has excellent consistency and stability under different testing conditions, making it stand out amongst the algorithms. Specifically, when facing the test functions of cec03, cec04, cec07, cec08, and cec12, the box-and-line plot box of FYDSE is narrower, which implies that the performance of FYDSE fluctuates less during multiple iterations, demonstrating its stability and reliability. As for the test functions cec01, cec02, cec05, cec09, cec10, and cec11, the box-and-line diagram of FYDSE shows a red line, which means that FYDSE can solve the problem effectively and achieve a high-performance level during each iteration. This performance once again proves the superiority and usefulness of FYDSE.
Figure 6 visually depicts the capability rankings of the FYDSE against a string of comparative algorithms across the twelve functions of the CEC2022 benchmark test set, utilizing an intuitive radar chart format. Upon examining Figure 6, it becomes evident that the FYDSE exhibits a significantly smaller coverage area within the radar chart, which not only accentuates its outstanding performance across various optimization scenarios but also attests to its exceptional optimization capabilities and stability. Regardless of whether it tackles simple or complex optimization problems, FYDSE can swiftly discover satisfactory solutions through efficient and accurate search strategies.

4.4. Comparison of Algorithm Complexity and Running Results

In order to validate the proposed FYDSE algorithm against other algorithms in terms of complexity and running results, we provide the algorithmic complexity of the proposed algorithm and other methods in Table 4, where Tmax represents the maximum iteration, M represents the number of function evaluations, n represents the number of populations, and d represents the dimension.
The table shows that the YDSE, SO, HBA, SSA, WSO, and COOT algorithms have the same complexity. Also, the DO, AO and SCSO algorithms have the same complexity. The proposed algorithm, only in the second term of the complexity, has a difference in coefficients with YDSE, SO, HBA, SSA, WSO, and COOT, mainly due to the introduced crossover strategy. In fact, real-world optimization problems are complex and nonlinear. Therefore, when confronted with real-world optimization problems, the complexity of the proposed FYDSE is comparable to that of YDSE, SO, HBA, SSA, WSO, COOT, and AHA and outperforms DO, AO, and SCSO and significantly outperforms ChOA. To demonstrate this conclusion more effectively, we provide the runtime results for the ten functions in the cec2022 suite in Table 5.
From the results in the table, the overall operational results of the proposed FYDSE are better than those of the ChOA, SCSO, and DO algorithms. Also, the gap between the running results of the proposed algorithm and the other algorithms is considered small. At the same time, this gap will gradually decrease as the complexity of the objective function of the problem under study increases. Therefore, the complexity of the proposed FYDSE is improved within the acceptable range, and it is important to note that it cannot be ignored that the proposed algorithm obtains better performance results in these benchmark suites.

4.5. Search Capability Analysis of Global Optimal Solutions

In order to verify the improvement in the introduced initialization and exploration strategies on the global solution searchability of the YDSE algorithm, we provide different methods to obtain the probability of the global optimal solution. First, considering that some test functions are difficult to converge to the optimal position, we consider that the candidate solution of the solution is considered optimal if it converges to the position of the neighborhood of the optimal solution, i.e., the relative error ((Candidate Solution − Optimal Solution)/Candidate Solution) is < 10−2. We provide the results of the probability of reaching the optimal solution for the twenty runs of the proposed FYDSE and YDSE, as well as the other methods, in Table 6.
The results in Table 6 show that the proposed FYDSE achieves the best probability of reaching the optimal solution for most of the test functions and has the best final average ranking. Compared with YDSE, the proposed FYDSE has a higher probability of searching for the optimal solution in cec02, cec07, cec08, and cec10. The main reason is that the probability of obtaining the global optimal solution is improved by introducing a dynamic adversarial strategy that effectively extends the full exploration of seldom-searched regions.

5. An Example of Complex Engineering Optimization: The Four-Stage Gearbox Problem

To further validate the performance of the proposed FYDSE in solving complex engineering optimization problems, the proposed algorithm is experimentally compared with other algorithms, including the original YDSE, in a complex engineering optimization example: a four-stage gearbox problem. A schematic diagram of the four-stage gearbox problem is shown in Figure 7. The objective of the problem is to minimize the weight of the gearbox, which contains 22 discrete independent variables [58]. They are categorized into four types, including gear position, pinion position, billet thickness, and number of teeth. In addition, the problem contains 86 nonlinear design constraints related to pitch, kinematics, contact ratio, gear strength, gear assembly, and gear size [44]. The mathematical model is shown below:
Minimize:
f ( x ¯ ) = ( π 1000 )   i = 1 4 b i c i 2 ( N p i 2 + N g i 2 ) ( N p i + N g i ) 2 ,     where ,   i = ( 1 , 2 , 3 , 4 ) .
Subject to:
g 1 ( x ¯ ) = ( 366000 π ω 1 + 2 c 1 N p 1 N p 1 + N g 1 ) ( ( N p 1 + N g 1 ) 2 4 b 1 c 1 2 N p 1 ) σ N J R 0.0167 W K o K m 0 ,
g 2 ( x ¯ ) = ( 366000 N g 1 π ω 1 N p 1 + 2 c 2 N p 2 N p 2 + N g 2 ) ( ( N p 2 + N g 2 ) 2 4 b 2 c 2 2 N p 2 ) σ N J R 0.0167 W K o K m 0 ,
g 3 ( x ¯ ) = ( 366000 N g 1 N g 2 π ω 1 N p 1 N p 2 + 2 c 3 N p 3 N p 3 + N g 3 ) ( ( N p 3 + N g 3 ) 2 4 b 3 c 3 2 N p 3 ) σ N J R 0.0167 W K o K m 0 ,
g 4 ( x ¯ ) = ( 366000 N g 1 N g 2 N g 3 π ω 1 N p 1 N p 2 N p 3 + 2 c 4 N p 4 N p 4 + N g 4 ) ( ( N p 4 + N g 4 ) 2 4 b 4 c 4 2 N p 4 ) σ N J R 0.0167 W K o K m 0 ,
g 5 ( x ¯ ) = ( 366000 π ω 1 + 2 c 1 N p 1 N p 1 + N g 1 ) ( ( N p 1 + N g 1 ) 3 4 b 1 c 1 2 N g 1 N p 1 2 ) ( σ H C p ) 2 ( sin ( φ ) cos ( φ ) 0.0334 W K o K m ) 0 ,
g 6 ( x ¯ ) = ( 366000 N g 1 π ω 1 N p 1 + 2 c 2 N p 2 N p 2 + N g 2 ) ( ( N p 2 + N g 2 ) 3 4 b 2 c 2 2 N g 2 N p 2 2 ) ( σ H C p ) 2 ( sin ( φ ) cos ( φ ) 0.0334 W K o K m ) 0 ,
g 7 ( x ¯ ) = ( 366000 N g 1 N g 2 π ω 1 N p 1 N p 2 + 2 c 3 N p 3 N p 3 + N g 3 ) ( ( N p 3 + N g 3 ) 3 4 b 3 c 3 2 N g 3 N p 3 2 ) ( σ H C p ) 2 ( sin ( φ ) cos ( φ ) 0.0334 W K o K m ) 0 ,
g 8 ( x ¯ ) = ( 366000 N g 1 N g 2 N g 3 π ω 1 N p 1 N p 2 N p 3 + 2 c 4 N p 4 N p 4 + N g 4 ) ( ( N p 4 + N g 4 ) 3 4 b 4 c 4 2 N g 4 N p 4 2 ) ( σ H C p ) 2 ( sin ( φ ) cos ( φ ) 0.0334 W K o K m ) 0 ,
g 9 12 ( x ¯ ) = N p i sin 2 ( φ ) 4 1 N p i + ( 1 N p i ) 2 + N g i sin 2 ( φ ) 4 + 1 N p i ( 1 N p i ) 2 + sin ( φ ) ( N p i + N g i ) 2 + C R min π cos ( φ ) 0 ,
g 13 16 ( x ¯ ) = d min 2 c i N p i N p i + N g i 0 ,
g 17 20 ( x ¯ ) = d min 2 c i N g i N p i + N g i 0 ,
g 21 ( x ¯ ) = x p 1 + ( ( N p 1 + 2 ) c 1 N p 1 + N g 1 ) L max 0 ,
g 22 24 ( x ¯ ) = L max + ( ( N p i + 2 ) c i N p i + N g i ) i = 2 , 3 , 4 + x g ( i 1 ) 0 ,
g 25 ( x ¯ ) = x p 1 + ( N p 1 + 2 ) c 1 N p 1 + N g 1 0 ,
g 26 28 ( x ¯ ) = ( ( N p i + 2 ) c i N p i + N g i x g ( i 1 ) ) i = 2 , 3 , 4 0 ,
g 29 ( x ¯ ) = y p 1 + ( N p 1 + 2 ) c 1 N p 1 + N g 1 L max 0 ,
g 30 32 ( x ¯ ) = L max + ( c i ( 2 + N p i ) N p i + N g i + y g ( i 1 ) ) i = 2 , 3 , 4 0 ,
g 33 ( x ¯ ) = ( 2 + N p 1 ) c 1 N p 1 + N g 1 y p 1 0 ,
g 34 36 ( x ¯ ) = ( c i ( 2 + N p i ) N p i + N g i y g ( i 1 ) ) i = 2 , 3 , 4 0 ,
g 37 40 ( x ¯ ) = L max + c i ( 2 + N g i ) N p i + N g i + x g i 0 ,
g 41 44 ( x ¯ ) = x g i + ( ( N g 1 + 2 ) c i N p i + N g i ) 0 ,
g 45 48 ( x ¯ ) = y g i + ( ( N g i + 2 ) c i N p i + N g i ) L max 0 ,
g 49 52 ( x ¯ ) = y g i + ( ( N g i + 2 ) c i N p i + N g i ) 0 ,
g 53 56 ( x ¯ ) = ( b i 8.255 ) ( b i 5.715 ) ( b i 12.70 ) ( N p i + 0.945 c i N g i ) ( 1 ) 0 ,
g 57 60 ( x ¯ ) = ( b i 8.255 ) ( b i 3.175 ) ( b i 12.70 ) ( N p i + 0.646 c i N g i ) 0 ,
g 61 64 ( x ¯ ) = ( b i 5.715 ) ( b i 3.175 ) ( b i 12.70 ) ( N p i + 0.504 c i N g i ) 0 ,
g 65 68 ( x ¯ ) = ( b i 5.715 ) ( b i 3.175 ) ( b i 8.255 ) ( 0 c i N p i N g i ) 0 ,
g 69 72 ( x ¯ ) = ( b i 8.255 ) ( b i 5.715 ) ( b i 12.70 ) ( N p i + N g i 1.812 c i ) ( 1 ) 0 ,
g 73 76 ( x ¯ ) = ( b i 8.255 ) ( b i 3.175 ) ( b i 12.70 ) ( 0.945 c i + N p i + N g i ) 0 ,
g 77 80 ( x ¯ ) = ( b i 5.715 ) ( b i 3.175 ) ( b i 12.70 ) ( 0.646 c i + N p i + N g i ) ( 1 ) 0 ,
g 81 84 ( x ¯ ) = ( b i 5.715 ) ( b i 3.175 ) ( b i 8.255 ) ( N p i + N g i 0.504 c i ) 0 ,
g 85 ( x ¯ ) = ω min ω 1 ( N p 1 N p 2 N p 3 N p 4 ) N g 1 N g 2 N g 3 N g 4 0 ,
g 86 ( x ¯ ) = ω 1 ( N p 1 N p 2 N p 3 N p 4 ) N g 1 N g 2 N g 3 N g 4 ω max 0 ,
where:
x ¯ = { N p 1 , N g 1 , N p 2 , N g 2 b 1 , b 2 x p 1 , x g 1 , x p 2 , x g 2 y p 1 , y g 1 , y p 2 , y g 2 y g 4 }   , c i = ( y g i y p i ) 2 + ( x g i x p i ) 2 , K o = 1.5 , d min = 25 , J R = 0.2 , φ = 120 , W = 55.9 , K M = 1.6 , C R min = 1.4 , L max = 127 , C p = 464 , σ H = 3290 , ω max = 255 , ω 1 = 5000 , σ N = 2090 , ω min = 245 ,
With bounds:
b i { 3.175 , 12.7 , 8.255 , 5.715 } , y p i , x p i , y g i , x g i { 12.7 , 38.1 , 25.4 , 50.8 , 76.2 , 63.5 , 88.9 , 114.3 , 101.6 } , 7 N g i , N p i 76
Table 7 gives the minimum, worst, and average gearbox weights and values of the design variables obtained after 20 runs of the proposed FYDSE with YDSE, AHA, SO, DO, AO, ChOA, HBA, SCSO, SSA, WSO, and COOT. Also, the running times of all the methods are given at the end of the table. The gearbox weight box plots for FYDSE and other methods at 20 runs are given in Figure 8. From the results, the difference between the running time of the proposed FYDSE method and YDSE is small when facing complex optimization problems due to the high complexity of the objective function. Also, the proposed algorithm obtains the best minimum and average gearbox weight with a significant difference compared to other methods and YDSE. Therefore, the introduced four strategies can effectively improve the algorithm’s ability to face complex engineering and realistic optimization problems.

6. Specific Truss Topology Optimization Problems (TTOP)

To validate the efficacy of FYDSE in tackling intricate engineering optimization challenges, it is employed to optimize the topology of the truss structure. Compared with the cross-section optimization of trusses, topology optimization attempts to streamline the mass of the truss by eliminating unnecessary members and nodes. Therefore, topology optimization can effectively reduce the cost when the cost of nodes is large. In our implementation, we impose frequency constraints on the intrinsic frequencies to ensure that they are effectively limited. Conventional optimization techniques, including sensitivity analysis, have garnered widespread application in addressing truss optimization challenges and have yielded notable outcomes. However, the efficacy of traditional methods in tackling intricate optimization challenges remains in need of enhancement. Meta-heuristic algorithms are an effective solution to this problem since they can fully explore the nonlinear and nonconvex space of topological optimization and maintain the topological results. Therefore, in this section, the efficacy of the FYDSE algorithm in addressing highly nonlinear and nonconvex TTOP problems is further validated. Moving forward, we present the mathematical model for topology optimization.

6.1. TTOP Model

The primary goal of the TTOP was to ascertain the best truss structure and layout to ensure that the ground structure could carry the minimum possible loads. This ground structure comprises both necessary and discretionary nodes. Necessary nodes are typically regarded as promptly bearing structural, load, and nodal stresses [59]. Conversely, including discretionary nodes enhances the stress distribution among the various components. To embark on this endeavor, we first define the key constraints and objective functions of the TTOP model.
Suppose variable X= {A1, A2, …, Ap} and the target function of TTOP is formulated as follows:
F ( A ) = i = 1 p B i A i ρ i L e n g t h i + j = 1 q b j , B i = { 0 , i f   A i < Critical   region 1 , i f   A i Critical   region
where p means numbers of variables and q means numbers of nodes. ρi, Lengthi, Ai signifies the elemental density, length, and transverse area of the ith variable. bj denotes the quantity of the jth node. We set a critical region with a small positive value to determine which elements to discard [60]. When the cross-sectional area of an element falls below a certain minimal threshold, it becomes necessary to discard it.
Further, we give multiple constraints for the TTOP problem, including stress, displacement, Euler buckling, intrinsic frequency, and upper and lower cross-section constraints. The specific definitions are shown below:
σ i min B i σ i σ i max
δ j min δ j δ j max
C i σ i σ i E , σ i E = k i A i E L i 2
f r min f r 0
A k min A k A k max
where σi implies the stress of the ith node.  σ i max  and  σ i min  represent the upper and lower boundary for stress constraints. δj means the displacement of the ith knot.  δ j max  and  δ j min  represent the upper and lower boundary for displacement constraints, respectively. fr means the intrinsic frequency of the truss structure in rth mode. In addition,  A k min  and  A k max  denote the scope of variables, respectively.
In order for the variables to satisfy the above constraints, we introduce a penalty term in the objective function [61]. As illustrated in Equation (39), when the constraints are obeyed, the penalty term will be reduced to zero. However, a positive penalty will be imposed in the event of a constraint violation. The following outlines the specific penalty function in detail:
F ( A ) = { 10 9 ,       i f       structure   is   not   satisfied                   10 8 ,       i f       degree   of   freedom   is   not   satisfied                 10 7 ,                 i f       positive   definiteness   is   not   satisfied             F ( A ) p e n a l t y ,           otherwise                                                                                                          
where the penalty term is set to zero if the independent variable adheres to the specified constraint. F(A) means the weighted objective function. However, if the constraint is violated, a positive penalty is imposed. The precise formulation of this penalty function is detailed as follows:
P e n a l t y = ( 1 + ε i = 1 a c | 1 g i g i | ) ε .
where g denotes the tendency of a variable to violate a constraint and be subject to a  g i , ac representation of active restraint. ε is set to 2 [62].
In order to fully validate the performance and usefulness of FYDSE in solving the optimized TTOPs for complex projects, three TTOPs with different numbers of trusses are discussed in this section. In these three TTOPs, all independent variables are considered to be continuous, and the Euler buckling coefficients and nodal masses are fixed to 4 kg and 5 kg. The cross-section size range is set between Amax and −Amax. To fairly assess the engineering optimization capabilities of FYDSE, we compare it with the original YDSE as well as other latest optimization algorithms, including AHA, SO, DO, AO, ChOA, HBA, SCSO, SSA, WSO, and COOT. To ensure the fairness and repeatability of the optimization TTOP, we made the population size of all algorithms 30 with a maximum iteration of 1000. In addition, all algorithms are tested with 30 independent repetitions.

6.2. 20-Truss Topology Optimization

Figure 9 illustrates the initial schematic ground architecture of a 20-rod truss consisting of 20 members and 9 knots. In this figure, all 20 members and 9 nodes are clearly labeled. Of particular note is that the weights of this truss structure are mainly carried by the supports of node 1 and node 9.
Furthermore, to enhance the optimization of this truss structure, we have compiled a comprehensive list of detailed constraints and inherent material properties in Table 8. Additionally, Table 9 comprehensively displays the experimental outcomes obtained by various comparative methods in addressing the 20-truss optimization problem. Among the experimental results, according to a model introduced for TTOP, we further optimized the design of the truss structure by deleting some unwanted parts (indicated by “-”) according to the requirements.
The outcomes of the independent variables and objective functions of YDSE, AHA, SO, DO, AO, ChOA, HBA, SCSO, SSA, WSO, COOT and the proposed FYDSE are given in Table 9. Upon examination of the research results outlined in Table 9, it is evident that the FYDSE and the AHA, HBA, and COOT all attained the desired minimum optimal weight, measuring precisely 154.799. However, among these algorithms, the proposed FYDSE algorithm shows its unique advantage, with an average weight of only 164.718, significantly smaller than the average weights of other algorithms. In contrast, the SSA and SO algorithms ranked second and third, with their optimal structural qualities of 155.347 and 155.574, which failed to surpass the performance of the FYDSE algorithm despite their proximity. This result indicates that the FYDSE algorithm has higher efficiency and stability in weight optimization.
Figure 10 illustrates the optimal topology achieved through the FYDSE method and a comparative analysis with alternative algorithms during optimization. Specifically, all eight algorithms, FYDSE, YDSE, AHA, SO, DO, HBA, SSA, and COOT, select eight members as the basis of the topology, and these members have the same numbering in their respective algorithms. However, the CHOA algorithm employs a different strategy for retaining only six members. Although the CHOA algorithm requires a more streamlined component structure than that suggested by the FYDSE algorithm, it is more stringent in terms of the quality of the structure. As shown in Figure 10a, YDSE, AHA, SO, DO, HBA, SSA, COOT, and FYDSE present consistency in topology, and their main difference lies in the fact that there are different sizes of cross-sections in the structure. These algorithms provide a valuable reference for solving practical problems, especially when applying metaheuristics. It is worth mentioning that our proposed FYDSE algorithm shows excellent performance in solving a topology optimization problem containing 20 trusses. In contrast to other SOTA algorithms, FYDSE performs better in terms of optimization efficiency and quality of results, further validating its effectiveness and usefulness in topology optimization.

6.3. 24-Truss Topology Optimization

Figure 11 shows a structural diagram for an original ground configuration of a 24-rod truss consisting of 24 members and 8 knots, all clearly labeled. Notably, node 3 supports a non-structural concentrated mass of 500 kg. While this additional weight does not contribute to the overall weight of the truss, it significantly influences the design considerations for the positioning and dimensions of both nodes and members.
To facilitate the optimization process for the 24-rod truss structure, we have compiled a list of pertinent constraints and the inherent material properties in Table 10. Moreover, the results of comparing the method with FYDSE in terms of independent variables and minimum weight is presented in Table 11. Among the experimental results, according to the introduction of the TTOP model, we further optimized the design of the truss structure by deleting some unwanted parts (indicated by “-”) according to the requirements.
The consequences of the independent variables and objective functions of YDSE, AHA, SO, DO, AO, ChOA, HBA, SCSO, SSA, WSO, COOT and the proposed FYDSE are shown in Table 11. Based on the results of the experiment, we can see that the HBA, COOT, and our proposed FYDSE algorithms achieve the same and the lowest result, i.e., 126.252, in seeking the optimal weights, and none of them violates the set constraints. However, among these algorithms, the proposed FYDSE algorithm shows its unique advantage, with an average weight of only 139.951, significantly smaller than the average weights of other algorithms. In contrast, the HBA and COOT algorithms positioned themselves in the second and third spots, respectively, boasting optimal structural qualities of 178.958 and 146.784, which failed to surpass the performance of the FYDSE algorithm despite their proximity. The FYDSE algorithm performs optimally in terms of average weights, a result that further proves the stability and reliability of the FYDSE algorithm.
In Figure 12, a comparative analysis is presented, showcasing the optimal topology attained by the FYDSE algorithm in contrast to that of other algorithms. FYDSE, COOT, SO, AO, and SSA successfully retain seven building blocks among these algorithms. FYDSE and COOT corresponded to the same number of building blocks, while SO, AO, and SSA chose different numbers. Although the HBA algorithm retains only six building blocks, this may imply that it is more stringent regarding structural quality. The topology of the FYDSE algorithm is more reasonable regarding layout and connections looking at the structure of the figure. When dealing with the problem of topology optimization with a 24-rod truss, our proposed FYDSE algorithm outperforms other recent methods.

6.4. 72-Truss Topology Optimization

Figure 13 comprehensively illustrates the original ground structure schematic of a 72-rod truss comprising 72 members and 20 knots, with select members and nodes labeled for clarity. It is crucial to consider the impact of the center on the positioning and dimensions of these members and knots when determining the ultimate total weight of this 72-rod truss structure.
In order to optimize this truss structure, we have detailed the materials’ associated constraints and inherent properties in Table 12.
In addition, the results of comparing the method with FYDSE in terms of independent variables and minimum weight are presented in Table 13. Among the experimental results, according to the introduction of the TTOP model, we further optimized the design of the truss structure by deleting some unwanted parts (indicated by “-”) according to the requirements.
Drawing from the experimental data presented in Table 13, we observe that the SCSO, COOT, and our proposed FYDSE algorithm attain the lowest optimal weight value of 450.388. Nevertheless, the FYDSE algorithm stands out in its performance, exhibiting a significantly lower average weight value of 454.209, surpassing all comparable algorithms. These data further highlight the stability and reliability of the FYDSE algorithm. Among the comparison algorithms, the SO algorithm also performs exceptionally well, with an average weight of 458.213, the second highest, and its structural quality is equally satisfactory.
To have a more intuitive understanding of the optimization effect of these algorithms, Figure 14 exhibits the optimal topologies derived through FYDSE and its comparative algorithms [63,64]. Figure 14a indicates that the YDSE, SO, HBA, SCSO, WSO, COOT, and FYDSE algorithms retain the same number of building blocks during optimization. In addition, the AHA and ChOA algorithms also retain the same number of building blocks. These algorithms remain consistent in the number of building blocks and labeling, but the cross-sectional area in the topology varies in size. In addressing the optimization challenge of truss topology, our FYDSE surpasses the performance of other cutting-edge algorithms, exhibiting remarkable efficiency and effectiveness. It proves once again its superiority in the field of optimal structural design.

7. Discussion

By comparing the results in this study with other methods, the proposed FYDSE shows excellent optimization performance. Based on two experiments, the advantages of FYDSE are specifically discussed. The first part compares FYDSE with other methods on the CEC2022 test function. Better results are obtained on 11 functions compared to YDSE, demonstrating that the introduction of the four strategies sufficiently improves the algorithmic shortcomings of YDSE. Meanwhile, the FYDSE algorithm optimized eight functions, which accounted for 66.6% of all functions. This result shows that the fractional-order modified mechanism, dynamic opposition, and vertical crossover operator enhance the effective search and comprehensive utilization of YDSE, which confirms that FYDSE optimizes the test problem with higher accuracy, better reliability, and faster conversion speed. The second part discusses the performance of FYDSE with other methods in three truss topology optimizations. FYDSE has the best average and optimal weights in all three cases. The results validate the effectiveness and reliability of FYDSE in engineering optimization. Therefore, the advantages of FYDSE in solving numerical and engineering optimization problems are summarized as follows:
(1)
FYDSE ranks better than the original YDSE in most CEC2022 tests. The improved strategy introduced by FYDSE can efficiently balance convergence speed and convergence accuracy and search for better solutions.
(2)
In the CEC2022 test, FYDSE outperforms other algorithms, which shows that FYDSE has a smooth development and exploration process, and is not easily trapped in localized solutions when facing multi-type problems.
(3)
The optimal average weights and minimum weights of FYDSE in the three truss topology optimization indicate the effectiveness, reliability, and stability of FYDSE in solving engineering optimization problems.

8. Conclusions

In order to solve the problems that the original YDSE will easily fall into local optimums, converge slowly, and have an imbalance between exploration and exploitation when dealing with sophisticated engineering optimal problems, this paper presents a fractional-order boosted hybrid YDSE for solving complex optimization problems. FYDSE introduces a piecewise chaotic mapping strategy, a fractional-order improvement strategy, a dynamic dyadic strategy, and a vertical operator. First, the fractional-order strategy is introduced in the dark edge position update of FYDSE to ensure that the search potential of a single neighborhood space is exploited more efficiently while decreasing the likelihood of trapping in a local optimum. Secondly, during the initialization phase, the incorporation of piecewise chaotic mapping is intended to yield a high-quality primary population, thereby enhancing algorithm convergence efficiency. Furthermore, the low exploration space of FYDSE is extended by using a dynamic inversion strategy, which improves the probability of obtaining a globally optimal solution. Finally, by introducing the vertical operator, FYDSE can better balance global exploration and local exploitation and explore new unknown areas. Comparative experiments of FYDSE with a series of state-of-the-art algorithms are conducted in the CEC2022 test suite, the four-stage gearbox problem, and three TTOP cases, respectively. The Wilcoxon rank-sum test statistically proves the effectiveness of FYDSE. In addition, the average ranking of the mean result and probability of FYDSE in cec2022 are 1.42 and 1.33, respectively, verifying that FYDSE can better converge to the optimal solution. Although FYDSE increases the constant complexity by introducing improvement strategies, this difference is small when facing complex problems as analyzed through experiments. The efficient performance of the four-stage gearbox problem and the three TTOP cases verifies that FYDSE is an effective method for solving complex engineering optimization challenges.
Many complex, nonlinear, and high-dimensional realistic optimization problems, such as complex path planning for UAVs, neural network parameter optimization, and curved shape parameter optimization problems, are yet to be solved in scientific and practical research. The better experimental results for numerical and engineering optimization problems show that the proposed FYDSE is an optimization method that can be extended in different fields. Therefore, FYDSE can be used as a potential solution to solve various complex real-world problems effectively.
As a direction for future work, we will strive to extend FYDSE to other domain-specific optimization problems, such as UAV path planning and image threshold segmentation. Moreover, we will also explore the integration of FYDSE with machine learning and deep learning techniques, such as using FYDSE to solve feature selection problems or for deep learning parameter optimization.

Author Contributions

Conceptualization, S.Q., J.L. and G.H.; Methodology, S.Q., J.L., X.B. and G.H.; Software, S.Q. and X.B.; Validation, S.Q., J.L., X.B. and G.H.; Formal analysis, J.L. and G.H.; Investigation, S.Q., J.L., X.B. and G.H.; Resources, S.Q., J.L. and G.H.; Data curation, X.B.; Writing—original draft, S.Q., J.L., X.B. and G.H.; Writing—review & editing, S.Q., J.L., X.B. and G.H.; Visualization, X.B. and G.H.; Supervision, S.Q. and J.L.; Project administration, S.Q., J.L. and X.B.; Funding acquisition, J.L. and G.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during the study are included in this published article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shang, R.; Zhong, J.; Zhang, W.; Xu, S.; Li, Y. Multilabel Feature Selection via Shared Latent Sublabel Structure and Simultaneous Orthogonal Basis Clustering. IEEE Trans. Neural Netw. Learn. Syst. 2024, 1–16. [Google Scholar] [CrossRef] [PubMed]
  2. Hu, G.; Huang, F.; Seyyedabbasi, A.; Wei, G. Enhanced multi-strategy bottlenose dolphin optimizer for UAVs path planning. Appl. Math. Model. 2024, 130, 243–271. [Google Scholar] [CrossRef]
  3. Jia, H.; Lu, C. Guided learning strategy: A novel update mechanism for metaheuristic algorithms design and improvement. Knowl.-Based Syst. 2024, 286, 111402. [Google Scholar] [CrossRef]
  4. Hu, G.; Zhu, X.; Wei, G.; Chang, C.-T. An improved marine predators algorithm for shape optimization of developable Ball surfaces. Eng. Appl. Artif. Intell. 2021, 105, 104417. [Google Scholar] [CrossRef]
  5. Wadood, A.; Park, H. A Novel Application of Fractional Order Derivative Moth Flame Optimization Algorithm for Solving the Problem of Optimal Coordination of Directional Overcurrent Relays. Fractal Fract. 2024, 8, 251. [Google Scholar] [CrossRef]
  6. Aldosary, A. Power Quality Conditioners-Based Fractional-Order PID Controllers Using Hybrid Jellyfish Search and Particle Swarm Algorithm for Power Quality Enhancement. Fractal Fract. 2024, 8, 140. [Google Scholar] [CrossRef]
  7. Zhong, J.; Shang, R.; Zhao, F.; Zhang, W.; Xu, S. Negative Label and Noise Information Guided Disambiguation for Partial Multi-Label Learning. IEEE Trans. Multimed. 2024, 1–16. [Google Scholar] [CrossRef]
  8. Jia, H.; Peng, X.; Lang, C. Remora optimization algorithm. Expert Syst. Appl. 2021, 185, 115665. [Google Scholar] [CrossRef]
  9. Hu, G.; Zhong, J.; Wang, X.; Wei, G. Multi-strategy assisted chaotic coot-inspired optimization algorithm for medical feature selection: A cervical cancer behavior risk study. Comput. Biol. Med. 2022, 151, 106239. [Google Scholar] [CrossRef]
  10. Hu, G.; Guo, Y.; Wei, G.; Abualigah, L. Genghis Khan shark optimizer: A novel nature-inspired algorithm for engineering optimization. Adv. Eng. Inform. 2023, 58, 102210. [Google Scholar] [CrossRef]
  11. Zhang, Q.; Gao, H.; Zhan, Z.-H.; Li, J.; Zhang, H. Growth Optimizer: A powerful metaheuristic algorithm for solving continuous and discrete global optimization problems. Knowl.-Based Syst. 2023, 261, 110206. [Google Scholar] [CrossRef]
  12. Abdel-Basset, M.; Mohamed, R.; Jameel, M.; Abouhawwash, M. Nutcracker optimizer: A novel nature-inspired metaheuristic algorithm for global optimization and engineering design problems. Knowl.-Based Syst. 2023, 262, 110248. [Google Scholar] [CrossRef]
  13. Zhao, W.; Wang, L.; Zhang, Z. Artificial ecosystem-based optimization: A novel nature-inspired meta-heuristic algorithm. Neural Comput. Appl. 2020, 32, 9383–9425. [Google Scholar] [CrossRef]
  14. Zhong, C.; Li, G.; Meng, Z.; He, W. Opposition-based learning equilibrium optimizer with Levy flight and evolutionary population dynamics for high-dimensional global optimization problems. Expert Syst. Appl. 2023, 215, 119303. [Google Scholar] [CrossRef]
  15. Lambora, A.; Gupta, K.; Chopra, K. Genetic Algorithm—A Literature Review. In Proceedings of the 2019 International Conference on Machine Learning, Big Data, Cloud and Parallel Computing (COMITCon), Faridabad, India, 14–16 February 2019; pp. 380–384. [Google Scholar]
  16. Das, S.; Suganthan, P.N. Differential Evolution: A Survey of the State-of-the-Art. IEEE Trans. Evol. Comput. 2011, 15, 4–31. [Google Scholar] [CrossRef]
  17. Kala, R.; Khan, A.; Diksha, D.; Shelly, S.; Sinha, S. Evolutionary Mission Planning. In Proceedings of the 2018 IEEE Congress on Evolutionary Computation (CEC), Rio de Janeiro, Brazil, 8–13 July 2018; pp. 1–8. [Google Scholar]
  18. Guilmeau, T.; Chouzenoux, E.; Elvira, V. Simulated Annealing: A Review and a New Scheme. In Proceedings of the 2021 IEEE Statistical Signal Processing Workshop (SSP), Rio de Janeiro, Brazil, 11–14 July 2021; pp. 101–105. [Google Scholar]
  19. Hashim, F.A.; Mostafa, R.R.; Hussien, A.G.; Mirjalili, S.; Sallam, K.M. Fick’s Law Algorithm: A physical law-based algorithm for numerical optimization. Knowl.-Based Syst. 2023, 260, 110146. [Google Scholar] [CrossRef]
  20. Rashedi, E.; Nezamabadi-pour, H.; Saryazdi, S. GSA: A Gravitational Search Algorithm. Inf. Sci. 2009, 179, 2232–2248. [Google Scholar] [CrossRef]
  21. Cheraghalipour, A.; Hajiaghaei-Keshteli, M.; Paydar, M.M. Tree Growth Algorithm (TGA): A novel approach for solving optimization problems. Eng. Appl. Artif. Intell. 2018, 72, 393–414. [Google Scholar] [CrossRef]
  22. Zhong, J.; Shang, R.; Xu, S.; Li, Y. Graph embedding orthogonal decomposition: A synchronous feature selection technique based on collaborative particle swarm optimization. Pattern Recognit. 2024, 152, 110453. [Google Scholar] [CrossRef]
  23. Mirjalili, S.; Mirjalili, S.M.; Lewis, A. Grey Wolf Optimizer. Adv. Eng. Softw. 2014, 69, 46–61. [Google Scholar] [CrossRef]
  24. Heidari, A.A.; Mirjalili, S.; Faris, H.; Aljarah, I.; Mafarja, M.; Chen, H. Harris hawks optimization: Algorithm and applications. Future Gener. Comput. Syst. 2019, 97, 849–872. [Google Scholar] [CrossRef]
  25. Wang, J.; Wang, W.-C.; Hu, X.-X.; Qiu, L.; Zang, H.-F. Black-winged kite algorithm: A nature-inspired meta-heuristic for solving benchmark functions and engineering problems. Artif. Intell. Rev. 2024, 57, 98. [Google Scholar] [CrossRef]
  26. Askari, Q.; Younas, I.; Saeed, M. Political Optimizer: A novel socio-inspired meta-heuristic for global optimization. Knowl.-Based Syst. 2020, 195, 105709. [Google Scholar] [CrossRef]
  27. Lian, J.; Hui, G. Human Evolutionary Optimization Algorithm. Expert Syst. Appl. 2024, 241, 122638. [Google Scholar] [CrossRef]
  28. Samareh Moosavi, S.H.; Bardsiri, V.K. Poor and rich optimization algorithm: A new human-based and multi populations algorithm. Eng. Appl. Artif. Intell. 2019, 86, 165–181. [Google Scholar] [CrossRef]
  29. Abdel-Basset, M.; El-Shahat, D.; Jameel, M.; Abouhawwash, M. Young’s double-slit experiment optimizer: A novel metaheuristic optimization algorithm for global and constraint optimization problems. Comput. Methods Appl. Mech. Eng. 2023, 403, 115652. [Google Scholar] [CrossRef]
  30. Hu, G.; Guo, Y.; Zhong, J.; Wei, G. IYDSE: Ameliorated Young’s double-slit experiment optimizer for applied mechanics and engineering. Comput. Methods Appl. Mech. Eng. 2023, 412, 116062. [Google Scholar] [CrossRef]
  31. Dong, Y.; Sun, Y.; Liu, Z.; Du, Z.; Wang, J. Predicting dissolved oxygen level using Young’s double-slit experiment optimizer-based weighting model. J. Environ. Manag. 2024, 351, 119807. [Google Scholar] [CrossRef] [PubMed]
  32. Wolpert, D.H.; Macready, W.G. No free lunch theorems for optimization. IEEE Trans. Evol. Comput. 1997, 1, 67–82. [Google Scholar] [CrossRef]
  33. Hu, G.; Zhong, J.; Wei, G. SaCHBA_PDN: Modified honey badger algorithm with multi-strategy for UAV path planning. Expert Syst. Appl. 2023, 223, 119941. [Google Scholar] [CrossRef]
  34. Lei, Y.; Fan, L.; Yang, J.; Si, W. Fractional-Order Boosted Jellyfish Search Optimizer with Gaussian Mutation for Income Forecast of Rural Resident. Comput. Intell. Neurosci. 2022, 2022, 3343505. [Google Scholar] [CrossRef] [PubMed]
  35. Solteiro Pires, E.J.; Tenreiro Machado, J.A.; de Moura Oliveira, P.B.; Boaventura Cunha, J.; Mendes, L. Particle swarm optimization with fractional-order velocity. Nonlinear Dyn. 2010, 61, 295–301. [Google Scholar] [CrossRef]
  36. Luo, K.; Jiao, Y.; Zhu, J. Perturbation Observer Based Fractional-Order Control for SMES Systems Based on Jellyfish Search Algorithm. Front. Energy Res. 2021, 9, 781774. [Google Scholar] [CrossRef]
  37. Zhang, P.; Yang, R.; Yang, R.; Ren, G.; Yang, X.; Xu, C.; Xu, B.; Zhang, H.; Cai, Y.; Lu, Y. Parameter estimation for fractional-order chaotic systems by improved bird swarm optimization algorithm. Int. J. Mod. Phys. C 2019, 30, 1950086. [Google Scholar] [CrossRef]
  38. Ortigueira, M.D.; Tenreiro Machado, J.A. What is a fractional derivative? J. Comput. Phys. 2015, 293, 4–13. [Google Scholar] [CrossRef]
  39. Naruei, I.; Keynia, F. Wild horse optimizer: A new meta-heuristic algorithm for solving engineering optimization problems. Eng. Comput. 2022, 38, 3025–3056. [Google Scholar] [CrossRef]
  40. Gandomi, A.H.; Yang, X.S.; Talatahari, S.; Alavi, A.H. Firefly algorithm with chaos. Commun. Nonlinear Sci. Numer. Simul. 2013, 18, 89–98. [Google Scholar] [CrossRef]
  41. Yuan, X.; Zhao, J.; Yang, Y.; Wang, Y. Hybrid parallel chaos optimization algorithm with harmony search algorithm. Appl. Soft Comput. 2014, 17, 12–22. [Google Scholar] [CrossRef]
  42. Xu, Y.; Yang, Z.; Li, X.; Kang, H.; Yang, X. Dynamic opposite learning enhanced teaching–learning-based optimization. Knowl.-Based Syst. 2020, 188, 104966. [Google Scholar] [CrossRef]
  43. Arini, F.Y.; Chiewchanwattana, S.; Soomlek, C.; Sunat, K. Joint Opposite Selection (JOS): A premiere joint of selective leading opposition and dynamic opposite enhanced Harris’ hawks optimization for solving single-objective problems. Expert Syst. Appl. 2022, 188, 116001. [Google Scholar] [CrossRef]
  44. Hu, G.; Zhong, J.; Du, B.; Wei, G. An enhanced hybrid arithmetic optimization algorithm for engineering applications. Comput. Methods Appl. Mech. Eng. 2022, 394, 114901. [Google Scholar] [CrossRef]
  45. Meng, A.-B.; Chen, Y.-C.; Yin, H.; Chen, S.-Z. Crisscross optimization algorithm and its application. Knowl.-Based Syst. 2014, 67, 218–229. [Google Scholar] [CrossRef]
  46. 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] [CrossRef]
  47. 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]
  48. Hashim, F.A.; Hussien, A.G. Snake Optimizer: A novel meta-heuristic optimization algorithm. Knowl.-Based Syst. 2022, 242, 108320. [Google Scholar] [CrossRef]
  49. Zhao, S.; Zhang, T.; Ma, S.; Chen, M. Dandelion Optimizer: A nature-inspired metaheuristic algorithm for engineering applications. Eng. Appl. Artif. Intell. 2022, 114, 105075. [Google Scholar] [CrossRef]
  50. Abualigah, L.; Yousri, D.; Abd Elaziz, M.; Ewees, A.A.; Al-qaness, M.A.A.; Gandomi, A.H. Aquila Optimizer: A novel meta-heuristic optimization algorithm. Comput. Ind. Eng. 2021, 157, 107250. [Google Scholar] [CrossRef]
  51. Khishe, M.; Mosavi, M.R. Chimp optimization algorithm. Expert Syst. Appl. 2020, 149, 113338. [Google Scholar] [CrossRef]
  52. Hashim, F.A.; Houssein, E.H.; Hussain, K.; Mabrouk, M.S.; Al-Atabany, W. Honey Badger Algorithm: New metaheuristic algorithm for solving optimization problems. Math. Comput. Simul. 2022, 192, 84–110. [Google Scholar] [CrossRef]
  53. Seyyedabbasi, A.; Kiani, F. Sand Cat swarm optimization: A nature-inspired algorithm to solve global optimization problems. Eng. Comput. 2023, 39, 2627–2651. [Google Scholar] [CrossRef]
  54. Mirjalili, S.; Gandomi, A.H.; Mirjalili, S.Z.; Saremi, S.; Faris, H.; Mirjalili, S.M. Salp Swarm Algorithm: A bio-inspired optimizer for engineering design problems. Adv. Eng. Softw. 2017, 114, 163–191. [Google Scholar] [CrossRef]
  55. Braik, M.; Hammouri, A.; Atwan, J.; Al-Betar, M.A.; Awadallah, M.A. White Shark Optimizer: A novel bio-inspired meta-heuristic algorithm for global optimization problems. Knowl.-Based Syst. 2022, 243, 108457. [Google Scholar] [CrossRef]
  56. Naruei, I.; Keynia, F. A new optimization method based on COOT bird natural life model. Expert Syst. Appl. 2021, 183, 115352. [Google Scholar] [CrossRef]
  57. Hussain, K.; Salleh, M.N.M.; Cheng, S.; Shi, Y. On the exploration and exploitation in popular swarm-based metaheuristic algorithms. Neural Comput. Appl. 2019, 31, 7665–7683. [Google Scholar] [CrossRef]
  58. Yokota, T.; Taguchi, T.; Gen, M. A solution method for optimal weight design problem of the gear using genetic algorithms. Comput. Ind. Eng. 1998, 35, 523–526. [Google Scholar] [CrossRef]
  59. Wang, J.; Li, Y.; Hu, G.; Yang, M. An enhanced artificial hummingbird algorithm and its application in truss topology engineering optimization. Adv. Eng. Inform. 2022, 54, 101761. [Google Scholar] [CrossRef]
  60. Savsani, V.J.; Tejani, G.G.; Patel, V.K. Truss topology optimization with static and dynamic constraints using modified subpopulation teaching–learning-based optimization. Eng. Optim. 2016, 48, 1990–2006. [Google Scholar] [CrossRef]
  61. Kaveh, A.; Zolghadr, A. Topology optimization of trusses considering static and dynamic constraints using the CSS. Appl. Soft Comput. 2013, 13, 2727–2734. [Google Scholar] [CrossRef]
  62. Hu, G.; Zhong, J.; Wei, G.; Chang, C.-T. DTCSMO: An efficient hybrid starling murmuration optimizer for engineering applications. Comput. Methods Appl. Mech. Eng. 2023, 405, 115878. [Google Scholar] [CrossRef]
  63. Jia, H.; Lu, C.; Xing, Z. Memory backtracking strategy: An evolutionary updating mechanism for meta-heuristic algorithms. Swarm Evol. Comput. 2024, 84, 101456. [Google Scholar] [CrossRef]
  64. Jia, H.; Rao, H.; Wen, C.; Mirjalili, S. Crayfish optimization algorithm. Artif. Intell. Rev. 2023, 56, 1919–1979. [Google Scholar] [CrossRef]
Figure 1. Schematic of the fractional-order modification strategy.
Figure 1. Schematic of the fractional-order modification strategy.
Biomimetics 09 00474 g001
Figure 2. Main flowchart of FYDSE.
Figure 2. Main flowchart of FYDSE.
Biomimetics 09 00474 g002
Figure 3. Iterative plots of exploration and exploitation rates of the FYDSE algorithm for 12 test functions in CEC2022.
Figure 3. Iterative plots of exploration and exploitation rates of the FYDSE algorithm for 12 test functions in CEC2022.
Biomimetics 09 00474 g003
Figure 4. Plot of convergence of FYDSE and another comparative algorithm on the CEC2022 benchmark function.
Figure 4. Plot of convergence of FYDSE and another comparative algorithm on the CEC2022 benchmark function.
Biomimetics 09 00474 g004aBiomimetics 09 00474 g004b
Figure 5. Box plot of FYDSE versus another comparative algorithm on the CEC2022 test function set.
Figure 5. Box plot of FYDSE versus another comparative algorithm on the CEC2022 test function set.
Biomimetics 09 00474 g005aBiomimetics 09 00474 g005b
Figure 6. Bar plot of FYDSE versus another comparative algorithm on the CEC2022 test function set.
Figure 6. Bar plot of FYDSE versus another comparative algorithm on the CEC2022 test function set.
Biomimetics 09 00474 g006aBiomimetics 09 00474 g006b
Figure 7. Schematic diagram of the four-stage gearbox problem.
Figure 7. Schematic diagram of the four-stage gearbox problem.
Biomimetics 09 00474 g007
Figure 8. Box plots of gearbox weights for FYDSE and other methods at 20 runs.
Figure 8. Box plots of gearbox weights for FYDSE and other methods at 20 runs.
Biomimetics 09 00474 g008
Figure 9. Schematic diagram of 20-rod truss structures.
Figure 9. Schematic diagram of 20-rod truss structures.
Biomimetics 09 00474 g009
Figure 10. FYDSE and comparison algorithms for 20-bar truss topology optimization structures.
Figure 10. FYDSE and comparison algorithms for 20-bar truss topology optimization structures.
Biomimetics 09 00474 g010
Figure 11. Schematic diagram of 24-rod truss structures.
Figure 11. Schematic diagram of 24-rod truss structures.
Biomimetics 09 00474 g011
Figure 12. FYDSE and comparison algorithms for 24-rod truss topology optimization structures.
Figure 12. FYDSE and comparison algorithms for 24-rod truss topology optimization structures.
Biomimetics 09 00474 g012
Figure 13. Schematic diagram of 72-rod truss structure.
Figure 13. Schematic diagram of 72-rod truss structure.
Biomimetics 09 00474 g013
Figure 14. FYDSE and comparison algorithms for 72-rod truss topology optimization structures.
Figure 14. FYDSE and comparison algorithms for 72-rod truss topology optimization structures.
Biomimetics 09 00474 g014
Table 1. Configuring algorithm parameters, where the parameter settings for the comparison experiments are all referenced to the results provided in the original references.
Table 1. Configuring algorithm parameters, where the parameter settings for the comparison experiments are all referenced to the results provided in the original references.
AlgorithmsYearParameter Setting
YDSE2023Wavelength (λ) = 5 × 10−6 m, distance between two slits (d) = 5 × 10−3 m, distance between the barrier and the projection screen (L) = 1 m, distance between light source and barrier (I) = 0.01 m, constant value (δ) = 0.38.
AHA2022Migration coefficient = 2n.
DO2022α = [0, 1], k = [0, 1].
ChOA2020m is chaotic.
HBA2022β (the ability of a honey badger to get food) = 6, C = 2.
SCSO2022Sensitivity range (rG) = [2, 0], phases control range (R) = [−2rG, 2rG].
SSA2020ST = 0.8.
WSO2022fmin = 0.07, fmax = 0.75, τ = 4.125, a0 = 6.25, a1 = 100, a2 = 0.0005.
Table 2. Results of FYDSE and various comparative methods on cec2022; the optimal value is in bold.
Table 2. Results of FYDSE and various comparative methods on cec2022; the optimal value is in bold.
FunctionIndexAlgorithms
YDSEAHASODOAOChOAHBASCSOSSAWSOCOOTFYDSE
cec01Best3.00E+023.00E+023.00E+023.00E+026.58E+028.39E+023.00E+023.10E+023.00E+023.00E+023.00E+023.00E+02
Worst3.00E+023.03E+023.06E+023.00E+026.03E+034.52E+033.00E+025.24E+033.00E+021.10E+033.00E+023.00E+02
Mean3.00E+023.00E+023.01E+023.00E+021.92E+032.25E+033.00E+021.64E+033.00E+024.60E+023.00E+023.00E+02
Std2.37E−045.96E−011.48E+008.21E−031.50E+031.21E+038.16E−061.65E+037.41E−102.11E+022.33E−025.38E−05
Rank478511122101963
cec02Min4.00E+024.00E+024.00E+024.00E+024.02E+024.34E+024.00E+024.00E+024.04E+024.00E+024.00E+024.00E+02
Max4.09E+024.71E+024.64E+024.75E+024.53E+022.05E+034.09E+024.74E+024.09E+024.88E+024.90E+024.00E+02
Mean4.04E+024.06E+024.06E+024.21E+024.19E+027.28E+024.04E+024.26E+024.06E+024.12E+024.17E+024.00E+02
Std2.50E+001.57E+011.40E+013.06E+011.47E+013.64E+022.86E+002.52E+011.59E+002.71E+012.90E+012.00E−02
Rank245109123116781
cec03Min6.01E+026.00E+026.00E+026.00E+026.05E+026.13E+026.00E+026.04E+026.01E+026.00E+026.00E+026.00E+02
Max6.06E+026.09E+026.06E+026.20E+026.39E+026.48E+026.01E+026.37E+026.37E+026.10E+026.08E+026.01E+02
Mean6.02E+026.01E+026.01E+026.05E+026.17E+026.29E+026.00E+026.17E+026.09E+026.02E+026.03E+026.00E+02
Std1.20E+002.01E+001.37E+004.50E+008.86E+008.69E+003.45E−018.36E+009.13E+002.80E+002.61E+001.57E−01
Rank634811121109572
cec04Min8.05E+028.16E+028.07E+028.15E+028.12E+028.20E+028.07E+028.10E+028.07E+028.03E+028.08E+028.04E+02
Max8.24E+028.39E+028.22E+028.55E+028.33E+028.48E+028.24E+028.34E+028.55E+028.18E+028.27E+028.18E+02
Mean8.13E+028.25E+028.16E+028.33E+028.22E+028.38E+028.18E+028.25E+028.28E+028.07E+028.15E+028.11E+02
Std4.98E+006.74E+004.51E+001.13E+015.52E+007.75E+004.10E+007.71E+001.22E+013.56E+004.96E+003.60E+00
Rank385117126910142
cec05Min9.00E+029.00E+029.00E+029.00E+029.35E+021.09E+039.00E+029.23E+029.00E+029.00E+029.00E+029.00E+02
Max9.17E+021.14E+031.04E+031.51E+031.33E+031.62E+039.53E+021.38E+031.29E+031.07E+031.01E+039.18E+02
Mean9.03E+029.37E+029.22E+021.01E+031.02E+031.28E+039.08E+021.12E+039.20E+029.28E+029.11E+029.04E+02
Std3.65E+005.55E+013.24E+011.90E+029.30E+011.59E+021.66E+011.43E+028.69E+014.76E+012.51E+015.19E+00
Rank186910123115742
cec06Min1.80E+031.82E+031.85E+031.97E+035.24E+034.82E+041.98E+031.88E+032.09E+031.81E+031.87E+031.80E+03
Max1.81E+035.06E+035.69E+038.11E+032.81E+043.58E+068.10E+038.16E+035.26E+041.89E+037.93E+031.80E+03
Mean1.80E+032.53E+033.26E+034.34E+031.40E+041.13E+064.28E+034.21E+031.13E+041.83E+033.29E+031.80E+03
Std2.66E+009.41E+021.31E+031.82E+037.23E+038.42E+051.96E+032.32E+031.26E+041.79E+011.67E+035.42E−01
Rank245911128710361
cec07Min2.02E+032.00E+032.00E+032.02E+032.03E+032.04E+032.00E+032.02E+032.02E+032.00E+032.01E+032.00E+03
Max2.03E+032.02E+032.06E+032.04E+032.13E+032.07E+032.05E+032.08E+032.06E+032.05E+032.05E+032.02E+03
Mean2.03E+032.01E+032.03E+032.03E+032.05E+032.06E+032.02E+032.05E+032.04E+032.01E+032.03E+032.01E+03
Std3.67E+009.87E+001.34E+017.37E+002.25E+016.04E+001.01E+011.77E+011.04E+011.17E+019.59E+006.01E+00
Rank627810124119351
cec08Min2.21E+032.20E+032.21E+032.22E+032.21E+032.23E+032.22E+032.21E+032.22E+032.20E+032.20E+032.20E+03
Max2.23E+032.22E+032.22E+032.23E+032.24E+032.37E+032.44E+032.24E+032.35E+032.23E+032.23E+032.22E+03
Mean2.22E+032.22E+032.22E+032.23E+032.23E+032.34E+032.25E+032.23E+032.25E+032.22E+032.22E+032.21E+03
Std5.35E+007.13E+003.43E+003.03E+006.75E+004.52E+016.11E+017.41E+004.06E+017.30E+006.85E+004.13E+00
Rank234791211810651
cec09Min2.41E+032.53E+032.53E+032.53E+032.53E+032.55E+032.53E+032.53E+032.49E+032.53E+032.53E+032.40E+03
Max2.53E+032.53E+032.53E+032.53E+032.63E+032.78E+032.53E+032.70E+032.71E+032.57E+032.68E+032.53E+03
Mean2.52E+032.53E+032.53E+032.53E+032.57E+032.60E+032.53E+032.58E+032.54E+032.54E+032.54E+032.52E+03
Std2.57E+011.13E−041.48E−133.42E−042.43E+015.13E+011.49E+004.99E+018.89E+019.29E+003.29E+012.84E+01
Rank243510126119781
cec10Min2.50E+032.50E+032.50E+032.41E+032.50E+032.50E+032.50E+032.50E+032.50E+032.50E+032.50E+032.40E+03
Max2.50E+032.63E+032.65E+032.64E+032.63E+034.11E+032.63E+032.64E+033.39E+032.63E+032.64E+032.50E+03
Mean2.50E+032.54E+032.54E+032.56E+032.57E+032.73E+032.55E+032.57E+032.63E+032.55E+032.52E+032.43E+03
Std8.88E−025.50E+016.18E+016.93E+015.95E+014.96E+025.75E+016.33E+012.56E+025.96E+014.59E+012.32E+01
Rank245810126911731
cec11Min2.60E+032.60E+032.60E+032.60E+032.61E+033.04E+032.60E+032.60E+032.60E+032.60E+032.60E+032.60E+03
Max2.60E+032.90E+032.90E+033.00E+032.75E+033.90E+033.18E+033.18E+032.90E+033.00E+032.75E+032.60E+03
Mean2.60E+032.65E+032.66E+032.75E+032.69E+033.53E+032.65E+032.82E+032.68E+032.67E+032.62E+032.60E+03
Std2.41E−018.83E+011.02E+021.60E+027.20E+012.67E+021.54E+021.79E+021.33E+021.15E+024.63E+012.48E−02
Rank256109124118731
cec12Min2.86E+032.86E+032.86E+032.86E+032.86E+032.87E+032.86E+032.86E+032.85E+032.87E+032.86E+032.86E+03
Max2.86E+032.88E+032.88E+032.89E+032.87E+032.93E+032.95E+032.95E+032.90E+032.94E+032.87E+032.86E+03
Mean2.86E+032.87E+032.87E+032.87E+032.87E+032.88E+032.89E+032.87E+032.89E+032.88E+032.87E+032.86E+03
Std1.31E+003.48E+003.66E+007.53E+002.16E+001.94E+013.35E+011.91E+012.34E+012.12E+011.47E+001.56E+00
Rank256741012811931
Mean ranks2.83E+004.75E+005.33E+008.08E+009.25E+001.18E+015.50E+009.67E+008.25E+005.92E+005.17E+001.42E+00
Final ranking235810126119741
Table 3. The p-values from the Wilcoxon test across various test functions.
Table 3. The p-values from the Wilcoxon test across various test functions.
FunctionsYDSEAHASODOAOChOAHBASCSOSSAWSOCOOT
cec015.17E−06/-1.66E−07/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-5.24E−08/-6.80E−08/-8.01E−09/-6.80E−08/-6.80E−08/-
cec027.90E−08/-1.80E−06/-2.55E−07/-1.05E−06/-6.80E−08/-6.80E−08/-3.89E−07/-6.80E−08/-6.80E−08/-2.36E−06/-1.05E−06/-
cec036.80E−08/-1.60E−05/-9.05E−03/-3.50E−06/-6.80E−08/-6.80E−08/-2.47E−04/-6.80E−08/-1.06E−07/-8.10E−02/=1.41E−05/-
cec041.64E−01/=1.05E−07/-1.12E−03/-9.17E−08/-2.96E−07/-6.80E−08/-2.57E−05/-1.05E−06/-5.16E−06/-1.23E−03/-2.07E−02/-
cec057.56E−01/+1.63E−03/-4.11E−02/-1.55E−02/-6.80E−08/-6.80E−08/-6.17E−01/=6.80E−08/-1.05E−03/-2.80E−03/-8.18E−01/=
cec063.99E−06/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-6.80E−08/-
cec073.94E−07/-7.76E−01/=9.75E−06/-1.23E−07/-6.80E−08/-6.80E−08/-8.29E−05/-1.06E−07/-7.90E−08/-2.08E−01/=3.07E−06/-
cec081.38E−06/-5.63E−04/-1.20E−06/-6.80E−08/-1.23E−07/-6.80E−08/-1.23E−07/-6.01E−07/-6.80E−08/-5.87E−06/-8.60E−06/-
cec091.94E−08/-4.63E−05/-5.74E−01/=1.95E−08/-1.95E−08/-1.95E−08/-5.02E−03/-1.95E−08/-6.52E−02/=1.95E−08/-1.69E−04/-
cec107.90E−08/-2.22E−07/-9.17E−08/-7.95E−07/-6.80E−08/-6.80E−08/-1.23E−07/-9.17E−08/-9.17E−08/-9.17E−08/-7.90E−08/-
cec119.17E−08/-3.15E−02/-3.15E−02/-2.69E−06/-6.80E−08/-6.80E−08/-6.04E−06/-6.80E−08/-3.15E−02/-6.01E−07/-5.65E−02/=
cec129.89E−01/=6.80E−08/-6.80E−08/-1.06E−07/-6.80E−08/-6.80E−08/-1.43E−07/-6.80E−08/-7.04E−03/-6.80E−08/-1.23E−07/-
+/=/-1/2/90/1/110/1/110/0/120/0/120/0/120/1/110/0/120/1/110/2/100/2/10
Table 4. Comparison of algorithmic complexity between the proposed algorithm and other methods.
Table 4. Comparison of algorithmic complexity between the proposed algorithm and other methods.
MethodsAlgorithmic Complexity
YDSEO(Tmax×M×n+Tmax×n×d)
AHAO(Tmax×M×n+Tmax×n×d+Tmax×d/2)
SOO(Tmax×M×n+Tmax×n×d)
DOO(Tmax×M×n×d)
AOO(Tmax×M×n×d)
ChOAO((Tmax×n2/2+Tmax×n×d/2M)
HBAO(Tmax×M×n+Tmax×n×d)
SCSOO(Tmax×M×n×d)
SSAO(Tmax×M×n+Tmax×n×d)
WSOO(Tmax×M×n+Tmax×n×d)
COOTO(Tmax×M×n+Tmax×n×d)
FYDSEO(Tmax×M×n+Tmax×5n/2×d)
Table 5. Comparison of runtime of the proposed algorithm and other methods in cec2022 suite.
Table 5. Comparison of runtime of the proposed algorithm and other methods in cec2022 suite.
FunctionsYDSEAHASODOAOChOAHBASCSOSSAWSOCOOTFYDSE
cec012.8534.7382.7328.7046.28830.3533.60915.4232.0015.8732.8567.785
cec022.6284.4322.3858.2565.76530.0223.21115.1111.8465.3482.4457.355
cec033.4965.3563.3489.2277.60531.1124.07916.0392.6206.1093.38710.453
cec042.9404.6072.6738.5936.36330.2903.54015.3252.0475.2502.7048.306
cec053.0074.7532.7378.6746.55632.6473.54615.4352.1915.2562.7999.510
cec062.7224.4572.4818.3835.99936.0593.25215.1031.8515.0682.4867.456
cec074.0225.7693.7659.6708.52636.0194.60816.4023.1396.5763.90711.638
cec084.4836.1474.15010.0699.48934.5384.96616.9343.5246.9314.19113.199
cec093.7505.5283.5169.3528.00631.2704.34716.1612.8785.6163.57110.752
cec103.6225.2233.3489.2207.76332.9084.17915.9932.7965.9943.43110.250
cec114.3386.1324.06310.0229.25333.4544.86417.1643.5086.9394.14412.795
cec124.5196.3414.19410.1029.59632.0735.04016.9853.6306.4554.34813.165
Table 6. Probability results of reaching the optimal solution for twenty runs of FYDSE and other methods.
Table 6. Probability results of reaching the optimal solution for twenty runs of FYDSE and other methods.
FunctionsYDSEAHASODOAOChOAHBASCSOSSAWSOCOOTFYDSE
cec01100.0%100.0%95.0%100.0%0.0%0.0%100.0%0.0%100.0%30.0%100.0%100.0%
cec0255.0%65.0%60.0%35.0%5.0%0.0%25.0%25.0%10.0%65.0%35.0%100.0%
cec0395.0%95.0%95.0%85.0%5.0%0.0%100.0%10.0%50.0%90.0%80.0%100.0%
cec040.0%0.0%10.0%0.0%0.0%0.0%5.0%0.0%5.0%75.0%5.0%20.0%
cec05100.0%40.0%40.0%45.0%0.0%0.0%80.0%0.0%95.0%45.0%65.0%85.0%
cec06100.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%30.0%0.0%100.0%
cec075.0%95.0%10.0%0.0%0.0%0.0%15.0%0.0%0.0%65.0%10.0%85.0%
cec0870.0%100.0%90.0%10.0%15.0%0.0%55.0%10.0%0.0%15.0%15.0%100.0%
cec090.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%
cec100.0%0.0%0.0%5.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%45.0%
cec11100.0%70.0%70.0%45.0%45.0%0.0%90.0%15.0%70.0%60.0%90.0%100.0%
cec120.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%0.0%
Mean rank2.9173.0834.0005.0836.7507.5833.2506.8334.9174.0833.9171.333
Table 7. Results for design variables, gearbox weight, and running time for four-stage gearbox problem.
Table 7. Results for design variables, gearbox weight, and running time for four-stage gearbox problem.
FYDSEYDSEAHASODOAOChOAHBASCSOSSAWSOCOOT
x115.95917.01618.27718.58015.9689.50214.77117.94513.76523.41119.50919.801
x234.10033.36632.79742.31536.37717.75229.78153.29724.28340.98345.47345.714
x313.18414.94013.96317.62515.5069.08814.23123.06113.05519.38818.77221.534
x429.36535.59130.83935.35931.70622.60831.47547.14932.64241.21939.49943.259
x513.41718.20216.91418.35413.6069.46613.56832.58613.27527.79019.72126.928
x628.59532.50635.10637.15630.05023.47929.47048.44830.59644.35138.31149.284
x714.66821.52423.67521.03817.11810.05214.11232.26914.30724.00628.69423.742
x829.22634.97139.47039.59634.00628.33227.77744.80931.35749.31843.68346.500
x90.8452.6022.9531.3181.4582.8020.9281.4621.2541.7771.5371.448
x100.9832.5892.8721.5361.7152.4140.9711.4210.9961.6121.7871.606
x111.0132.8592.7401.2531.6672.6040.9111.4421.0501.6041.7411.405
x121.1892.4842.3291.2071.2982.4350.9031.5231.2681.2851.3661.533
x132.5325.9574.0604.5883.1454.4132.2684.5291.9654.9985.1434.669
x142.8595.2515.0474.7553.7914.5011.1874.9472.2605.9884.5255.310
x153.1655.0865.4024.7783.9725.2971.2025.0012.5575.1985.3965.142
x162.1475.1165.7624.4164.5265.5251.0905.0172.6264.6684.7595.615
x172.3845.2824.5024.6873.5684.3861.1625.4722.5375.0855.6925.274
x182.4614.4394.8174.9513.4114.0612.1835.2293.0014.3444.7334.306
x192.7705.5634.8764.5733.3945.0331.4095.2462.2775.2144.3984.756
x202.2484.9884.9954.2593.4644.4181.5144.7523.1815.0604.3024.716
x212.7424.9775.9374.3044.0954.7501.1954.9812.4965.6284.9454.941
x222.3125.3665.8334.2683.6724.4660.9544.4142.7364.1824.4765.081
Min14.73777.35668.08547.81758.58946.760107.321107.25541.50044.77250.43741.787
Max98.804197.889373.641135.408373.641350.620143.407192.523192.523280.607183.075224.306
Mean61.861136.076203.85290.482130.787126.405118.456134.148106.606112.777109.18197.105
Std23.35535.98190.23828.93180.96776.1666.80620.34041.65262.52034.81443.278
Time69.38267.93371.87671.27777.522141.335131.25274.92597.36767.87076.76869.814
Table 8. Optimization restraints for 20-rod truss structures.
Table 8. Optimization restraints for 20-rod truss structures.
VariablesParameter
Multiple loadConditions 1: F1 is 5 × 105 N, F2 is 0, Conditions 2: F1 is 0, F2 is 5 × 105 N.
Stressσimax is 173.43 MPa.
Displacement constraintsδ4ymax is 60 mm.
Natural frequency constraintsf1 ≥ 60 Hz and f2 ≥ 100 Hz.
Continuous cross-sectional area[Xlower, Xupper] is [−100, 100], and critical area is 1 cm2.
Material propertiesE is 6.9 × 1010 Pa and ρ is 2740 kg/m3.
Table 9. Optimization results for 20-rod truss structures.
Table 9. Optimization results for 20-rod truss structures.
FYDSEYDSEAHASODOAOChOAHBASCSOSSAWSOCOOT
1 (A1)1.54E+012.43E+011.50E+011.51E+011.52E+011.73E+013.66E+011.52E+011.44E+011.46E+011.81E+011.47E+01
2 (A2)-----2.10E+01--2.26E+00---
3 (A3)------------
4 (A4)-----2.28E+013.23E+01---1.28E+01-
5 (A5)2.02E+012.38E+011.98E+011.96E+011.95E+011.86E+01-2.00E+011.99E+011.99E+013.47E+012.04E+01
6 (A6)-----1.92E+01--8.89E+00-1.25E+00-
7 (A7)------------
8 (A8)2.03E+012.16E+012.02E+011.96E+012.03E+011.66E+013.36E+012.05E+011.93E+012.05E+012.06E+012.04E+01
9 (A9)-----6.06E+00--1.42E+01---
10 (A10)------------
11 (A11)2.00E+012.88E+012.00E+011.99E+012.05E+012.42E+01-1.97E+012.04E+012.13E+011.49E+011.96E+01
12 (A12)--------8.24E+00---
13 (A13)2.11E+012.16E+012.14E+012.19E+012.13E+012.51E+013.04E+012.13E+012.19E+012.12E+012.23E+012.09E+01
14 (A14)------------
15 (A15)2.10E+012.31E+012.07E+012.13E+012.08E+011.47E+014.65E+012.13E+012.01E+012.13E+012.05E+012.14E+01
16 (A16)-----2.06E+01------
17 (A17)------------
18 (A18)3.26E+014.33E+013.35E+013.35E+013.30E+013.27E+014.78E+013.27E+013.09E+013.32E+013.25E+013.28E+01
19 (A19)-----1.22E+01--4.35E+00---
20 (A20)3.25E+013.48E+013.26E+013.35E+013.27E+013.26E+01-3.29E+013.07E+013.31E+013.25E+013.30E+01
Best weight1.55E+021.80E+021.55E+021.56E+021.55E+022.28E+021.78E+021.55E+021.82E+021.55E+021.69E+021.55E+02
Mean weight1.65E+021.94E+021.76E+021.81E+021.84E+022.61E+021.78E+021.87E+022.11E+022.41E+022.05E+021.80E+02
Table 10. Optimization restraints for 24-rod truss structures.
Table 10. Optimization restraints for 24-rod truss structures.
VariablesParameter
Multiple loadConditions 1: F1 is 5 × 104 N, F2 is 0, Conditions 2: F1 is 0, F2 is 5 × 104 N.
lumped mass on nodes 3500 kg.
Stress and displacement constraintsσimax is 173.43 MPa, δ5y&6ymax is 10 mm.
Natural frequency constraintsf1 ≥ 30 Hz.
Continuous cross-sectional area[Xlower, Xupper] is [−40, 40] cm2, critical area is 1 cm2.
Material propertiesE is 6.9 × 1010 Pa and ρ is 2740 kg/m3.
Table 11. Optimization results for 24-rod truss structures.
Table 11. Optimization results for 24-rod truss structures.
FYDSEYDSEAHASODOAOChOAHBASCSOSSAWSOCOOT
1 (A1)-1.32E+011.22E+01---------
2 (A2)-5.13E+004.41E+00---------
3 (A3)------------
4 (A4)--------1.43E+00---
5 (A5)------------
6 (A6)-1.83E+011.08E+01-----2.39E+00---
7 (A7)2.13E+01--2.11E+012.18E+012.85E+012.23E+012.14E+012.15E+012.43E+011.91E+012.12E+01
8 (A8)2.66E+005.42E+001.34E+003.12E+001.02E+012.51E+005.36E+002.76E+003.36E+001.45E+013.17E+003.28E+00
9 (A9)-2.05E+00--1.63E+00-2.23E+00-----
10 (A10)1.37E+00--1.34E+00-1.47E+01-1.41E+001.42E+001.49E+00-1.47E+00
11 (A11)------------
12 (A12)----1.28E+013.47E+01----4.45E+00-
13 (A13)1.81E+012.16E+011.82E+011.86E+01--2.03E+011.79E+011.78E+011.75E+012.05E+011.77E+01
14 (A14)---1.37E+001.42E+01-------
15 (A15)3.84E+007.28E+007.08E+004.18E+004.83E+006.86E+007.54E+004.39E+004.09E+003.90E+004.24E+003.84E+00
16 (A16)2.42E+012.42E+012.43E+012.42E+012.41E+012.31E+012.43E+012.39E+012.38E+012.44E+012.44E+012.42E+01
17 (A17)----------1.37E+00-
18 (A18)------------
19 (A19)------------
20 (A20)------------
21 (A21)------------
22 (A22)-1.20E+01--7.22E+00-1.38E+00-2.26E+001.10E+00--
23 (A23)--3.09E+00------1.37E+003.36E+00-
24 (A24)1.42E+007.30E+002.50E+00-1.04E+001.51E+011.79E+00-1.17E+00--1.17E+00
Best weight1.26E+021.59E+021.31E+021.29E+021.35E+021.62E+021.38E+021.26E+021.36E+021.39E+021.35E+021.26E+02
Mean weight1.40E+021.84E+021.50E+021.54E+021.67E+022.03E+021.78E+021.79E+022.06E+021.89E+022.08E+021.47E+02
Table 12. Optimization restraints for 72-rod truss structures.
Table 12. Optimization restraints for 72-rod truss structures.
VariablesParameter
Multiple load conditionsConditions 1: F1x and F1y is 22.25 kN, F1z is −22.25 kN,
Conditions 2: F1z, F2z, F3z, and F4z is −22.25 kN.
Lumped mass on nodes 1, 2, 3, 42270 kg.
Stress and displacement constraintsσimax is 173.43 MPa, and δmax1x&1y&2x&2y&3x&3y&4x&4y is 6.35 mm.
Natural frequency constraintsf1 ≥ 4 Hz and f3 ≥ 6 Hz.
Continuous cross-sectional area[Xlower, Xupper] is [−30, 30], and critical area is 1 cm2.
Material propertiesE is 6.9 × 1010 Pa and ρ is 2767.99 kg/m3.
Table 13. Optimization results for 72-rod truss structures.
Table 13. Optimization results for 72-rod truss structures.
FYDSEYDSEAHASODOAOChOAHBASCSOSSAWSOCOOT
1 (A1A4)4.88E+006.37E+005.40E+004.60E+005.19E+007.49E+004.96E+004.86E+005.29E+002.54E+014.72E+006.23E+00
2 (A5A12)1.09E+011.10E+018.69E+001.12E+011.10E+018.69E+001.22E+011.15E+011.13E+011.21E+011.08E+011.08E+01
3 (A13A16)---------6.33E+00--
4 (A17A18)--8.79E+00--9.94E+001.00E+01-----
5 (A19A22)9.58E+001.09E+011.09E+011.14E+018.16E+008.98E+001.02E+011.05E+018.91E+002.39E+011.14E+017.82E+00
6 (A23A30)8.31E+007.88E+008.86E+007.93E+008.03E+008.06E+008.73E+009.33E+008.47E+001.19E+018.47E+008.14E+00
7 (A31A34)--3.07E+00-3.23E+00-3.87E+00-----
8 (A35A36)3.76E+003.69E+00-4.00E+005.70E+00--3.51E+003.76E+00-4.28E+004.36E+00
9 (A37A40)1.08E+017.73E+001.42E+011.13E+011.37E+019.69E+001.25E+018.64E+001.34E+015.48E+011.53E+011.12E+01
10 (A41A48)9.42E+008.50E+007.72E+007.77E+007.15E+007.96E+009.26E+007.64E+008.27E+001.33E+019.49E+007.88E+00
11 (A49A52)------------
12 (A53A54)------------
13 (A55A58)1.54E+012.33E+011.40E+011.46E+011.48E+011.82E+011.89E+011.82E+011.37E+015.74E+011.44E+011.64E+01
14 (A59A66)8.35E+008.94E+006.52E+009.50E+007.83E+008.17E+005.68E+007.54E+008.74E+001.20E+017.53E+009.03E+00
15 (A67A70)-----3.23E+00------
16 (A71A72)------------
Best weight4.50E+024.62E+024.55E+024.52E+024.52E+024.57E+024.90E+024.52E+024.50E+027.62E+024.52E+024.50E+02
Mean weight4.54E+024.97E+024.80E+024.58E+024.69E+024.68E+025.99E+024.70E+024.91E+021.30E+031.01E+034.60E+02
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

Qin, S.; Liu, J.; Bai, X.; Hu, G. Fractional-Order Boosted Hybrid Young’s Double-Slit Experimental Optimizer for Truss Topology Engineering Optimization. Biomimetics 2024, 9, 474. https://doi.org/10.3390/biomimetics9080474

AMA Style

Qin S, Liu J, Bai X, Hu G. Fractional-Order Boosted Hybrid Young’s Double-Slit Experimental Optimizer for Truss Topology Engineering Optimization. Biomimetics. 2024; 9(8):474. https://doi.org/10.3390/biomimetics9080474

Chicago/Turabian Style

Qin, Song, Junling Liu, Xiaobo Bai, and Gang Hu. 2024. "Fractional-Order Boosted Hybrid Young’s Double-Slit Experimental Optimizer for Truss Topology Engineering Optimization" Biomimetics 9, no. 8: 474. https://doi.org/10.3390/biomimetics9080474

Article Metrics

Back to TopTop