1. Introduction
Variable camber wings (VCWs) can provide optimal aerodynamic performance and enhance the flight envelope by smoothly and continuously deforming their leading and trailing edges (LATEs), depending on flight conditions and mission objectives. VCWs have emerged as a focal point in morphing aircraft design owing to their low energy usage and superior cruising economy [
1]. In recent years, some companies such as Boeing, NASA, and Airbus have sequentially advanced research on Variable Camber Flexible Wings (VCFWs) [
2], demonstrating the viability of continuous VCWs through the application of smart materials and flexible structures.
Numerous academics are presently engaged in comprehensive investigations examining the aerodynamic performance and structural deformation of VCW. Kaul et al. investigated the aerodynamic effects of trailing-edge camber in the Variable Camber Continuous Trailing Edge Flap (VCCTEF) project [
3,
4], with the results demonstrating that appropriate trailing-edge deflection improves the lift-to-drag ratio and the numerical lift increments are in good agreement with the theoretical predictions [
5,
6]. However, excessive downward deflection angles increase the drag. For this purpose, drag reduction studies in various flight conditions are conducted in order to examine the performance of different flap configurations with selected camber deflection profiles [
7]. To obtain the optimal profiles of VCWs under various flight conditions, the software environment Multidisciplinary Integrated Conceptual Aircraft Design and Optimization (MICADO) was developed. The software also features a sophisticated mission analysis to assess the impact on fuel planning [
8]. Aimed at the structural deformation of VCWs, Keidel et al. [
9,
10] proposed an innovative structural deformation technique to tackle the difficulties encountered in the flight control of flying-wing aircraft. The deformation of the trailing edge was proven through experiments by improving internal flexible components and electromechanical actuators.
Aerodynamic optimization design of variable camber airfoils (VCAs) has yielded substantial scientific outcomes. Typically, three optimization search algorithms are incorporated into aerodynamic optimization design methodologies: (1) gradient-based optimization transformation techniques, such as the Quasi-Newton method [
11]; (2) non-gradient optimization algorithms, such as the genetic algorithm (GA) [
12,
13,
14,
15]; and (3) surrogate model optimization methods [
16,
17,
18]. The gradient technique identifies the search direction of the objective function based on the gradient information of the design variables and iteratively calculates until the objective function converges to a minimum or maximum value. However, the derivative value is calculated repeatedly, necessitating the invocation of the objective function analysis program, resulting in computing complexity that correlates with the quantity of design variables. As the quantity of design variables increases, the computational complexity of the optimization design process intensifies swiftly. Based on the principles of survival of the fittest, as articulated in Darwin’s Theory of Evolution, GA can emulate the process of biological genetic evolution and conduct a global search inside the optimization landscape. It is difficult to fall into local optima during the search process. However, GA necessitates a substantial quantity of objective function evaluations. The application of Euler or Navier–Stokes (NS) equations for the aerodynamic optimization of complicated forms with numerous design variables necessitates substantial processing resources, resulting in low computational efficiency.
The surrogate model is an approximate mathematical representation constructed by sampling the design space. It features low computational cost while maintaining a level of accuracy comparable to that of high-fidelity models, such as those based on the NS equations. Compared with CFD evaluations, surrogate modeling greatly improves the efficiency of aerodynamic optimization and facilitates global optimization searches. The most widely used surrogate models currently include the Polynomial Response Surface model, the Artificial Neural Network model, the Radial Basis Function model, and the Kriging surrogate model [
19]. The Polynomial Response Surface model is suitable for linear or weakly nonlinear problems; however, its accuracy is limited in high-dimensional or strongly nonlinear cases. Although the Artificial Neural Network model offers strong fitting capability, it involves complex training. The Radial Basis Function model provides high interpolation accuracy, but it is sensitive to parameters and requires huge amounts of resources. In contrast, the Kriging model combines prediction accuracy with error estimation, making it the choice for this study due to fewer samples and high-precision modeling. The integration of the Kriging surrogate model with optimization methods has been extensively researched in the field of aircraft design, particularly concerning wings [
20]. Rajagopal et al. [
21] examined the multi-objective design optimization of low-speed, long-endurance UAV wings. They utilized the Kriging surrogate model to substitute high-fidelity analysis tools, hence decreasing computational time, and addressed the optimization problem employing the Non-dominated Sorting Genetic Algorithm II (NSGA-II). The findings disclosed several advantageous Pareto-optimal design methods that can inform the initial design of UAV wings. Qiu et al. [
22] proposed a novel approach for optimizing supersonic airfoils by integrating orthogonal decomposition with data dimensionality reduction to develop a Kriging surrogate model. The results showed that this method reduced the number of design variables by 50% and enhanced optimization efficiency by 200%. Zhao et al. [
23] tackled the multi-objective optimization problem in UAV flying wing control surface design by proposing a multi-objective control allocation method based on the Kriging surrogate model. This approach effectively solved the control allocation issue for UAVs’ continuously deformable trailing edges. Zhao et al. [
24] subsequently used the Kriging surrogate model with NSGA-II for aerodynamic optimization. The optimization results of the RAE2822 airfoil at low Reynolds numbers in transonic and subsonic regimes were compared, and the mechanisms underlying the airfoil’s lift generation were analyzed. Ju et al. [
25] proposed an optimization technique that integrates the particle swarm optimization algorithm with a Kriging surrogate model for the design of high-lift devices. The method was employed to optimize the parameters of flap configuration position, resulting in identification of the ideal flap deflection position with merely a 1% reduction in the lift-to-drag ratio. To enhance the computational efficiency of aerodynamic shape optimization, Raul et al. [
26] proposed a least-squares programming technique utilizing the Kriging surrogate model. This method assisted in alleviating the dynamic stall characteristics of the airfoil. However, there are comparatively few research advancements in the optimal design of VCWs utilizing the Kriging surrogate model. Weaver-Rosen et al. [
27] presented a parameter optimization method for the continuous variable camber design of lightweight aircraft wings. They applied the Kriging surrogate model to the output of a genetic algorithm to obtain the optimal solution. The results indicated that the parameter optimization strategy was applicable under diverse operating situations. Wang et al. [
28] presented a deformation method that combines piezoelectric actuation with flexible dynamic shape control, based on the Kriging surrogate model and using only 150 sampling points. This approach provided an efficient solution for the swift optimization of flexible trailing-edge VCWs. Du et al. [
29] investigated a variable camber airfoil design by integrating the X-foil application with a Kriging surrogate model to elucidate the correlation between driving variables and the aerodynamic characteristics of the airfoil, as summarized in
Table 1.
Although many researchers have studied the aerodynamic characteristics of VCWs, most of these studies have been based on the trailing edge. There are few studies on the influence of leading-edge camber on the aerodynamic characteristics of wings. The Kriging surrogate model is primarily utilized for single-objective optimization and is infrequently employed in multi-objective optimization design of VCAs with LATEs. Consequently, it is essential to establish a Kriging prediction model for the multi-objective optimization of VCAs utilizing LATEs. This work focuses on the supercritical RAE2822 airfoil as the research object. First, the aerodynamic performance of the airfoil is investigated by varying the camber of the LATE. A Kriging surrogate model is established with LATE deflection angles as input variables and the lift coefficient and drag coefficient as output variables. Utilizing NSGA-II multi-objective optimization, the maximum lift coefficient, and the lift-to-drag ratio as optimization objectives, the optimal airfoil configurations for Mach numbers 0.74 and 0.76 are obtained. The discrepancy between the predicted optimization model and the traditional CFD optimization model is within 2%. The reliability and the efficiency of the surrogate model are demonstrated.
4. Multi-Objective Airfoil Optimization Using Kriging Surrogate Model
From the aerodynamic analysis above, it can be concluded that
Macr of the RAE2822 airfoil is 0.73 (
Figure 11d), Re = 1.7 × 10
7, α = 2°. To improve the aerodynamic characteristics of the airfoil beyond the critical Mach number, two distinct flight conditions at
Ma = 0.74 and 0.76 will be chosen for multi-objective optimization. NSGA-II is a widely utilized multi-objective genetic algorithm that optimizes multiple objective functions concurrently via rapid non-dominated sorting and crowding distance computation. It generates a Pareto optimal solution set with good uniform distribution, making it easier to achieve diversity, uniformity, and robustness within the solution set [
33]. The optimization process is illustrated in
Figure 12.
(1) Define the optimization objective functions F (max CL, max K) and design constraints and establish the design space X for the LATE deflections.
(2) Select initial sample points using the optimal Latin hypercube sampling technique and perform parallel CFD calculations to obtain the performance data.
(3) In the surrogate model construction process, the Expected Improvement (EI) criterion is used to guide the automatic dynamic addition of sample points, and EI < 10−3 serves as the convergence criterion to improve the accuracy of the Kriging surrogate model.
(4) Use the NSGA-II algorithm for optimization, generate a Pareto solution set and select the optimal solutions along the Pareto front as the final optimized airfoil configuration.
4.1. Interpolation Accuracy Verification for Kriging Surrogate Model
The OLHS is performed on the deflection angles of the LATE, where the sampling ranges for both angles are [−5°, 5°]. Based on the EI criterion, 30 sample points are collected under two different flight conditions to meet the accuracy requirements. The aerodynamic Kriging surrogate model is then fitted. The fitting accuracy is presented in
Table 3, as indicated by Equations (10) and (11). The values of
R2 for two Mach numbers are greater than 0.95, and the values of
RMSE are less than 0.07, the results demonstrate that the surrogate model has remarkable reliability in handling complex aerodynamic data.
Figure 13 and
Figure 14 show the surrogate models for aerodynamic coefficients and the corresponding fitting validation plots for
Ma = 0.74 and 0.76.
CL and
CD surrogate models are in good agreement with the aerodynamic analysis results (
Figure 8 and
Figure 11).
The solid line in the fitting validation graphs denotes the locus where the true values coincide with the anticipated values. The closer the sample points are to this solid line, the smaller the prediction deviation, the more accurate the model. The percentage of high-confidence sample points exceeds 95% for Ma = 0.74 and 0.76.
4.2. Multi-Objective Optimization
The optimization processes at two Mach numbers emphasize maximizing the lift-to-drag ratio and the lift coefficient, utilizing a population size of 40 for the NSGA-II algorithm, with 200 evolutionary generations, a crossover probability of 0.9, and a mutation probability of 0.01. If the optimization is performed by sequential loop iterations, the numerical model may be called as many as 8000 times. The optimization process utilizing the Kriging surrogate model only requires 30 iterations of the numerical model, significantly reducing the computational burden.
The optimization model for
Ma = 0.74 and 0.76 is as follows:
where
represents the optimization objective of the lift coefficient and lift-to-drag ratio,
denotes the design space, and
represents the constraints: (1) the optimized airfoil area
S1 is greater than and equal to the basic airfoil area
S0 and (2) the aerodynamic coefficients of the optimized airfoil are greater than and equal to those of the basic airfoil. It can be seen from
Figure 15 and
Figure 16 that the most feasible solution is uniformly distributed across the constraint interval with high density, and both optimization objective values are improved.
Figure 17 illustrates the optimized solutions at
Ma = 0.74 and
Ma = 0.76. The solutions include the Pareto front and the dominated solutions. Optimization solution 1 is closer to the lift-to-drag ratio, optimization solution 2 embodies a trade-off between the lift-to-drag ratio and the lift coefficient, and optimization solution 3 is closer to the lift coefficient. This study eventually selects the compromise represented by optimization solution 2.
Table 4 and
Table 5 demonstrate that the discrepancies between the approximate results of the Kriging surrogate model and the CFD simulation results are less than 2%, indicating the high reliability of the surrogate model.
The deflection angles of the LATE for the optimized airfoil at
Ma = 0.74 and
Ma = 0.76 are
and
, and
and
, respectively. The optimization results are shown in
Table 6.
CL increased by 7.55% and 7.37%, and
K increased by 6.97% and 10.27%, respectively. The results achieve the desired multi-objective aerodynamic performance optimization.
The pressure coefficient contours of the basic and optimized airfoils are shown in
Figure 18. As shown in the figure, the lift of the wing mainly depends on the suction effect generated by the strong negative pressure zone in front of the upper wing, rather than the positive pressure on the lower surface. From
Figure 18b,d, it can be seen that the negative pressure effect becomes stronger as the incoming Mach number increases. Compared with the basic airfoils, two negative pressure peaks appear on the upper surface of optimized airfoils, indicating an increase in lift. Both the basic and optimized airfoils exhibit significant shock wave phenomena on their upper surfaces, indicating a typical supersonic flight state. The optimized airfoil has significantly expanded the negative pressure zone in the leading-edge region of the upper surface and moved forward. It is obvious that the suction center shows a trend of moving forward compared to the basic airfoil. It indicates that the optimized design achieves local flow control through the deflections of the LATE, accelerating the airflow earlier and obtaining stronger suction effects, thereby effectively enhancing the lift performance of the wing.
The pressure coefficient curves of the basic airfoil and the optimized airfoil are shown in
Figure 19. It can be seen that the suction peak of the optimized airfoil in the leading-edge region has significantly increased, indicating that the airflow acceleration at this location is stronger, while the shock wave position moves forward and the intensity is slightly weakened. This indicates that the adverse effects of shock wave interference on the boundary layer are reduced, thereby contributing to a decrease in overall drag levels. However, the weakening of shock waves can also change the boundary layer state on the wing surface, making it easier for the airflow to separate, subsequently resulting in a decrease in lift coefficient.
5. Conclusions
Aiming at the supercritical RAE2822 airfoil, the aerodynamic effects of variable camber for the LATE are investigated. A multi-objective optimization strategy is established based on the Kriging surrogate model and NSGA II, and the following conclusions are achieved:
(1) The upward deflection of the leading edge marginally increases the lift-to-drag ratio, whilst the downward deflection of the leading edge improves the value of the critical angle of attack and the stall characteristics of the airfoil. The trailing-edge deflection has a significant influence on the critical angle of attack, while the downward deflection of the trailing edge increases the lift coefficient. Moderate upward deflection of both LATEs can delay the critical Mach number, while downward deflection of the LATE results in a reduction in the critical Mach number, which adversely affects the aerodynamic performance.
(2) Aerodynamic coefficient Kriging surrogate models are established for Ma = 0.74 and 0.76. The prediction deviations of the aerodynamic coefficients are fitted and investigated, and the R2 of the surrogate models is more than 0.95, while the RMSE is less than 0.1. The results of the Kriging surrogate model and CFD simulation are within 2%, satisfying the accuracy requirements.
(3) The results show that the optimized airfoil lift increased by 7.55% and 7.37%, respectively, and the lift-to-drag ratio increased by 6.97% and 10.27%, respectively. The feasibility and reliability are demonstrated. The method proposed in this paper reduces the computational cost of numerical simulation.
(4) In this paper, a predictive optimization method that combines the Kriging surrogate model with NSGA-II optimization is proposed. Theoretically, the lift and the lift-to-drag are improved to some extent. When optimizing solely based on given objectives, there is a risk of rapid separation of the boundary layer due to the forward movement of shock waves on the upper wing surface. Subsequently, we further investigate the optimization algorithm and expand the selection of variable curvature airfoils and study the impact of parameters such as camber position and camber radius on wing aerodynamics. In addition, wind tunnel tests will be conducted to verify the accuracy of the theoretical model.