Next Article in Journal
Comprehensive Analysis of HY-2B/2C/2D Satellite-Borne GPS Data Quality and Reduced-Dynamic Precise Orbit Determination
Previous Article in Journal
A Comparative Analysis of Multi-Criteria Decision-Making Methods and Normalization Techniques in Holistic Sustainability Assessment for Engineering Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comparison of Modern Metaheuristics for Multi-Objective Optimization of Transonic Aeroelasticity in a Tow-Steered Composite Wing

1
Sustainable Infrastructure Research and Development Center (SIRDC), Khon Kaen University, Khon Kaen 42000, Thailand
2
Department of Mechanical Engineering, Faculty of Engineering, Mahasarakham University, Maha Sarakham 44150, Thailand
*
Author to whom correspondence should be addressed.
Aerospace 2025, 12(2), 101; https://doi.org/10.3390/aerospace12020101
Submission received: 25 December 2024 / Revised: 18 January 2025 / Accepted: 22 January 2025 / Published: 30 January 2025

Abstract

:
This study proposes a design procedure for the multi-objective aeroelastic optimization of a tow-steered composite wing structure that operates at transonic speed. The aerodynamic influence coefficient matrix is generated using the doublet lattice method, with the steady-state component further refined through high-fidelity computational fluid dynamics (CFD) analysis to enhance accuracy in transonic conditions. Finite element analysis (FEA) is used to perform structural analysis. A multi-objective transonic aeroelastic optimization problem is formulated for the tow-steered composite wing structure, where the objective functions are designed for mass and critical speed, and the design constraints include structural and aeroelastic limits. A comparative analysis of eight state-of-the-art algorithms is conducted to evaluate their performance in solving this problem. Among them, the Multi-Objective Multi-Verse Optimization (MOMVO) algorithm stands out, demonstrating superior performance and achieving the best results in the aeroelastic optimization task.

1. Introduction

Improving aircraft design technology for performance, limitations, and sustainability is increasingly required. These are important to reach new frontiers in commercial aircraft [1,2,3,4,5] and unmanned aerial vehicle (UAV) development [6,7]. To manufacture these aircraft, ensuring aircraft design is in accordance with aviation standards and regulations, such as those of the Federal Aviation Administration (FAA) and European Union Aviation Safety Agency (EASA) [8,9], is essential. Preliminary aircraft design is crucial because it is the center of the design process. Conceptual and detailed design can designate and actualize important aircraft components. At this design stage, aeroelasticity analysis is vital in achieving reliable design and has rudimentary design requirements. Aeroelastic optimization is essential for minimizing mass and construction costs while maximizing safety. Since most commercial transport aircraft operate at transonic speeds, aeroelastic optimization in this regime is essential.
Aeroelasticity has multiple mechanics applications that require understanding the interaction between elastic solid bodies and fluid. Aeroelastic analysis considers interactions of inertial, elastic, and aerodynamic forces in aircraft design. Both aerodynamic and structural forces are considered to analyze each interaction. Aerodynamic analysis is commonly conducted using panel methods that include medium-fidelity and classical techniques. The Vortex Lattice Method (VLM) [10,11,12,13,14,15] and the Doublet Lattice Method (DLM) [16,17,18,19] are especially important. Other methods are based on the lifting-line theory [20,21]. These aerodynamic methods can perform calculations rapidly; however, they cannot capture complex phenomena such as flow separation, oscillation, and shock, which commonly occur at transonic speeds. Therefore, a higher-fidelity method is needed. CFD [22] is usually employed to model complex problems and accurately capture aerodynamic phenomena, but it is highly time-consuming. Currently, both classical and CFD methods have been employed, and they have improved aerodynamic analysis for aircraft design [14,15,23,24]. Structural aircraft analysis is regularly performed with FEA, a classical structural engineering approach. Numerous aeroelastic analyses are required in the optimization process. For this reason, using CFD as an aerodynamic solver may be too time-consuming. Previous studies [25,26,27,28] demonstrated a correction technique that uses CFD results to improve DLM accuracy at transonic speeds.
Composite materials are commonly used to reduce the mass of aircraft. Normally, they are based on carbon or fiberglass. Not only does this material reduce mass, but tailoring the sequence stacking and orientation of the material layers can produce a designed strength that can withstand expected loads. The layout of the composite needs to adapt to the direction of stress. Conventionally, the direction of fibers in a ply is likely straight, parallel, or in a specified orientation in layered structures. The tow-steered technique is continuously applied to lay fiber with varying fiber orientations to address the stress field. This can be successful by employing automated fiber-placement technology. Presently, the design and optimization of composite tow steering involving aeroelastic analysis [29,30,31,32,33,34] have primarily focused on predefined layering configurations during manufacturing. These studies have contributed to innovative aircraft design utilizing composite materials.
Optimization has been performed in engineering problems to more rapidly determine optimal solutions than can be experimentally achieved, especially with aircraft [35,36,37]. These optimization methods often employ gradient-based [38,39] and stochastic-based methods [40,41]. Gradient-based and rapidly converging methods are normally applied, while stochastic-based methods use randomization to generate solutions that converge slowly. However, real-world problems may not be convex, so the gradients of objective and constraint functions may be unacceptable because of discrete variables. Likewise, the design variables are discrete in aeroelastic optimization for aircraft design. Stochastic-based methods are appropriate for handling discrete and continuous variables. Furthermore, non-gradient problems can be solved. This development requires metaheuristics (MHs) to be implemented for performance improvement in seeking optimal solutions to higher-complexity problems. Recently developed MHs can be categorized by their algorithms and classified into several groups. Evolutionary algorithms [42,43,44,45,46,47,48] are the most common MH type. Other forms, like swarm intelligence [49,50,51,52,53,54], mathematical and physical-inspired algorithms [55,56,57,58], and human society-inspired algorithms [59,60,61,62], are further MH types. The use of MHs has been studied in a variety of applications where the problem definitions can be optimized, typically using single and multiple objectives. MH algorithms are inspired by different techniques, but they are primarily focused on population-based algorithms, and their overall procedures are identical. Generally, initial solutions, called populations, are randomly generated in the optimization procedure. Afterward, the population reproduction becomes iterative by regulating the reproducing operators, which define the algorithm and motivation. In every iteration, the iterative population proceeds with domination sorting techniques to evaluate the non-dominated population. The calculations are terminated when the maximum number of function evaluations is reached. Non-dominated solutions can be structured to support the optimal Pareto front.
There are many recent studies in aeroelastic optimization employing MH for formulating and solving wing optimization problems [63,64,65,66,67,68,69,70,71,72,73,74,75,76,77]. Nevertheless, outdated MH algorithms like Genetic Algorithms (GA) [68,71,78,79], Particle Swarm Optimization (PSO) [68,80,81,82], and the Non-dominated Sorting Genetic Algorithm (NSGA-II) [83,84,85] have been used to solve these problems. Several studies utilized more recent optimization algorithms [67] and developed new algorithms [66]. Most aircraft optimization studies focus on designing optimal wings using isotropic or composite materials at low speeds. Studies have been performed examining aeroelastic optimization of a composite wing at transonic speeds [67], but none account for tow-steering. Recent works have explored tow-steered wing design [68,69,86,87,88,89,90,91,92,93]. In these studies, the primary design variables include laminate thickness and orientation, which are used to define the tow-steering pattern [92]. The primary objective of the optimization problems addressed in these studies is to minimize mass, while some have also incorporated multi-objective formulations. Typical constraints include allowable stresses, critical aeroelastic effects (such as divergence, flutter, and gust responses [68,69,87,91]), and mechanical failure modes (such as buckling [68,88,89] and fatigue [68]). However, the optimization of composite wings with tow steering at designed transonic speeds remains unexplored. Additionally, many existing studies rely on outdated optimization algorithms.
To address these gaps, our study aims to investigate the aeroelastic optimization of a tow-steered composite wing at transonic speed, utilizing aerodynamic transonic corrections from [28]. MHs offer superior global search capabilities compared to gradient-based methods, but their slower convergence is a drawback. To mitigate this, aerodynamic transonic correction is applied to enhance the aerodynamic influence coefficient (AIC) from the DLM using CFD results, as detailed in [67]. The contributions of this study are twofold:
  • The development of a design procedure for tow-steered aeroelastic optimization at transonic speed.
  • A comparative performance study of recent MHs in tow-steered aeroelastic optimization at transonic speed.
The research presented here introduces a novel composite wing design procedure and evaluates the optimization of performance for solving these challenges. This design procedure can be further developed, making it more practical and applicable for aircraft design, offering a foundation for future researchers to enhance its utilization. The comparative analysis of algorithms in this study demonstrates their effectiveness in solving the formulated problems, leading to the development of new algorithms and advancements in metaheuristic optimization.
A problem formulation and design variable encoding procedure specifically for tow-steering is detailed in Section 2, while the numerical test settings and experimental setup are presented in Section 3. The results and discussion of the study are given in Section 4, and study conclusions are offered in Section 5.

2. Aeroelastic Design Problem

Multiobjective optimization aims to simultaneously identify all possible solutions for a given set of scenarios. In this study, two objectives, mass and critical speed, are considered. The former is a cost indicator, while the latter reflects reliability. They tend to conflict. Thus, multiobjective design is a tool to explore all possible optimal solutions. Notably, all the designs are feasible, but they have varying levels of construction costs and reliability.
The practical benefit of obtaining these trade-off solutions is that designers can visualize all possible options related to the objective functions during the design process. This allows them to freely select a design for construction using decision-making techniques, such as a weight rating evaluation method. Moreover, selecting multiple designs for different scenarios is also viable, as all possible solutions are readily available.
In this work, recent and well-known multi-objective metaheuristic optimization algorithms are employed to solve the aeroelastic optimization problem of a tow-steered Common Research Model (CRM) wing [94,95,96,97]. The wing has a half-span of 26.23 m. Structural and aerodynamic analyses are performed using FEA and the DLM with AIC correction [28,67], respectively. The CFD results of this CRM wing model are from [96] and used to improve the aerodynamic AIC. Mindlin shell elements with a drilling degree of freedom [98] and wrapped correction [32] are applied. The transverse shear elasticity matrix for preventing a shear-locking phenomenon is computed based on [99,100,101]. The structural components, including the front spar, rear spar, and all ribs, are made of an isotropic material (Aluminum 2024-T3 [102]). Both the upper and lower skins consist of five layers of composite material (carbon fiber) [103]. The mechanical properties of the structural and skin components are presented in Table 1.
The wing FEA model used for structural analysis is shown in Figure 1. Since this is a preliminary design phase, a simplified wing mesh is used to reduce computational effort. The design variables for tow-steered layup orientation are presented in Figure 2 to address variations in carbon fiber orientation. Two design variables, root ( θ i n ) and tip ( θ t i p or θ o u t ) angles, are assigned. The angles of carbon fiber are linearly distributed along the spanwise direction. The ply orientation of each element can be computed using Algorithm 1.
Algorithm 1. Calculation of orientation distribution of a tow-steered wing.
Input: orientations at the root chord ( O r i   r o o t ), orientations at the tip chord ( O r i   t i p ), FEA element data ( e l e m e n t )
 
    - reshape the FEA element data ( e l e m e n t ) into groups of span elements ( e l e   s p a n )
    - get the total span elements ( n e l e   s p a n ), minimum span location ( y   m i n ), maximum span location ( y   m a x )
       f o r   i 1 = 1 : 1 : n e l e s p a n
        Compute span location ( y )
O r i @ s p a n i 1 = O r i r o o t + O r i t i p O r i r o o t ( y y m i n y m a x y m i n )
       e n d
 
Output:  O r i @ s p a n
The CRM wing is simulated at transonic speed at a 0.85 Mach number (corresponding to a flight speed of 251.26 m/s). The wing is analyzed at a 2° angle of attack (AOA) and a 0.0155272 kg/m3 air density. The reference length of the wing is 7.0053 m. In this proposed problem, mass and critical speed are objective functions. A lower mass results in a lighter structure and reduced construction costs, while a higher critical speed contributes to a more reliable and safer design. The critical speed is defined as the lower value of the divergence and flutter speeds, representing static and dynamic aeroelastic phenomena, respectively. The goal of this multi-objective optimization is to obtain the optimal solutions or an optimal Pareto front that demonstrates a trade-off between construction costs and safety. The aeroelastic multi-objective optimization problem for the CRM wing is as follows:
M i n i m i z e : f x = [ f 1 x ,   f 2 x ]
f 1 x = t o t a l   w i n g   m a s s   W ,   f 2 x = C r i t i c a l   s p e e d   ( V c r )
subject to
Lift effectiveness: 0.8 η L 1.2 ;
Critical speed: V c r 1.2 × V ;
Displacement: u m a x 10 % of half span;
Stress: σ v o n ,   m a x   σ y (isotropic material);
T H   m a x     1 (composite material);
Buckling: α b f     1 ;
Boundary: x L   x   x U .
where
σ v o n = σ 1 2 σ 1 σ 2 + σ 2 2 T H = σ 11 X 11 2 + σ 11 σ 22 X 11 2 σ 22 X 22 2 τ 12 S 12 2 η L = L F L R = q S [ A I C ] F { α i } F L R x   is   a   vector   of   design   variables .
The aeroelastic problem has a total of 46 design variables, including the thicknesses and orientations of wing components, as listed below:
x 1 =The root thickness of the front spar (spar No. 1);
x 2 =The root thickness of the rear spar (spar No. 2);
x 3 =The root thickness of the ribs;
x 4 =The tip thickness ratio of the front spar (spar no. 1);
x 5 =The tip thickness ratio of the rear spar (spar No. 2);
x 6 =The tip thickness ratio of the ribs;
x 7 11 =The thicknesses at the root chord of the upper skin (the inner layer to the outer layer);
x 12 16 =The thicknesses ratio at the tip chord of the upper skin (the inner layer to the outer layer);
x 17 21 =Orientations at the root chord of the upper skin (the inner layer to the outer layer);
x 22 26 =Orientations at the tip chord of the upper skin (the inner layer to the outer layer);
x 27 31 =The thicknesses at the root chord of the lower skin (the inner layer to the outer layer);
x 32 36 =The thicknesses ratio at the tip chord of the lower skin (the inner layer to the outer layer);
x 37 41 =Orientations at the root chord of the lower skin (the inner layer to the outer layer);
x 42 46 =Orientations at the tip chord of the lower skin (the inner layer to the outer layer).
All thicknesses are linearly distributed, with the maximum thickness at the wing root, decreasing in the spanwise direction, and reaching a minimum at the wing tip. The thickness ratio represents the tip thickness to the root thickness ratio. Most design variables are discrete. The thicknesses of the composite material are selected from the set {0.25, 0.5, 0.75, 1.0, …, 9.75, 10.0} mm, while the list of ply orientations is {0, 5, 10, …, 170, 175} degrees. The ribs and spars are made of an isotropic material, with thicknesses chosen from the set {0.25, 0.5, 0.75, …, 4.75, 5.0, 6.0, 7.0, …, 19, 20, 25, …, 45.0, 50.0} mm.
The function evaluation process for the aeroelastic wing optimization problem is outlined in Algorithms 1 and 2. Since the wing shape remains unchanged during optimization, the AIC matrix is pre-calculated using the DLM. Design variables containing all structural data are then input to perform FEA. Displacements, stresses, and buckling factors are obtained at this stage. Then, aeroelastic analysis is conducted to determine the lift effectiveness and critical speed.
Algorithm 2. The tow-steered CRM wing aeroelastic function evaluation
Input: design variable ( x )
 
    - Load the aerodynamic results data including
    - Transform the design variable into the real geometric values including skin, rib and spar thicknesses and orientation of tow steer composite wing skin (detailed in Algorithm 1).
    - Create the finite element model, assemble the stiffness matrix and mass matrix
    - Solve the free vibration analysis and static analysis:
        -     Free vibration analysis results including mode shape, lambda, omega.
        -     Static analysis results including the structural deformation ( u m a x ), Tsai-Hill criteria stress ( T H   m a x ) and Von-Mises stress ( σ v o n ,   m a x ).
    - Solve the static aeroelastic: obtain the lift effectiveness ( η L ) and divergence speed ( V d )
    - Solve for the flutter speed ( V f ) using v-g-method
    - Compute the critical speed:
V c r = m i n ( V d , V f )
    - Perform the objective function and constraints
        -      f x = f 1 x , f 2 x , f 1 x = t o t a l   w i n g   m a s s ,
        -      f 2 x = C r i t i c a l   s p e e d
      Subject to
        -      g 1 x : u m a x 2.623 m (10% of half span)
        -      g 2 x : σ v o n , m a x 410 MPa (isotropic material)
        -      g 3 x : T H m a x 1 (composite material)
        -      g 4 x :   α   b f     1 (Buckling factor constraint)
        -      g 5 x :   V c r   301.512 m/s (1.2 times flight speed)
        -      g 6 x : 0.8 η L 1.2
        -      x L x x U
 
Output:   f x , g i x

3. Numerical Setup

In the aeroelastic analysis procedure, the FEA mesh is completely fixed in both static and dynamic analyses. A grid-independence test is conducted. In this study, we deliberately chose a coarser mesh over a fully converged one since this research centers on crafting a design approach specialized for enhancing the tow-steered composite wing structures. Thus, achieving absolute precision in analytical outcomes is not the primary aim.
The integration of the function evaluation and an optimization algorithm is the key to this study. The common procedure for multi-objective metaheuristics (MHs) is illustrated in Figure 3. At the beginning of the optimization process, a population or a set of design solutions is randomly generated, while the Pareto archive, a matrix for collecting non-dominated solutions, is generated. Then, a new population is iteratively derived using reproduction operators, which vary between algorithms. Subsequently, function evaluations are performed. The members in the Pareto archive and those in a new population are sorted using a non-dominated sorting technique to identify non-dominated solutions and saved to the Pareto archive.
When the archive size is too large, a clustering technique is employed to remove some solutions to conserve computer memory. This clustering technique preserves the diversity of the members in the archive while addressing memory constraints. The updated Pareto archive at the last iteration is regarded as the Pareto set of optimal solutions. An equal number of function evaluations is used to terminate all algorithms to ensure that this comparison is fair.
During the optimization process, the wing shape remains fixed. The FEA mesh, aerodynamic loadings, and AIC are pre-calculated and imported for structural and aeroelastic analysis. Aerodynamic loadings are derived using the VLM, while the AIC is constructed using the DLM with steady-state corrections based on transonic CFD results.
Each reproduced population contains solution vectors used during function evaluations. These vectors contain structural data such as material thicknesses and orientations. The structural parameters, combined with the imported FEA mesh, are used to generate the FEA model. Pre-calculated aerodynamic loadings are then applied to perform structural static analysis. Subsequently, aeroelastic analysis is conducted using the FEA model and the imported AIC to determine divergence and flutter speeds. The critical speed is the lower of these two values.
The performance of the state-of-the-art algorithms on multi-objective aeroelastic optimization of the CRM composite wing is investigated. They are evaluated based on the convergence rate, consistency, and versatility. The hypervolume (HV), which can measure both the spread and advancement of a Pareto front, is used as a performance indicator in this study [104,105]. The statistical parameters of the HV, including the average, standard deviation, maximum, minimum, and rank of average, are evaluated for performance comparison. The average HV, maximum, and rank are used as convergence rate indicators, while the standard deviation and minimum HV are consistency indicators. Higher values of average, maximum, and minimum HV are better, while lower values of standard deviation and rank are more favorable. Ten independent runs of each algorithm are evaluated in the aeroelastic optimization problem. The population size and maximum function evaluations are 100 and 20,000, respectively. The parameter settings of all algorithms are detailed in Table 2.
In the aeroelastic analysis procedure, the FEA mesh is fully fixed for both static and dynamic analyses. A grid independence test is conducted to ensure accuracy. In this study, we intentionally selected a coarser mesh rather than a fully converged one because our primary focus is on developing a design methodology to enhance tow-steered composite wing structures. So, achieving absolute precision in analytical results is not the primary objective. The performance of state-of-the-art algorithms for multi-objective aeroelastic optimization of the CRM composite wing is investigated. Algorithms are evaluated based on the convergence rate and consistency. Hypervolume (HV), which quantifies both the spread and advancement of the Pareto front, is used as a performance metric. Statistical results, including the average, standard deviation, maximum, minimum, and rank of HV, are analyzed for comparative assessment.
The average HV, maximum HV, and rank indicate the convergence rate, while the standard deviation and minimum HV reflect consistency. Higher average and maximum HV values, along with a lower rank, indicate better performance. Lower standard deviation and higher minimum HV values signify greater consistency and robustness. Each algorithm is evaluated over 10 independent runs for the aeroelastic optimization problem. The population size is uniformly set to 100 across all algorithms. Each algorithm is subjected to 20,000 function evaluations to ensure a fair comparison. Parameter settings, adopted from pre-tuned configurations specified in earlier research, are detailed in Table 2. These settings represent optimal values tested across various problems in the respective studies.

4. Results and Discussion

In this section, the optimum solutions from several algorithms are presented and discussed. HV is used as an indicator of the diversity and spread of the Pareto solution. Figure 4 presents HV convergence plots for all algorithms, comparing the average hypervolume results from all optimization runs. MOMVO outperforms all other algorithms, while MOGOA, MOMA, and MSSA are the next best. Early in the optimization, MOGOA demonstrates the fastest convergence compared to other approaches. However, its HV shows minimal improvement during the second half of the optimization runs. This suggests that MOGOA excels in exploitation but suffers from poor exploration, leading to premature convergence at local optima. In contrast, MOMVO exhibits outstanding performance in the latter half of the optimization, with its rapidly increasing hypervolume. By the end of the process, it achieves nearly double the HV of the worst-performing algorithm, MOCryStAl. These results highlight MOMVO as the best-performing algorithm, striking a well-balanced trade-off between exploitation and exploration. Another notable observation from the convergence plots is the inconsistent convergence pattern of MOMA. This inconsistency suggests that the solutions provided by MOMA cannot be considered optimal, as the algorithm did not fully converge.
A comparison of the best Pareto fronts for this test problem is shown in Figure 5, which corresponds to the best HV results in Table 3. The best Pareto front in this study is obtained by MOMVO. Upon examining the Pareto fronts in Figure 5, the results can be categorized into two groups based on the critical speed. The first group includes Pareto fronts with critical speeds ranging from 500 m/s to approximately 600 m/s, while the second group features critical speeds greater than 700 m/s. However, regardless of the group, the Pareto front obtained by MOMVO exhibits significantly greater spread and advancement compared to all other examined approaches. Considering all aspects of the results, MOMVO is the best-performing algorithm for this problem, while MOGOA and MOMA provide comparable results and show potential for further improvement.
Statistical HV results are presented in Table 3 and summarized as a box plot in Figure 6 for a deeper analysis of performance in terms of both speed and consistency. In terms of speed, MOMVO achieved the best average HV and ranking, as well as the highest hypervolume in this study, confirming its position as the fastest algorithm. MOJS achieved the lowest standard deviation, while MOMA delivered the best results in the worst-case scenario. The poorest MOMA run still had the greatest hypervolume, slightly better than the worst run of MOMVO. Although MOJS exhibited a lower standard deviation, both its average hypervolume and that of MOMA were significantly lower than those of MOMVO. When comparing the average hypervolumes, MOMVO, MOJS, and MOMA have standard deviations of 11.9, 5.9, and 10.4%, respectively. So, MOJS demonstrates the best consistency, while MOMVO and MOMA are comparable. However, in the worst-case scenario, algorithms such as MOGOA, MOMRFO, and MSSA performed poorly, with a minimum hypervolume value of zero, indicating their failure to capture any feasible solutions for the test problem.
Overall, MOMVO remains the best algorithm based on these statistical results, as it offers the fastest speed and is one of the most consistent algorithms in this study. However, it is noteworthy that MOGOA and MOMA produce comparable results. Based on the observations in this study, the performance of these algorithms can be enhanced in future research.
MOGOA exhibits rapid convergence during the early stages of the optimization, indicating strong local search capabilities. However, its poor global search capability causes convergence at local optima. This algorithm could be improved by enhancing its global search capabilities. Potential improvements include introducing additional reproduction operators to increase population diversity or integrating adaptive strategies to dynamically readjust reproduction parameters during optimization.
MOMA demonstrates comparable performance to MOGOA. It maintains a good balance between exploration and exploitation, as it continuously improves the hypervolume throughout the optimization. However, MOMA encounters an issue in its clustering process. When the number of non-dominated solutions exceeds the maximum Pareto archive limit, a clustering technique is used to eliminate excessive solutions. If the algorithm employs an effective clustering technique, the hypervolume should consistently increase rather than fluctuate, as observed in the hypervolume history of MOMA. Its performance could be enhanced by addressing this issue.
The details of the selected solutions from the best Pareto front, highlighted in Figure 7, are provided in Table 4 and Table 5. Table 4 presents the values of all objective and constraint functions for the selected solutions, while Table 5 provides the corresponding design variable values. These details are provided to demonstrate the trade-off between mass and critical speed. In Table 5 and Figure 8, the first solution has less than half the mass of the third, but it has a significantly lower critical speed. The first solution is lighter but less reliable than the third solution, while the second solution is intermediate.
All selected solutions are feasible, as all constraints—including displacement, von Mises stress (for aluminum), Tsai–Hill (for carbon fiber), buckling factor, critical speed, and lift effectiveness—are within the allowable ranges. Additionally, the trend shows that overall thickness increases with greater mass and critical speed. This trade-off reflects the goal of multi-objective design optimization. Thicker structures may provide greater safety but require higher manufacturing costs.
In Table 5 and Figure 8, the details of the thicknesses and orientations of the selected solutions are displayed. Notably, the color scale in each sub-figure is different, as each solution thickness varies considerably. The solutions, from the first to third, range from the lightest to the heaviest. The optimum orientation of each solution is not perfectly matched. However, the orientations of each solution show a similar trend in more than half of the layers. This suggests that the optimum orientation may vary slightly depending on differences in ply thickness.
A comparison of composite wing aeroelastic results [67] with tow-steered aeroelastic results is presented using the best Pareto fronts from both studies, as shown in Figure 9. This figure highlights three solutions: (1) the tow-steered composite wing from the current study with 20,000 FEs, (2) the tow-steered composite wing from the current study with 5000 FEs, and (3) the Pareto front of composite wing aeroelastic results from previous studies [67] with 5000 FEs. The Pareto front at 5000 FEs with tow-steer design is more advanced than that obtained from constant ply orientation. The Pareto front at 20,000 FEs of the two-steer design dominates that of a previous study.
Since all optimal trade-off solutions have been obtained, any design can be selected for construction using a decision-making technique, such as a weight rating evaluation method. This selection can be made by designers or customers. Moreover, a comparison between the original transonic design problem presented in previous work [67] and the current study indicates significant improvements, as illustrated in Figure 9. The tow-steered composite wing design solution demonstrates superior performance to the original problem. As expected, more optimal solutions are achievable from the tow-steered design concept, which provides greater design diversity.

5. Conclusions

A design procedure for tow-steered aeroelastic optimization is developed in the current study. A comparative analysis of recent multi-objective algorithms for the multi-objective aeroelastic optimization of a tow-steered composite wing is conducted. These recent algorithms, which are rarely used in the aerospace field, are employed to solve a complex, real-world aircraft design optimization problem. Their performance and consistency are investigated.
Overall, MOMVO is considered the best algorithm in this study, with MOGOA, MOJS, and MOMA having slightly lower performance. MOMVO is the fastest algorithm, while MOGOA ranks second in this aspect. MOJS is the most consistent, slightly outperforming MOMVO and MOMA. MOMA performs best in the worst-case scenario, slightly better than MOMVO. While MOGOA demonstrates good speed, it fails in the worst-case scenario as it cannot obtain feasible solutions in some cases.
MOMVO is the best-performing algorithm in this study, with outstanding speed and best results in both consistency and the worst-case scenario. Optimal solutions showcasing the trade-off have been identified, allowing designers or customers to select a design for construction using decision-making techniques. Additionally, the tow-steered composite wing design from this study outperforms the original transonic design problem [67], significantly increasing the potential for achieving optimal solutions.
For future research, we recommend three key steps: (1) incorporating nonlinearity effects for high aspect ratio wings and considering additional flight conditions, for example, the maneuver load and gust load cases, among others; (2) dealing with randomness in linear/nonlinear computational simulation and manufacturing employing automated fiber placement and material properties using reliability optimization; (3) developing a faster optimizer by integrating machine learning or deep learning models with metaheuristic algorithms to address more detailed problems; and (4) conducting further experiments to confirm the reliability of the optimized designs. Ground vibration tests and updates to FEA models can be experimentally performed to apply the optimized design to a physical wing in future work.

Author Contributions

Conceptualization, K.W., N.P. (Natee Panagant) and S.B.; methodology, K.W., N.P. (Natee Panagant) and S.B.; software, K.W., N.P. (Natee Panagant), and S.B.; validation, K.W., and N.P.(Natee Panagant); formal analysis, K.W., N.P. (Natee Panagant), and S.B.; investigation, K.W., N.P. (Natee Panagant), and S.B.; resources, N.P. (Natee Panagant) and S.B.; data curation, R.K.; writing—original draft preparation, K.W., K.P., R.K., and N.P. (Natee Panagant); writing—review and editing, K.W., K.P., S.B., N.P. (Nantiwat Pholdee), and N.P. (Natee Panagant); visualization, N.P. (Nantiwat Pholdee) and S.B.; supervision, S.B.; project administration, S.B.; funding acquisition, N.P. (Nantiwat Pholdee) and S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work (Grant No. RGNS 65-140) was financially supported by Office of the Permanent Secretary, Ministry of Higher Education, Science, Research and Innovation.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors upon request.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
MOMVOMultiobjective Multi-Verse Optimization.
MOGOAMultiobjective Grasshopper Optimization Algorithm.
MOLAMultiobjective Lichtenberg Algorithm.
MOJSMultiobjective Jellyfish Search.
MOCryStAlMultiobjective Crystal Structure Algorithm.
MOMRFOMultiobjective Manta Ray Foraging Optimizer.
η L Lift effectiveness.
L F The actual lift force at flight speed operation.
L R The lift force at ground state shape with flight speed operation.
V Flight speed.
u Transverse displacement vector of the wing.
u m a x Maximum transverse displacement of the wing.
σ 1 Stress in direction 1 of isotropic elements.
σ 2 Stress in direction 2 of isotropic elements.
σ v o n The Von Mises stress of isotropic elements.
σ v o n ,   m a x The maximum Von Mises stress of isotropic elements.
σ y The yield stress of isotropic elements.
σ 11 Stress in direction 11 of composite elements.
σ 22 Stress in direction 22 of composite elements.
σ 12 Stress in direction 12 of composite elements.
X 11 Transverse displacement in direction 11 of composite elements.
X 22 Transverse displacement in direction 22 of composite elements.
X 12 Transverse displacement in direction 12 of composite elements.
T H   m a x Maximum Tsai–Hill failure index of composite elements.
α b f Buckling factor.
x A vector of design variables.
x L The lower bounds of design variables.
x U The upper bounds of design variables.

References

  1. Mcdonald, R.A.; German, B.J.; Takahashi, T.; Bil, C.; Anemaat, W.; Chaput, A.; Vos, R.; Harrison, N. Future aircraft concepts and design methods. Aeronaut. J. 2022, 126, 92–124. [Google Scholar] [CrossRef]
  2. Sarlioglu, B.; Morris, C.T. More Electric Aircraft: Review, Challenges, and Opportunities for Commercial Transport Aircraft. IEEE Trans. Transp. Electrif. 2015, 1, 54–64. [Google Scholar] [CrossRef]
  3. Ma, Y.; Elham, A. Designing high aspect ratio wings: A review of concepts and approaches. Prog. Aerosp. Sci. 2024, 145, 100983. [Google Scholar] [CrossRef]
  4. Liu, Y.; Zhang, S.; Zhang, J.; Yao, K.; Luo, M.; Liu, Y. Towards the Design of Future Aircraft: A Critical Review on the Tools and Methodologies. Aerosp. Res. Commun. 2024, 2, 13096. [Google Scholar] [CrossRef]
  5. Boden, F.; Barth, H.P.; Hein, S.; Heitmann, D.; Kirmse, T.; Klein, C.; Krüger, W.; Lemarechal, J.; Norambuena, M.; Schäfer, D.; et al. Recent developments of common numerical methods and common experimental means within the framework of the large passenger aircraft program. CEAS Aeronaut. J. 2023, 14, 281–293. [Google Scholar] [CrossRef]
  6. Elmeseiry, N.; Alshaer, N.; Ismail, T. A Detailed Survey and Future Directions of Unmanned Aerial Vehicles (UAVs) with Potential Applications. Aerospace 2021, 8, 363. [Google Scholar] [CrossRef]
  7. Rabiu, L.; Ahmad, A.; Gohari, A. Advancements of Unmanned Aerial Vehicle Technology in the Realm of Applied Sciences and Engineering: A Review. J. Adv. Res. Appl. Sci. Eng. Technol. 2024, 40, 74–95. [Google Scholar] [CrossRef]
  8. Missili, A.; Mirabella, C.; Tuccillo, M.; Della Vecchia, P. A Model-Based Systems Engineering Digital Certification Framework for General Aviation Aircraft. Aerotec. Missili Spaz. 2024. [Google Scholar] [CrossRef]
  9. Cook, S.P. Digital Airworthiness Certification: Opportunities for Unmanned Aircraft Systems. In Proceedings of the 2024 International Conference on Unmanned Aircraft Systems (ICUAS), Chania, Greece, 4–7 June 2024; pp. 1432–1437. [Google Scholar]
  10. Lambert, T.; Razak, N.A.; Dimitriadis, G. Vortex Lattice Simulations of Attached and Separated Flows around Flapping Wings. Aerospace 2017, 4, 22. [Google Scholar] [CrossRef]
  11. Parenteau, M.; Laurendeau, E. A general modal frequency-domain vortex lattice method for aeroelastic analyses. J. Fluids Struct. 2020, 99, 103146. [Google Scholar] [CrossRef]
  12. Hang, X.; Su, W.; Fei, Q.; Jiang, D. Analytical sensitivity analysis of flexible aircraft with the unsteady vortex-lattice aerodynamic theory. Aerosp. Sci. Technol. 2020, 99, 105612. [Google Scholar] [CrossRef]
  13. Joshi, H.; Thomas, P. Review of vortex lattice method for supersonic aircraft design. Aeronaut. J. 2023, 127, 1869–1903. [Google Scholar] [CrossRef]
  14. Dagilis, M.; Kilikevičius, S. Aeroelasticity Model for Highly Flexible Aircraft Based on the Vortex Lattice Method. Aerospace 2023, 10, 801. [Google Scholar] [CrossRef]
  15. Gulizzi, V.; Benedetti, I. Computational aeroelastic analysis of wings based on the structural discontinuous Galerkin and aerodynamic vortex lattice methods. Aerosp. Sci. Technol. 2024, 144, 108808. [Google Scholar] [CrossRef]
  16. Albano, E.; Hodden, W.P. A doublet-lattice method for calculating lift distributions on oscillating surfaces in subsonic flows. AIAA J. 2012, 7, 279–285. [Google Scholar] [CrossRef]
  17. Schmidt, H.; Neumann, J.; Mai, H. Analysis of forced response in a wind tunnel based on Doublet-Lattice method. CEAS Aeronaut. J. 2011, 2, 271–277. [Google Scholar] [CrossRef]
  18. Martínez, M.S.; Dimitriadis, G. Subsonic source and doublet panel methods. J. Fluids Struct. 2022, 113, 103624. [Google Scholar] [CrossRef]
  19. Gao, W.; Liu, Y.; Li, Q.; Lu, B. Aerodynamic Modeling and Simulation of Multi-Lifting Surfaces Based on the Unsteady Vortex Lattice Method. Aerospace 2023, 10, 203. [Google Scholar] [CrossRef]
  20. Boutet, J.; Dimitriadis, G. Unsteady Lifting Line Theory Using the Wagner Function for the Aerodynamic and Aeroelastic Modeling of 3D Wings. Aerospace 2018, 5, 92. [Google Scholar] [CrossRef]
  21. Cole, J.A.; Maughmer, M.D.; Bramesfeld, G.; Melville, M.; Kinzel, M. Unsteady Lift Prediction with a Higher-Order Potential Flow Method. Aerospace 2020, 7, 60. [Google Scholar] [CrossRef]
  22. Luckring, J.M.; Rizzi, A. Prediction of concentrated vortex aerodynamics: Current CFD capability survey. Prog. Aerosp. Sci. 2024, 147, 100998. [Google Scholar] [CrossRef]
  23. Hilger, J.; Ritter, M.R. Nonlinear Aeroelastic Simulations and Stability Analysis of the Pazy Wing Aeroelastic Benchmark. Aerospace 2021, 8, 308. [Google Scholar] [CrossRef]
  24. Li, X.; Wan, Z.; Wang, X.; Yang, C. Aeroelastic Optimization Design of the Global Stiffness for a Joined Wing Aircraft. Appl. Sci. 2021, 11, 11800. [Google Scholar] [CrossRef]
  25. Chen, P.C.; Gao, X.W.; Tang, L. Overset Field-Panel Method for Unsteady Transonic Aerodynamic Influence Coefficient Matrix Generation. AIAA J. 2012, 42, 1775–1787. [Google Scholar] [CrossRef]
  26. Silva, R.G.A.; Mello, O.A.F.; Azevedo, J.L.F.; Chen, P.C.; Liu, D.D. Investigation on Transonic Correction Methods for Unsteady Aerodynamics and Aeroelastic Analyses. J. Aircr. 2012, 45, 1890–1903. [Google Scholar] [CrossRef]
  27. Friedewald, D.; Thormann, R.; Kaiser, C.; Nitzsche, J. Quasi-steady doublet-lattice correction for aerodynamic gust response prediction in attached and separated transonic flow. CEAS Aeronaut. J. 2018, 9, 53–66. [Google Scholar] [CrossRef]
  28. Thormann, R.; Dimitrov, D. Correction of aerodynamic influence matrices for transonic flow. CEAS Aeronaut. J. 2014, 5, 435–446. [Google Scholar] [CrossRef]
  29. Rahmanian, M.; Farsadi, T.; Kurtaran, H. Nonlinear flutter of tapered and skewed cantilevered plates with curvilinear fiber paths. J. Sound. Vib. 2021, 500, 116021. [Google Scholar] [CrossRef]
  30. Asadi, D.; Farsadi, T.; Kayran, A. Flutter Optimization of a Wing–Engine System with Passive and Active Control Approaches. AIAA J. 2020, 59, 1422–1440. [Google Scholar] [CrossRef]
  31. Akhavan, H.; Ribeiro, P. Reduced-Order Models for Nonlinear Flutter of Composite Laminates with Curvilinear Fibers. AIAA J. 2019, 57, 3026–3039. [Google Scholar] [CrossRef]
  32. Kefal, A.; Tessler, A.; Oterkus, E. An enhanced inverse finite element method for displacement and stress monitoring of multilayered composite and sandwich structures. Compos. Struct. 2017, 179, 514–540. [Google Scholar] [CrossRef]
  33. Najmi, J.; Khan, H.A.; Javaid, S.S.; Hameed, A.; Siddiqui, F. Aeroelastic tailoring for aerospace applications. Heliyon 2024, 10, e24151. [Google Scholar] [CrossRef] [PubMed]
  34. Fraihat, B.I.; Ajaj, R.M. Aeroelastic tailoring of stiffened cantilever plate using composites and structural layouts: A parametric study. Aeronaut. J. 2024, 128, 2073–2091. [Google Scholar] [CrossRef]
  35. Leng, J.X.; Wang, Z.G.; Huang, W.; Shen, Y.; An, K. Multidisciplinary Design Optimization Processes for Efficiency Improvement of Aircraft: State-of-the-Art Review. Int. J. Aeronaut. Sp. Sci. 2024. [Google Scholar] [CrossRef]
  36. Sedelnikov, A.; Kurkin, E.; Quijada-Pioquinto, J.G.; Lukyanov, O.; Nazarov, D.; Chertykovtseva, V.; Kurkina, E.; Hoang, V.H. Algorithm for Propeller Optimization Based on Differential Evolution. Comput. 2024, 12, 52. [Google Scholar] [CrossRef]
  37. Hosseini, S.; Vaziry-Zanjany, M.A.; Ovesy, H.R. A Framework for Aircraft Conceptual Design and Multidisciplinary Optimization. Aerospace 2024, 11, 273. [Google Scholar] [CrossRef]
  38. Daoud, M.S.; Shehab, M.; Al-Mimi, H.M.; Abualigah, L.; Zitar, R.A.; Shambour, M.K.Y. Gradient-Based Optimizer (GBO): A Review, Theory, Variants, and Applications. Arch. Comput. Methods Eng. 2023, 30, 2431–2449. [Google Scholar] [CrossRef] [PubMed]
  39. Ahmadianfar, I.; Bozorg-Haddad, O.; Chu, X. Gradient-based optimizer: A new metaheuristic optimization algorithm. Inf. Sci. 2020, 540, 131–159. [Google Scholar] [CrossRef]
  40. Abd Elaziz, M.; Dahou, A.; Abualigah, L.; Yu, L.; Alshinwan, M.; Khasawneh, A.M.; Lu, S. Advanced metaheuristic optimization techniques in applications of deep neural networks: A review. Neural Comput. Appl. 2021, 33, 14079–14099. [Google Scholar] [CrossRef]
  41. Gürgen, S.; Kahraman, H.T.; Aras, S.; Altın, İ. A comprehensive performance analysis of meta-heuristic optimization techniques for effective organic rankine cycle design. Appl. Therm. Eng. 2022, 213, 118687. [Google Scholar] [CrossRef]
  42. Carvalho, J.P.G.; Vargas, D.E.C.; Jacob, B.P.; Lima, B.S.L.P.; Hallak, P.H.; Lemonge, A.C.C. Multi-objective structural optimization for the automatic member grouping of truss structures using evolutionary algorithms. Comput. Struct. 2024, 292, 107230. [Google Scholar] [CrossRef]
  43. Wang, J.; Zheng, Y.; Huang, P.; Peng, H.; Wu, Z. A stable-state multi-objective evolutionary algorithm based on decomposition. Expert. Syst. Appl. 2024, 239, 122452. [Google Scholar] [CrossRef]
  44. Salgotra, R.; Gandomi, A.H. A novel multi-hybrid differential evolution algorithm for optimization of frame structures. Sci. Rep. 2024, 14, 4877. [Google Scholar] [CrossRef]
  45. Zapotecas-Martínez, S.; Armas, R.; García-Nájera, A. A multi-objective evolutionary approach for the electric vehicle charging stations problem. Expert. Syst. Appl. 2024, 240, 122514. [Google Scholar] [CrossRef]
  46. Xu, X.; Xie, C.; Luo, Z.; Zhang, C.; Zhang, T. A multi-objective evolutionary algorithm based on dimension exploration and discrepancy evolution for UAV path planning problem. Inf. Sci. 2024, 657, 119977. [Google Scholar] [CrossRef]
  47. Bai, Z.; Zhou, H.; Shi, J.; Xing, L.; Wang, J. A hybrid multi-objective evolutionary algorithm with high solving efficiency for UAV defense programming. Swarm Evol. Comput. 2024, 87, 101572. [Google Scholar] [CrossRef]
  48. Cai, T.; Wang, H. A general convergence analysis method for evolutionary multi-objective optimization algorithm. Inf. Sci. 2024, 663, 120267. [Google Scholar] [CrossRef]
  49. Nadimi-Shahraki, M.H.; Taghian, S.; Mirjalili, S. An improved grey wolf optimizer for solving engineering problems. Expert. Syst. Appl. 2021, 166, 113917. [Google Scholar] [CrossRef]
  50. Chou, J.S.; Truong, D.N. A novel metaheuristic optimizer inspired by behavior of jellyfish in ocean. Appl. Math. Comput. 2021, 389, 125535. [Google Scholar] [CrossRef]
  51. Got, A.; Zouache, D.; Moussaoui, A. MOMRFO: Multi-objective Manta ray foraging optimizer for handling engineering design problems. Knowl.-Based Syst. 2022, 237, 107880. [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. Zhong, C.; Li, G.; Meng, Z. Beluga whale optimization: A novel nature-inspired metaheuristic algorithm. Knowl.-Based Syst. 2022, 251, 109215. [Google Scholar] [CrossRef]
  54. Makhadmeh, S.N.; Al-Betar, M.A.; Doush, I.A.; Awadallah, M.A.; Kassaymeh, S.; Mirjalili, S.; Zitar, R.A. Recent Advances in Grey Wolf Optimizer, its Versions and Applications: Review. IEEE Access 2024, 12, 22991–23028. [Google Scholar] [CrossRef]
  55. Luiz Junho Pereira, J.; Antônio Oliver, G.; Brendon Francisco, M.; Simões Cunha, S.; Ferreira Gomes, G. Multi-objective lichtenberg algorithm: A hybrid physics-based meta-heuristic for solving engineering problems. Expert. Syst. Appl. 2022, 187, 115939. [Google Scholar] [CrossRef]
  56. Nematollahi, A.F.; Rahiminejad, A.; Vahidi, B. MOGROM: Multiobjective Golden Ratio Optimization Algorithm. In Multi-Objective Combinatorial Optimization Problems and Solution Methods; Academic Press: Cambridge, MA, USA, 2022; pp. 91–119. [Google Scholar]
  57. Ghasemi, M.; Golalipour, K.; Zare, M.; Mirjalili, S.; Trojovský, P.; Abualigah, L.; Hemmati, R. Flood algorithm (FLA): An efficient inspired meta-heuristic for engineering optimization. J. Supercomput. 2024, 80, 22913–23017. [Google Scholar] [CrossRef]
  58. Verij kazemi, M.; Fazeli Veysari, E. A new optimization algorithm inspired by the quest for the evolution of human society: Human felicity algorithm. Expert. Syst. Appl. 2022, 193, 116468. [Google Scholar] [CrossRef]
  59. Xu, Z.; Zhang, K. Multiobjective multifactorial immune algorithm for multiobjective multitask optimization problems. Appl. Soft Comput. 2021, 107, 107399. [Google Scholar] [CrossRef]
  60. Fomin, F.V.; Fraigniaud, P.; Golovach, P.A. Present-biased optimization. Math. Soc. Sci. 2022, 119, 56–67. [Google Scholar] [CrossRef]
  61. Oladejo, S.O.; Ekwe, S.O.; Mirjalili, S. The Hiking Optimization Algorithm: A novel human-based metaheuristic approach. Knowl.-Based Syst. 2024, 296, 111880. [Google Scholar] [CrossRef]
  62. Sowmya, R.; Premkumar, M.; Jangir, P. Newton-Raphson-based optimizer: A new population-based metaheuristic algorithm for continuous optimization problems. Eng. Appl. Artif. Intell. 2024, 128, 107532. [Google Scholar] [CrossRef]
  63. Stanford, B.; Ifju, P. Multi-Objective Topology Optimization of Wing Skeletons for Aeroelastic Membrane Structures. Int. J. Micro Air Veh. 2009, 1, 51–69. [Google Scholar] [CrossRef]
  64. Guo, S.; Li, D.; Liu, Y. Multi-objective optimization of a composite wing subject to strength and aeroelastic constraints. Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 2011, 226, 1095–1106. [Google Scholar] [CrossRef]
  65. Phiboon, T.; Khankwa, K.; Petcharat, N.; Phoksombat, N.; Kanazaki, M.; Kishi, Y.; Bureerat, S.; Ariyarit, A. Experiment and computation multi-fidelity multi-objective airfoil design optimization of fixed-wing UAV. J. Mech. Sci. Technol. 2021, 35, 4065–4072. [Google Scholar] [CrossRef]
  66. Wansasueb, K.; Pholdee, N.; Panagant, N.; Bureerat, S. Multiobjective meta-heuristic with iterative parameter distribution estimation for aeroelastic design of an aircraft wing. Eng. Comput. 2022, 38, 695–713. [Google Scholar] [CrossRef]
  67. Wansasueb, K.; Panagant, N.; Bureerat, S.; Sabangban, N.; Pholdee, N. Comparative study of recent metaheuristics for solving a multiobjective transonic aeroelastic optimization of a composite wing. Acta Mech. 2024, 235, 391–407. [Google Scholar] [CrossRef]
  68. Rajpal, D.; Mitrotta, F.M.A.; Socci, C.A.; Sodja, J.; Kassapoglou, C.; De Breuker, R. Design and testing of aeroelastically tailored composite wing under fatigue and gust loading including effect of fatigue on aeroelastic performance. Compos. Struct. 2021, 275, 114373. [Google Scholar] [CrossRef]
  69. Farsadi, T.; Ahmadi, M.; Sahin, M.; Haddad Khodaparast, H.; Kayran, A.; Friswell, M.I. High Aspect Ratio Composite Wings: Geometrically Nonlinear Aeroelasticity, Multi-Disciplinary Design Optimization, Manufacturing, and Experimental Testing. Aerospace 2024, 11, 193. [Google Scholar] [CrossRef]
  70. Gemma, S.; Mastroddi, F. Genetic and gradient-based algorithms for the multi-objective optimization of aircraft design with aeroelastic constraints. In Proceedings of the International Forum on Aeroelasticity and Structural Dynamics (IFASD 2015), Saint Petersburg, Russia, 28 June–2 July 2015; Central Aerohydrodynamic Institute: Zhukovskiy, Russia, 2017; pp. 692–712. [Google Scholar]
  71. Caixeta, P.R.; Marques, F.D. Multiobjective optimization of an aircraft wing design with respect to structural and aeroelastic characteristics using neural network metamodel. J. Braz. Soc. Mech. Sci. Eng. 2018, 40, 17. [Google Scholar] [CrossRef]
  72. Elyasi, M.; Roudbari, A.; Hajipourzadeh, P. Multi-objective robust design optimization (MORDO) of an aeroelastic high-aspect-ratio wing. J. Braz. Soc. Mech. Sci. Eng. 2020, 42, 560. [Google Scholar] [CrossRef]
  73. Gasparetto, V.E.L.; Reid, J.; Parsons, W.P.; ElSayed, M.S.A.; Saad, M.; Shieldand, S.; Brown, G.L.; Hilliard, L.M. Multi-Objective Design Optimization of Multiple Tuned Mass Dampers for Attenuation of Dynamic Aeroelastic Response of Aerospace Structures. Aerospace 2023, 10, 235. [Google Scholar] [CrossRef]
  74. Zappino, E. Multi-Objective Aeroelastic Analysis and Optimization Using Surrogate Models. Ph.D. Thesis, Politecnico di Torino, Turin, Italy, 2023. [Google Scholar]
  75. Nogueira, F.S.; Pereira, J.L.J.; Simões Cunha, S. Multi-objective sensor placement optimization and damage identification for an aircraft wing using Lichtenberg algorithm. Eng. Comput. 2024, 41, 438–467. [Google Scholar] [CrossRef]
  76. Colella, M. Integrated Aeroservoelastic Modeling for Preliminary Aircraft Design with Multi-Objective Optimization. Available online: https://iris.uniroma1.it/handle/11573/1711034 (accessed on 25 December 2024).
  77. Sleesongsom, S.; Bureerat, S. New conceptual design of aeroelastic wing structures by multi-objective optimization. Eng. Optim. 2013, 45, 107–122. [Google Scholar] [CrossRef]
  78. Ali, N.; Behdinan, K. Optimal Geometrical Design of Aircraft Using Genetic Algorithms. Trans. Can. Soc. Mech. Eng. 2018, 26, 373–388. [Google Scholar] [CrossRef]
  79. Singh, V.; Sharma, S.K.; Vaibhav, S. Transport Aircraft Conceptual Design Optimization Using Real Coded Genetic Algorithm. Int. J. Aerosp. Eng. 2016, 2016, 2813541. [Google Scholar] [CrossRef]
  80. Wang, D.; Tan, D.; Liu, L. Particle swarm optimization algorithm: An overview. Soft Comput. 2018, 22, 387–408. [Google Scholar] [CrossRef]
  81. Ahmadi, M.; Farsadi, T.; Haddad Khodaparast, H. Enhancing gust load alleviation performance in an optimized composite wing using passive wingtip devices: Folding and Twist approaches. Aerosp. Sci. Technol. 2024, 147, 109023. [Google Scholar] [CrossRef]
  82. Zhang, Y.; Zhou, J.; Shi, L.; Cui, B. Design optimization for the entire aircraft structure of civil aircraft with blended-wing-body layout. Struct. Multidiscip. Optim. 2024, 67, 106. [Google Scholar] [CrossRef]
  83. Bharti, S.; Frecker, M.; Lesieutre, G. Optimal structural design of a morphing aircraft wing using parallel non-dominated sorting genetic algorithm II (NSGA II). In Proceedings of the Smart Structures and Materials 2006: Modeling, Signal Processing, and Control, San Diego, CA, USA, 26 February–2 March 2006; Volume 6166, pp. 12–23. [Google Scholar] [CrossRef]
  84. Zandavi, S.M.; Pourtakdoust, S.H. Multidisciplinary design of a guided flying vehicle using simplex nondominated sorting genetic algorithm II. Struct. Multidiscip. Optim. 2018, 57, 705–720. [Google Scholar] [CrossRef]
  85. Gunpinar, E.; Khan, S. A multi-criteria based selection method using non-dominated sorting for genetic algorithm based design. Optim. Eng. 2020, 21, 1319–1357. [Google Scholar] [CrossRef]
  86. Shrivastava, S.; Mohite, P.M.; Yadav, T.; Malagaudanavar, A. Multi-objective multi-laminate design and optimization of a Carbon Fibre Composite wing torsion box using evolutionary algorithm. Compos. Struct. 2018, 185, 132–147. [Google Scholar] [CrossRef]
  87. Ahmadi, M.; Farsadi, T. Multidisciplinary optimization of high aspect ratio composite wings with geometrical nonlinearity and aeroelastic tailoring. Aerosp. Sci. Technol. 2024, 145, 108849. [Google Scholar] [CrossRef]
  88. Huang, Y.; Wang, Z.; Chen, P. Buckling optimization of variable stiffness composite wing boxes with manufacturing defects. Struct. Multidiscip. Optim. 2024, 67, 109. [Google Scholar] [CrossRef]
  89. Arsenyeva, A.; Duddeck, F.; Thompson, H.M. An iso-contour method for automated fiber placement optimization of composite structures. Compos. Struct. 2024, 327, 117628. [Google Scholar] [CrossRef]
  90. Jamil, T.; Iqbal, A.; Allauddin, U.; Ahmad, E.; Hashmi, S.A.; Saleem, S.; Ikhlaq, M. Fluid Coupled Structural Analysis and Optimization of Expanded Polystyrene-Fiber-Reinforced Composite Wing of an Unmanned Aerial Vehicle. Mech. Compos. Mater. 2024, 60, 211–226. [Google Scholar] [CrossRef]
  91. Mitrotta, F.M.A.; Rajpal, D.; Sodja, J.; De Breuker, R. Multi-fidelity design of an aeroelastically tailored composite wing for dynamic wind-tunnel testing. In Proceedings of the AIAA Scitech 2020 Forum, Orlando, FL, USA, 6–10 January 2020. [Google Scholar] [CrossRef]
  92. Farsadi, T.; Bozkurt, M.O.; Coker, D.; Kayran, A. Improvement of structural characteristics of composite thin-walled beams using variable stiffness concept via curvilinear fiber placement. Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 2021, 235, 2017–2032. [Google Scholar] [CrossRef]
  93. Nikolaos, Z.; Spyridon, K.; Vassilis, P.D.K. Multidisciplinary Aeroelastic Optimization of High Aspect-Ratio Composite Aircraft Wings via Tow-Steering. PREPRINT (Version 1) available at Research Square. 2024. Available online: https://www.researchsquare.com/article/rs-4764693/v1 (accessed on 25 November 2024). [CrossRef]
  94. Brooks, T.R.; Kenway, G.K.W.; Martins, J.R.R.A. Benchmark aerostructural models for the study of transonic aircraft wings. AIAA J. 2018, 56, 2840–2855. [Google Scholar] [CrossRef]
  95. DPW6 Geometries|NASA Common Research Model. Available online: https://commonresearchmodel.larc.nasa.gov/geometry/dpw6-geometries/ (accessed on 25 December 2024).
  96. Computational Results|NASA Common Research Model. Available online: https://commonresearchmodel.larc.nasa.gov/computational-results/ (accessed on 25 December 2024).
  97. Valencia, E.; Alulema, V.; Hidalgo, V.; Rodriguez, D. A CAD-free methodology for volume and mass properties computation of 3-D lifting surfaces and wing-box structures. Aerosp. Sci. Technol. 2021, 108, 106378. [Google Scholar] [CrossRef]
  98. Kefal, A.; Oterkus, E.; Tessler, A.; Spangler, J.L. A quadrilateral inverse-shell element with drilling degrees of freedom for shape sensing and structural health monitoring. Eng. Sci. Technol. Int. J. 2016, 19, 1299–1313. [Google Scholar] [CrossRef]
  99. Nguyen-Van, H.; Mai-Duy, N.; Karunasena, W.; Tran-Cong, T. Buckling and vibration analysis of laminated composite plate/shell structures via a smoothed quadrilateral flat shell element with in-plane rotations. Comput. Struct. 2011, 89, 612–625. [Google Scholar] [CrossRef]
  100. Figueiras, J.d.A. Ultimate Load Analysis of Anisotropic and Reinforced Concrete Plates and Shells. Ph.D. Thesis, Swansea University, Sketty, UK, 1983. [Google Scholar]
  101. Rolfes, R.; Rohwer, K. Improved transverse shear stresses in composite finite elements based on first order shear deformation theory. Int. J. Numer. Methods Eng. 1997, 40, 51–60. [Google Scholar] [CrossRef]
  102. Ünlüsoy, L. Structural Design and Analysis of the Mission Adaptive Wings of an Unmanned Aerial Vehicle. Master’s Thesis, Middle East Technical University, Çankaya, Türkiye, 2010. Available online: https://open.metu.edu.tr/handle/11511/19160 (accessed on 25 December 2024).
  103. Narayanan, R.M. Microwave Nondestructive Testing of Galvanic Corrosion and Impact Damage in Carbon Fiber Reinforced Polymer Composites. Int. J. Microw. Appl. 2018, 7, 1–15. [Google Scholar] [CrossRef]
  104. Guerreiro, A.P.; Fonseca, C.M.; Paquete, L. The Hypervolume Indicator: Computational Problems and Algorithms. ACM Comput. Surv. 2021. Epub ahead of print. [Google Scholar] [CrossRef]
  105. Zitzler, E.; Deb, K.; Thiele, L. Comparison of multiobjective evolutionary algorithms: Empirical results. Evol. Comput. 2000, 8, 173–195. [Google Scholar] [CrossRef] [PubMed]
Figure 1. FEA model of the CRM wing.
Figure 1. FEA model of the CRM wing.
Aerospace 12 00101 g001
Figure 2. Assignment of design variables for tow-steered orientation distribution.
Figure 2. Assignment of design variables for tow-steered orientation distribution.
Aerospace 12 00101 g002
Figure 3. Optimization flow chart.
Figure 3. Optimization flow chart.
Aerospace 12 00101 g003
Figure 4. Convergence plots of all MHs for a tow-steered aeroelastic problem.
Figure 4. Convergence plots of all MHs for a tow-steered aeroelastic problem.
Aerospace 12 00101 g004
Figure 5. The best Pareto front of all MHs for the tow-steered aeroelastic problem.
Figure 5. The best Pareto front of all MHs for the tow-steered aeroelastic problem.
Aerospace 12 00101 g005
Figure 6. Box plots of all MHs in the tow-steered aeroelastic problem.
Figure 6. Box plots of all MHs in the tow-steered aeroelastic problem.
Aerospace 12 00101 g006
Figure 7. The best Pareto front for tow-steered aeroelastic results from MOMVO.
Figure 7. The best Pareto front for tow-steered aeroelastic results from MOMVO.
Aerospace 12 00101 g007
Figure 8. Three samples of tow-steered composite wings from the best Pareto front.
Figure 8. Three samples of tow-steered composite wings from the best Pareto front.
Aerospace 12 00101 g008
Figure 9. Comparison of the best Pareto front of composite wing aeroelastic results [67] and tow-steered aeroelastic results.
Figure 9. Comparison of the best Pareto front of composite wing aeroelastic results [67] and tow-steered aeroelastic results.
Aerospace 12 00101 g009
Table 1. Material properties.
Table 1. Material properties.
MaterialProperty ValueUnit
AluminumYoung’s modulus ( E )7.31 × 1010Pa.
Poisson’s ratio ( ϑ )0.33-
Density ( ρ )2780kg/m3
Yield stress ( σ y )4.10 × 107Pa.
Shear modulus ( G )2.70 × 1010Pa.
Carbon fiberYoung’s modulus ( E 11 )1.35 × 1011Pa.
Young’s modulus ( E 22 )1.00 × 1010Pa.
Shear modulus ( G 12 )5.00 × 109Pa.
Shear modulus ( G 13 )5.00 × 109Pa.
Shear modulus ( G 23 )5.00 × 109Pa.
Poisson’s ratio ( ϑ )0.3-
Density ( ρ )1600kg/m3
Tensile strength ( S t y , 11 )1.50 × 109Pa.
Tensile strength ( S t y , 22 )5.00 × 107Pa.
Tensile shear strength ( S t y , 12 )7.00 × 107Pa.
Compressive strength ( S c y , 11 )1.20 × 109Pa.
Compressive strength ( S c y , 22 )2.50 × 108Pa.
Compressive shear strength ( S c y , 12 )7.00 × 107Pa.
Stress of bonding ( S b )1.00 × 107Pa.
Table 2. Parameter settings of MHs.
Table 2. Parameter settings of MHs.
AlgorithmParameter NameAbbreviationValue
Multobjective Mayfly
Optimization Algorithm
(MOMA)
Inertia weight g 0.8
Inertia weight damping ratio g d a m p 1
Personal learning coefficient a 1 1
Global learning coefficient a 2 & a 3 1.5
Distance sight coefficient β 2
Mutation coefficient d 0.77
Mutation coefficient damping ratio d d a m p 0.99
Random flight f 0.77
Random flight damping ratio f d a m p 0.99
Mutation rate μ 0.02
Multiobjective Salp Swarm
Algorithm (MSSA)
---
Multiobjective Grasshopper
Optimization Algorithm
(MOGOA)
Exploration and exploitation proportional C [0.00004, 1]
Multiobjective Multi-Verse Optimization
(MOMVO)
Wormhole existence probability W E P [0.2, 1]
Multiobjective Lichtenberg
Algorithm (MOLA)
Reference Lichtenberg point r e f 0.4
Number of particles N P 100,000
Creation radius R c 150
Stick probability S c 1
Number of grids in each dimension N G r i d 30
Multiobjective Jellyfish
Search (MOJS)
Hypercube limitation grid point N G r i d 20
Multiobjective Crystal
Structure Algorithm
(MOCryStAl)
Grid inflation parameter α 0.1
Number of grids per each dimension N G r i d 30
Leader selection pressure parameter β 4
Extra repository member selection pressure γ 2
Multiobjective Manta Ray
Foraging Optimizer
(MOMRFO)
Grid inflation parameter α 0.1
Number of grids per each dimension N G r i d 30
Leader selection pressure parameter β 4
Extra archive member selection pressure γ 2
Table 3. Statistical results of a tow-steered aeroelastic problem considering a hypervolume indicator.
Table 3. Statistical results of a tow-steered aeroelastic problem considering a hypervolume indicator.
AlgorithmMOCryStAlMOGOAMOJSMOLAMOMAMOMRFOMOMVOMSSA
average0.427290.474620.524410.437540.561440.406380.654950.44987
standard deviation0.064880.230920.030940.066180.058340.154150.077780.18921
maximum0.507950.679490.568710.524110.657020.529660.767480.62961
minimum0.2941900.463140.292340.47009500.461450
rank74362815
Bold text presents the average, best, worst, standard deviation, and rank of hypervolume indicator.
Table 4. Aeroelastic results of three optimal solutions sampled from the Pareto front.
Table 4. Aeroelastic results of three optimal solutions sampled from the Pareto front.
Optimum Solution NumberNo. 1No. 2No. 3
1st objective
total mass [kg.]
32,639.1443944,057.8484069,370.10817
2nd objective
critical speed [m/s]
511.90569832.25059876.06638
total mass [kg.]32,639.1443944,057.8484069,370.10817
divergent speed [m/s]65,53565,53565,535
flutter speed [m/s]511.90569832.25059876.06638
maximum
deflection [m.]
0.029920.026390.01727
maximum
Von-Misses stress [Pa.]
200,433,275.012142,468,641.42058,506,362.433
maximum
Tsai Hill [-]
0.000840.000630.00028
lift effectiveness [-]1.031901.022891.01704
minimum
buckling factor [-]
1.158961.023241.73896
Table 5. Design variables of three optimal solutions sampled from Pareto front.
Table 5. Design variables of three optimal solutions sampled from Pareto front.
Optimum Solution NumberSolution No. 1Solution No. 2Solution No. 3
1st objective
total mass [kg.]
32,639.1443944,057.8484069,370.10817
2nd objective
critical speed [m/s]
511.9056913832.2505855876.0663793
thickness at root chordleading spar [mm.]87.57
tailing spar [mm.]8.51729
ribs [mm.]11.54.5
thickness at tip chordleading spar [mm.]2.30.71.4
tailing spar [mm.]6.613.67.6
ribs [mm.]11.44.3
thickness at upper skin root chordlayer 1, inner layer [mm.]112
layer 2 [mm.]998
layer 3, middle layer [mm.]0.20.235
layer 4 [mm.]0.4111
layer 5, outer layer [mm.]201123
thickness at upper skin tip chordlayer 1, inner layer [mm.]0.20.20.3
layer 2 [mm.]6.17.25.6
layer 3, middle layer [mm.]0.10.16.4
layer 4 [mm.]0.20.55.8
layer 5, outer layer [mm.]12.810.115.4
orientation at upper skin root chordlayer 1, inner layer [mm.]100105105
layer 2 [mm.]135130140
layer 3, middle layer [mm.]125155150
layer 4 [mm.]165165120
layer 5, outer layer [mm.]304535
orientation at upper skin tip chordlayer 1, inner layer [mm.]145145120
layer 2 [mm.]120130155
layer 3, middle layer [mm.]145160165
layer 4 [mm.]453545
layer 5, outer layer [mm.]2000
thickness at lower skin root chordlayer 1, inner layer [mm.]0.218
layer 2 [mm.]183029
layer 3, middle layer [mm.]121616
layer 4 [mm.]193140
layer 5, outer layer [mm.]7712
thickness at lower skin tip chordlayer 1, inner layer [mm.]0.217.8
layer 2 [mm.]7.729.627.8
layer 3, middle layer [mm.]8.711.49.1
layer 4 [mm.]18.73140
layer 5, outer layer [mm.]2.94.25
orientation at lower skin root chordlayer 1, inner layer [mm.]13012585
layer 2 [mm.]115110145
layer 3, middle layer [mm.]101510
layer 4 [mm.]150150150
layer 5, outer layer [mm.]165160115
orientation at lower skin tip chordlayer 1, inner layer [mm.]145145145
layer 2 [mm.]755045
layer 3, middle layer [mm.]1500
layer 4 [mm.]115115120
layer 5, outer layer [mm.]30305
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

Phuekpan, K.; Khammee, R.; Panagant, N.; Bureerat, S.; Pholdee, N.; Wansasueb, K. A Comparison of Modern Metaheuristics for Multi-Objective Optimization of Transonic Aeroelasticity in a Tow-Steered Composite Wing. Aerospace 2025, 12, 101. https://doi.org/10.3390/aerospace12020101

AMA Style

Phuekpan K, Khammee R, Panagant N, Bureerat S, Pholdee N, Wansasueb K. A Comparison of Modern Metaheuristics for Multi-Objective Optimization of Transonic Aeroelasticity in a Tow-Steered Composite Wing. Aerospace. 2025; 12(2):101. https://doi.org/10.3390/aerospace12020101

Chicago/Turabian Style

Phuekpan, Kantinan, Rachata Khammee, Natee Panagant, Sujin Bureerat, Nantiwat Pholdee, and Kittinan Wansasueb. 2025. "A Comparison of Modern Metaheuristics for Multi-Objective Optimization of Transonic Aeroelasticity in a Tow-Steered Composite Wing" Aerospace 12, no. 2: 101. https://doi.org/10.3390/aerospace12020101

APA Style

Phuekpan, K., Khammee, R., Panagant, N., Bureerat, S., Pholdee, N., & Wansasueb, K. (2025). A Comparison of Modern Metaheuristics for Multi-Objective Optimization of Transonic Aeroelasticity in a Tow-Steered Composite Wing. Aerospace, 12(2), 101. https://doi.org/10.3390/aerospace12020101

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop