This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
1. Introduction
Due to its high maneuverability and other advantages, the canard-controlled missile is widely used in vertical launching air-defense missiles. Vertical launching missiles may face the issue of uncertain roll angle (
A revolutionary body is a commonly used simplified model in research on missiles with large angles of attack. It has been confirmed in numerous studies that slender bodies produce unexpected lateral forces at sufficiently large angles of attack [1, 2]. At such large angles of attack, the vortices at the head of the slender body transition from symmetric to asymmetric as they develop downstream [3]. Micro disturbances on the head, such as surface roughness [4], manufacturing defects [5, 6], and even dust particles [7], are considered one of the causes of the natural asymmetry phenomenon. Extensive simulations and experiments have demonstrated the influence of micro disturbances on the asymmetric direction of the flow field by introducing micro perturbation [8–10] or changing the roll angle [11, 12]. However, when additional disturbances on the head are significantly stronger than surface roughness and manufacturing defects, the impact of micro disturbances on the flow field diminishes. Therefore, introducing additional disturbances on the head becomes an important means of controlling asymmetrical flow at large angles of attack. Asymmetric disturbances on the head determine the asymmetric direction of the flow field by eliminating the bistable phenomenon of asymmetric flow around the slender body [13, 14], while symmetric disturbances on the head, such as tip disturbance [15], nose vortex generators [16], and weak jets [17, 18], enhance flow field stability, allowing a stable symmetric flow field to extend to larger angles of attack. Additionally, periodic variations in disturbances, such as rotating tips perturbation [19], eliminate unexpected lateral forces through time-averaging techniques.
Compared to the vortices generated by weak disturbances, which are significantly weaker than the head vortices, the vortices produced by strong disturbance sources in missiles, such as canards, strakes, drag rings, and wings, are comparable to or even stronger than the head vortices. The introduction of strong disturbance induces interaction between vortices and interaction between vortices and the wall, further affecting the aerodynamic force of the missile. Johnson et al. [20] conducted research on wing-body combinations and found that the interaction between asymmetric vortices and wing significantly affects lateral force and roll moment and may lead to severe control issues. Kumar [21] carried out wind tunnel experiments on a slender body with a circular drag ring and found that adding a ring to the nose significantly reduces lateral force but greatly increases the roll moment. Karn et al. [22] investigated the unsteady characteristic of a slender body with circular drag rings using CFD methods and found that the ring reconstructs the flow field around the slender body and generates significant fluctuating loads behind the ring. Yuan et al. [23] conducted experiments to measure the pressure distribution and flow field structure of a slender body with strakes at
Current researches on missiles at large angles of attack typically focused on explaining and managing the natural asymmetry of slender revolutionary bodies, with less emphasis on the multivortex structures that arise from canard. To investigate the causes of unexpected aerodynamic changes in canard-configured missiles at low speeds and large angles of attack, this study focuses on the interaction between canard vortices and body vortices at different roll angles, examining the evolution of these multivortex structures and their effects on missile aerodynamics. This study is aimed at clarifying the mechanisms behind unexpected aerodynamic variations in canard-configured missiles at large angles of attack and informing flight control strategies in these conditions. The existing literature covers the study of aerodynamic mechanisms in simple models and the engineering applications of aerodynamics in complex models. It lacks an in-depth look at the aerodynamic mechanisms within complex models. This paper enhances the understanding of aerodynamic forces in canard-configured missiles at large angles of attack and extends the research on missile aerodynamics from basic revolutionary bodies to the more practically relevant canard configurations, offering valuable insights for engineering applications.
2. Methodology
2.1. Governing Equations
After decades of development, numerical simulation is now considered to be a mature and reliable means of analyzing the characteristics of flow, which can yield more information on the flow field than experiments. The Reynolds-averaged Navier–Stokes (RANS) equations use a turbulence model to solve for the viscous term and are closed to reduce the difficulty of obtaining the solution. The
The equations of continuity, momentum, and energy can be written in differential form as follows:
The Boussinesq hypothesis [35] is used to relate the Reynolds mean pressure to the mean velocity.
The
Equation of effective diffusion term:
Among them,
In Equation (5),
Definition of coefficient a:
Turbulent kinetic energy divergence term:
In incompressible flows,
Divergence in
In incompressible flows,
Correction for orthogonal divergence:
The constant term in this equation is as follows:
2.2. Validation and Verification
2.2.1. Validation of Numerical Algorithm
The k-ω SST turbulence model balances the model’s effectiveness with computational efficiency. Spalart [36] recognizes the practical value of unsteady Reynolds-averaged Navier–Stokes (URANS) equations. As depicted in Figure 1, 3D URANS simulation is capable of capturing the main vortex structures in the flow field and provides drag coefficient
[figure(s) omitted; refer to PDF]
Table 1
Comparison of methods for solving viscous terms.
Method | Advantage | Disadvantaged |
LES | High precision | Requires very high precision mesh and very high computational costs |
DES | Reduce the high computational cost of LES in the near-wall region while maintaining the accuracy of large-scale vortices | The parameters need to be carefully adjusted to ensure a smooth transition |
RANS | Mature, stable, and low computational cost | The transient structure of turbulence cannot be captured |
RSM | Provides anisotropic and nonlinear effects of turbulent pulsation | The large number of equations increases the uncertainty of the model |
A validation methodology for numerical approaches that align with the techniques detailed in reference [26] was employed in this study. To verify the feasibility of the numerical algorithm for the canard-controlled missile at a large angle of attack, the three-dimensional flow around a cylinder and an airfoil under a large angle of attack were compared with references. The case in Ref. [39] was used to examine the flow around a cycle cylinder. The diameter of the cylinder was 150 mm. Symmetric boundary conditions were used on both sides. The velocity of incoming flow was 35 m/s, pressure was 0.1 MPa, temperature was 60°C, turbulence was 0.7%, and the Reynolds number was
[figure(s) omitted; refer to PDF]
The distribution of the coefficient of surface pressure of the NACA0021 airfoil at
[figure(s) omitted; refer to PDF]
2.2.2. Model and Verification
The model used in the calculations and experiments is shown in Figure 5(a). The canards and fins were added based on a pointed arch-shaped slender body with a slenderness ratio of 17.5. To demonstrate the multivortex interactions that may occur under the influence of large angles of attack and roll and to derive universally applicable conclusions, this study employs a common design for canard-configured missiles. These designs encompass trapezoidal canard planforms and hexagonal airfoils with the tips of leading and trailing edges removed. For canard-configured missiles launched in a tube, the canards are restricted to a location on the ogive segment, with a limitation that the canard span should not exceed the missile diameter. The chord length and axial position of the canards are determined based on the requirement of static stability. When the roll angle of the model was 0°, the form of the canard was a + shape. The diameter of the model was
[figure(s) omitted; refer to PDF]
The domain of calculation is shown in Figure 5(b). The front face, 55.6D from the vertex, and the lower half of the cylinder were used as the inlet of velocity. The back face, 83.3D from the tail, and the upper half of the cylinder were used as the outlet of pressure. A Cartesian coordinate system with the origin of the vertex was established. The
Table 2
The definitions of the configurations under different roll angles.
Configuration | A | B | C | D | E |
To ensure the availability of data, it was necessary to verify mesh independence. Calculations were performed with a coarse mesh with 3.05 million cells, a medium-mesh with 6.32 million cells, and a fine mesh with 8.71 million cells, and their results were compared. In addition, the
Table 3
Comparison of results of three kinds of cells.
Number of cells | CA | CN | CZ | Cm | GCI of CA | GCI of CN | GCI of CZ | GCI of Cm |
3.05 million | 0.1648 | 13.7501 | −1.1346 | −6.2406 | ||||
6.53 million | 0.1216 | 15.3227 | −1.4467 | −7.9860 | 0.8362 | 0.2415 | 4.1987 | 0.5143 |
8.71 million | 0.1255 | 15.2706 | −1.5742 | −7.9132 | 0.0841 | 0.0092 | 0.2185 | 0.0248 |
2.2.3. Experimental Validation
To further validate the results of numerical calculations, PIV experiments are conducted to obtain the velocity distribution of the cross-section for comparison with the results obtained from numerical simulations. The PIV experiments were carried out in an open-jet return-flow wind tunnel with turbulence of lower than 1%. The experimental arrangement is shown in Figure 6. A 532 nm solid double-pulse laser with a maximum frequency of 10 Hz and a laser controller was connected through cables and cooling pipes. Following reflection by the light guide arm, the laser generator generated sheet light through refraction from a group of laser lenses. The synchronizer was, respectively, connected to the computer, the laser controller, and the CCD camera. The CCD camera and the computer were also connected through a cable and an image acquisition card. In the experiment, the particles generated by the particle generator for imaging were dispersed into the wind tunnel and diffused evenly. The computer simultaneously triggered the generation of a solid pulse laser and the exposure of the CCD camera through the synchronizer. The double-pulse laser triggered two laser pulses each time, and the interval between them was 10 μs. The CCD camera captured two frames of images of particles through two instances of exposure and transferred them to the Dantec PIV software for cross-correlation analysis. The two frames of images are divided into multiple interrogation zones. Within each interrogation zone, cross-correlation statistical analysis is used to obtain the average displacement vector of particles within the zone. This further allows for the calculation of the velocity vector field across the entire solution domain.
[figure(s) omitted; refer to PDF]
Lazar et al. [47] categorizes the uncertainties in PIV measurements into the equipment, particle dynamics, data acquisition, and image analysis and, subsequently, provides a method for quantifying the uncertainty in PIV measurements. In the context of the current PIV measurements, uncertainties are quantified using the method from the literature [47]. Since the experiment was conducted in an open-return low-speed wind tunnel without observation windows, optical distortions caused by window glass are not present and, thus, are neglected. The velocity uncertainty induced by the equipment is as follows:
In the equation,
Table 4
Parameters and uncertain values.
Category | Parameter | Description | Values | |
Calibration | Calibration scale physical length | 36 mm | 1 mm | |
Calibration scale image plane length | 2048 pixels | 1 pixel | ||
Distance from calibration scale to lens | 1 m | 0.5 mm | ||
Timing | Laser pulse timing | 200 ns | 2 ns | |
Accuracy of delay generator | 200 ns | 2 ns |
From Equation (32), the velocity uncertainty caused by the equipment is obtained as 0.075 m/s, with a relative uncertainty of 0.54%.
The uncertainty in particle dynamics primarily stems from the particle slip phenomenon caused by large velocity gradients or strong velocity fluctuations, which is predominantly observed in compressible flows. In PIV experiments, atomized paraffin oil is utilized as tracer particles, with a density of 780 kg/m3 and droplet diameters of approximately 2 to 3 μm. Given the minuscule size of the droplets, the effect of gravity is neglected. Due to the significantly higher density of paraffin oil relative to air, the forces acting on the particles can be simplified to Stokes’ drag. The particle slip velocity can be expressed as follows:
In the equation,
Sampling of each set of images can be considered as being conducted independently within the same spatial window [48]. The samples can be assumed to follow a normal distribution [49, 50]. For samples that fall within two standard deviations from the mean, the confidence level is 95%. For a sample size of
In the equation,
The uncertainty in image analysis arises from the image processing techniques and the PIV algorithms used. Factors such as particle intensity, particle density, image size, turbulence fluctuations, velocity gradients, background noise, and the size of the interrogation windows can all impact the level of uncertainty. The PIV algorithm employs bilinear interpolation on the images. To achieve a smooth flow field, noise is filtered and corrected by comparing the average values of adjacent points. In this experiment, a PIV acquisition and processing system from Dantec Inc was employed. Cross-correlation computations were performed on the 50 sets of images using interrogation zones of
The error propagation formula for uncertainty is given as follows [47]:
In the equation,
Sources of error in PIV measurements are commonly categorized into installation and alignment errors, system component errors, flow-induced errors, and measurement errors [51, 52]. Installation and alignment errors involve model machining inaccuracies, discrepancies in mounting angles, and the precise alignment of the laser plane. System component errors encompass laser synchronization delays and image distortion. Flow-induced errors are attributed to the particle slip phenomenon, which can arise from significant velocity gradients, intense velocity fluctuations, or high streamline curvature. Within the uncertainty analysis, the influence of particle slip at low velocities has been discussed and is deemed insignificantly small, thus ignorable. Measurement errors include invalid measurements, systemic error, and random error. In the experiment, the sideslip angle was utilized as a proxy for the angle of attack. To rigorously mitigate errors, a laser level was employed in the calibration of the model and experimental apparatus, ensuring that the model’s longitudinal axis was aligned within a horizontal plane. The sideslip angle was fine-tuned with an accuracy of 0.05° using a precision angle adjustment mechanism, and a similar calibration approach with a 0.5° precision was applied to the roll angle, aided by a laser level. To address random measurement errors, the average and standard deviation of 50 sets of PIV images were computed, and the resulting curves and error bands were charted and compared with CFD results as depicted in Figure 7.
[figure(s) omitted; refer to PDF]
Configuration C was used for the PIV experiment at
3. Results and Discussion
3.1. The Evolution of the Multivortex Structure
At large angles of attack, the flow around a slender body typically consists of a pair of main vortices and multiple smaller secondary vortices. However, when there is a canard present on the head, it introduces additional pairs of vortices. These corotating vortices tend to merge due to viscous diffusion. This merging process occurs within a certain downstream distance of the canard, which is defined as the region of vortex interaction. The multivortex structure undergoes multiple cycles of merging until it eventually converges into a simpler structure dominated by two main vortices. Throughout this process, the reconstruction of the flow field structure from a multivortex configuration to a steady Karman-like vortex structure affects the surface pressure coefficient, the downstream flow field structure, and the overall aerodynamic force coefficients.
3.1.1. The Merging of Near-Wall Streamwise Vortices
In canard-controlled missile, there are multiple vortices present such as the tip vortices of canard. Unlike at low angles of attack where these vortices operate independently, neighboring vortices with the same rotational direction may merge due to the effects of viscous diffusion at large angles of attack. Within the region of vortex interaction, typical flow phenomena such as strong vortex entrainment of weaker vortices, two-vortex merging, and three-vortex merging occur, leading to the transformation of the flow field into relatively simpler separated flows. These vortex interactions are primarily observed in the range of
The larger the value of
In the theoretical framework proposed by Dizes and Verga [28], the merging of two parallel vortices with equal intensity begins when their relative distance, normalized by the vortex radius, reaches 0.22. However, in practical flow situations, it is rare to have two parallel vortices with exactly equal intensities. For configuration D at
[figure(s) omitted; refer to PDF]
In the region of vortices interaction, it is also possible to observe the phenomenon of a strong vortex entraining a weaker vortex and the participation of a third vortex during the merging process of two vortices, as depicted in Figure 9. In Figure 9(a), the vorticity distribution of configuration A at
[figure(s) omitted; refer to PDF]
Figure 10 presents the variation in vorticity
[figure(s) omitted; refer to PDF]
3.1.2. Interaction Between Vortices and Walls
In the region of vortex interaction, there are various cases of interaction between vortices and the wall, such as the reattachment and reseparation of vortex, the Conda effect of curved surfaces on near-wall vortices, the interaction between vortices and canards, and the influence of vortices on surface pressure coefficients.
Figure 11 illustrates the process of reattachment and separation of a canard separation vortex. The definition of the dimensionless velocity
[figure(s) omitted; refer to PDF]
The interaction between the vortex and canard at large
[figure(s) omitted; refer to PDF]
The influence of the vortex on the surface
[figure(s) omitted; refer to PDF]
The multiple vortices generated by canards in a canard-controlled missile create a complex flow field structure at the head. At large angles of attack, these vortices interact with each other through vortex merging and other interactions, eventually forming a structure of a flow field similar to the flow field around a slender body. In addition to the direct impact of the low-pressure vortex on the body, strong separated vortices on one side induce high velocities and low pressures in near-wall flow, resulting in a large area of low-pressure on the surface, defined as the regions indirectly affected by the vortex. The regions indirectly affected by the vortices, which are located on the side of the body and significantly affect the lateral aerodynamic forces, are larger compared to the regions directly affected by the vortices.
3.2. The Influence of the Interaction of Vortices on Flow Field and Aerodynamic Characteristics of the Missile
3.2.1. The Influence of the Head Vortices on the Downstream Structure of Vortices
The multivortex structure in the flow field of a canard-controlled missile not only affects the surface
[figure(s) omitted; refer to PDF]
To analyze the structure of downstream vortices, the configuration C at
[figure(s) omitted; refer to PDF]
The vortex distribution of configuration C at
Figure 17 presents a comparison of the pressure coefficient across three sections for configuration C at an angle of attack of 55°. Sections 1 and 3 are each oriented at an angle of 60° relative to Section 2. The distribution of the pressure coefficient on both sides corresponds to the surface pressure coefficient distribution depicted in Figure 18, with alternating low and high-pressure regions and a significant drop in the pressure coefficient within the vortex region. On the leeward side, the presence of a strong vortex at the initial separation is associated with a very low-pressure coefficient, which corresponds to the large velocity regions indicated by the contour plots in Figure 16.
[figure(s) omitted; refer to PDF]
The comparison of vortex distribution between configuration A and configuration E at
[figure(s) omitted; refer to PDF]
For symmetric canard-controlled missiles, the merging of vortices in the region of vortices interaction eventually leads to the generation of a symmetric pair of separated vortices. The separation of these symmetric main vortices reduces the velocity gradient in the free shear layer, thereby decreasing the asymmetry in the downstream flow field structure and the asymmetry in the surface pressure distribution on the body. However, for asymmetric canard-controlled missiles, the merging of vortices in the region of vortices interaction results in the generation of a determined asymmetric flow field structure. The downstream flow field structure and the surface
3.2.2. The Integration of the Surface Pressure Coefficient and the Coefficients of Aerodynamic Force
The interaction of vortices and the alternating separation of vortices both result in the changes of surface
In the equation,
A comparative analysis of the
The overall coefficient of lateral force (
[figure(s) omitted; refer to PDF]
By integrating the pressure along the normal direction of the entire missile, the variation of the coefficient of normal force (
[figure(s) omitted; refer to PDF]
4. Conclusions
In this study, the vortex interaction process at the head of a canard missile at large angle of attack was investigated, and its influence on the flow field and aerodynamic forces was analyzed. The main conclusions are as follows:
At large angles of attack, the flow field over a canard-controlled missile generates a complex multivortex structure due to the disturbance caused by the canards. The evolution of the multivortex structure significantly impacts the pressure distribution, the flow field, and the aerodynamic forces. The multivortex structure in the head evolves into a steady-state structure of flow field dominated by a pair of main vortices through vortex merging. This evolutionary process occurs within a certain range in the head.
The configurations with symmetric geometry maintain the symmetry of the flow field during the evolution process. The process of vortex merging is dominated by the separated vortices of the canard, which results in the merged main vortices further away from the body and maintaining downstream symmetry. In comparison to the “+” sharp configuration of canard missiles, the X-sharp configuration exhibits stronger main vortices after mergence in the region of vortices interaction. It maintains a longer symmetric region in the downstream flow field, demonstrating higher flow field stability.
The configurations with asymmetric geometry induce inconsistent vortex intensities and positions on both sides, leading to a strong asymmetry in the main vortices after merging. These intense vortices not only directly influence the surface pressure distribution near the vortices but also create large areas of low pressure on the sides by affecting the velocity. The asymmetric vortices in the downstream flow field cause alternating separation and shedding on the left and right sides, resulting in fluctuating sectional lateral force coefficients. As the angle of attack increases, the amplitude of the fluctuations in the sectional lateral force coefficient increases and gradually dominates the lateral force coefficient, leading to a significant decrease in the lateral force coefficient.
As a comparison, the symmetric configuration without canards exhibits a naturally asymmetric flow at large angles of attack. When the main vortex becomes asymmetric, it also induces alternating separation and shedding of vortices on both sides of the body. However, the separation process is delayed compared to the asymmetric canard configuration. This is because the asymmetric canards generate a pair of asymmetric main vortices, leading to earlier detachment of the vortices.
In conclusion, the flow field in the head of a canard-controlled missile exhibits a multivortex structure. This multivortex structure not only affects the surface pressure coefficients within the region of vortex interaction but also determines the downstream flow field structure, pressure distribution, and dominates the lateral force.
Author Contributions
Conceptualization: Kai Wei and Shaosong Chen; methodology: Kai Wei and Yihang Xu; validation: Yihang Xu and Xujian Lyu; formal analysis: Dongdong Tang; investigation: Dongdong Tang; resources: Shaosong Chen; data curation: Qing Chen; writing—original draft preparation: Kai Wei; writing—review and editing: Shaosong Chen, Qing Chen, and Kai Wei; visualization: Xujian Lyu; supervision: Shaosong Chen; project administration: Kai Wei. All authors have read and agreed to the published version of the manuscript.
Funding
The present work is financially supported by the Science and Technology on Underwater Information and Control Laboratory (No. 2021-JCJQ-LB-030-05).
Acknowledgments
The support of the Science and Technology on Underwater Information and Control Laboratory is greatly appreciated.
[1] D. Degani, G. G. Zilliac, "Experimental study of nonsteady asymmetric flow around an ogive-cylinder at incidence," AIAA Journal, vol. 28 no. 4, pp. 642-649, DOI: 10.2514/3.10441, 1990.
[2] W. Letko, A Low-Speed Experimental Study of the Directional Characteristics of a Sharp-Nosed Fuselage Though a Large Angle-of-Attack Range at Zero Angle of Sideslip, 1953.
[3] D. Degani, "Development of nonstationary side forces along a slender body of revolution at incidence," Physical Review Fluids, vol. 7 no. 12,DOI: 10.1103/PhysRevFluids.7.124101, 2022.
[4] C. Moskovitz, F. Dejarnette, R. Hall, "Effects of nose bluntness, roughness, and surface perturbations on the asymmetric flow past slender bodies at large angles of attack," AIAA 7th Applied Aerodynamics Conference, .
[5] D. Degani, "Effect of geometrical disturbance on vortex asymmetry," AIAA Journal, vol. 29 no. 4, pp. 560-566, DOI: 10.2514/3.59929, 1991.
[6] P. M. Hartwich, R. M. Hall, M. J. Hemsch, "Navier-Stokes computations of vortex asymmetries controlled by smallsurface imperfections," Journal of Spacecraft and Rockets, vol. 28 no. 2, pp. 258-264, DOI: 10.2514/3.26239, 1991.
[7] G. G. Zilliac, D. Degani, M. Tobak, "Asymmetric vortices on a slender body of revolution," AIAA Journal, vol. 29 no. 5, pp. 667-675, DOI: 10.2514/3.59934, 1991.
[8] Z. Qi, S. Zong, Y. Wang, Q. Li, J. Wang, "Sources of asymmetric flow over the blunt-nose slender body," European Journal of Mechanics - B/Fluids, vol. 75, pp. 372-381, DOI: 10.1016/j.euromechflu.2018.10.010, 2019.
[9] Z. Qi, Y. Wang, H. Bai, Y. Sha, Q. Li, "Effects of micro-perturbations on the asymmetric vortices over a blunt-nose slender body at a high angle of attack," European Journal of Mechanics - B/Fluids, vol. 68, pp. 211-218, DOI: 10.1016/j.euromechflu.2017.06.005, 2018.
[10] Y. Zhu, H. Yuan, C. Lee, "Experimental investigations of the initial growth of flow asymmetries over a slender body of revolution at high angles of attack," Physics of Fluids, vol. 27 no. 8,DOI: 10.1063/1.4928313, 2015.
[11] Z. Qi, S. Zong, Y. Wang, "Bi-stable asymmetry on a pointed-nosed slender body at a high angle of attack," Journal of Applied Physics, vol. 130 no. 2,DOI: 10.1063/5.0056494, 2021.
[12] P. Dexter, B. Hunt, "The effect of roll angle on the flow over a slender body of revolution at high angles of attack," AlAA 19th Aerospace Sciences Meeting, .
[13] L. Zhao, Y. Wang, Z. Qi, "Investigation of periodic characteristics of perturbed flow Over a slender body," Heliyon, vol. 9 no. 5, article e16194,DOI: 10.1016/j.heliyon.2023.e16194, 2023.
[14] S. Zong, Y. Wang, Z. Qi, "Structural control of asymmetric forebody vortices over a slender body," Physics of Fluids, vol. 33 no. 11, pp. 112021-117111, DOI: 10.1063/5.0068576, 2021.
[15] D. Degani, M. Tobak, "Experimental study of controlled tip disturbance effect on flow asymmetry," Physics of Fluids, vol. 4 no. 12, pp. 2825-2832, DOI: 10.1063/1.858338, 1992.
[16] J. Zhai, W. Zhang, Q. C. Gao, Y. H. Zhang, Z. Y. Ye, "Side force control on slender body by vortex generators," Chinese Journal of Theoretical and Applied Mechanics, vol. 46 no. 2, pp. 308-312, 2014.
[17] G. H. Kamakoli, K. Mansour, "Determining the side force appearance and its magnitude over a slender body at high angle of attack," Thermophysics and Aeromechanics, vol. 28 no. 6, pp. 919-930, DOI: 10.1134/S0869864321060160, 2021.
[18] E. Lee, Y. Huang, B. Vukasinovic, A. Glezer, "Aerodynamic flow control of an unstable slender cylindrical body at high incidence," AIAA AVIATION 2021 Forum,DOI: 10.2514/6.2021-2610, .
[19] B.-F. Ma, H. Yu, X.-Y. Deng, "Dynamic responses of asymmetric vortices over slender bodies to a rotating tip perturbation," Experiments in Fluids, vol. 57 no. 4,DOI: 10.1007/s00348-016-2139-3, 2016.
[20] G. Johnson, "Overall forces and moments on wing-bodies at high incidence," 15th Atmospheric Flight Mechanics Conference,DOI: 10.2514/6.1988-4354, .
[21] P. Kumar, "Rolling moment of slender body at high incidence for air to air missile rocket applications," Defence Science Journal, vol. 70 no. 1, pp. 18-22, DOI: 10.14429/dsj.70.13612, 2020.
[22] P. K. Karn, S. Das, P. Kumar, "Effect of ring on the flow unsteadiness of slender body at Α =50," Journal of the Brazilian Society of Mechanical Sciences and Engineering, vol. 45 no. 3,DOI: 10.1007/s40430-023-04101-4, 2023.
[23] Q. Yuan, Y. Wang, Z. Qi, "Strakes effects on asymmetric flow over a blunt-nosed slender body at a high angle of attack," Journal of Fluids Engineering, vol. 141 no. 6,DOI: 10.1115/1.4041815, 2019.
[24] S. W. Xu, X. Y. Deng, "Effects of angle of attack on wing rock motion induced by the flows over slender body with low swept wing," Science China. Physics, Mechanics & Astronomy, vol. 58 no. 4,DOI: 10.1007/s11433-014-5630-y, 2015.
[25] D. Maynes, G. A. Gebert, "High-angle-of-attack aerodynamics of a missile geometry at low speed," Journal of Spacecraft and Rockets, vol. 36 no. 5, pp. 772-774, DOI: 10.2514/2.3493, 1999.
[26] F. Wang, J. Liu, H. Qin, Y. Song, L. Chen, H. Jing, "Unsteady aerodynamic characteristics of slender body at extra-wide angle-of-attack range," Aerospace Science and Technology, vol. 110, article 106477,DOI: 10.1016/j.ast.2020.106477, 2021.
[27] A. R. Nielsen, M. Andersen, J. S. Hansen, M. Brøns, "Topological bifurcations of vortex pair interactions," Journal of Fluid Mechanics, vol. 917,DOI: 10.1017/jfm.2021.191, 2021.
[28] S. Le Dizes, A. Verga, "Viscous interactions of two co-rotating vortices before merging," Journal of Fluid Mechanics, vol. 467, pp. 389-410, DOI: 10.1017/S0022112002001532, 2002.
[29] P. Meunier, T. Leweke, "Three-dimensional instability during vortex merging," Physics of Fluids, vol. 13 no. 10, pp. 2747-2750, DOI: 10.1063/1.1399033, 2001.
[30] M. J. B. Dam, J. S. Hansen, M. Andersen, "Viscous merging of three vortices," European Journal of Mechanics - B/Fluids, vol. 99, pp. 17-22, DOI: 10.1016/j.euromechflu.2022.12.014, 2023.
[31] C. McKenna, M. Bross, D. Rockwell, "Structure of a streamwise-oriented vortex incident upon a wing," Journal of Fluid Mechanics, vol. 816, pp. 306-330, DOI: 10.1017/jfm.2017.87, 2017.
[32] Z. Sun, G. Yunsong, H. Zhao, "Experimental investigation on the streamwise vortex–surface interaction," Journal of Visualization, vol. 22 no. 3, pp. 477-488, DOI: 10.1007/s12650-018-00544-3, 2019.
[33] F. R. Menter, "Zonal two equation K-W turbulence models for aerodynamic flows," 23rd Fluid Dynamics, Plasmadynamics, and Lasers Conference, .
[34] F. R. Menter, "Two-equation eddy-viscosity turbulence models for engineering applications," AIAA Journal, vol. 32 no. 8, pp. 1598-1605, DOI: 10.2514/3.12149, 1994.
[35] J. Boussinesq, Essai Sur La Théorie Des Eaux Courantes, 1877.
[36] P. Spalart, "Reflections On RANS modelling," In the 3rd Symposium on Hybrid RANS-LES Methods, .
[37] D. Teso-Fz-Betoño, M. Juica, K. Portal-Porras, U. Fernandez-Gamiz, E. Zulueta, "Estimating the reattachment length by realizing a comparison between URANS K-omega SST and LES WALE models on a symmetric geometry," Symmetry-Culture and Science, vol. 13 no. 9,DOI: 10.3390/sym13091555, 2021.
[38] A. Dejoan, Y. J. Jang, M. A. Leschziner, "Comparative LES and unsteady RANS computations for a periodically-perturbed separated flow over a backward-facing step," Journal of Fluids Engineering, vol. 127 no. 5, pp. 872-878, DOI: 10.1115/1.2033012, 2005.
[39] E. Achenbach, "Distribution of local pressure and skin friction around a circular cylinder in cross-flow up toRe= 5 × 106," Journal of Fluid Mechanics, vol. 34 no. 4, pp. 625-639, DOI: 10.1017/S0022112068002120, 1968.
[40] A. Fage, V. M. Falkner, Further Experiments On the Flow Around a Circular Cylinder, 1931.
[41] W. H. Giedt, "Effect of turbulence level of incident air stream on local heat transfer and skin friction on a cylinder," Journal of the Aeronautical Sciences, vol. 18 no. 11, pp. 725-730, DOI: 10.2514/8.2092, 1951.
[42] W. Haase, M. Braza, A. Revell, DESider–A European Effort on Hybrid RANS-LES Modelling: Results of the European-Union Funded Project, 2009.
[43] A. Gerasimov, Modeling Turbulent Flows with FLUENT, 2006.
[44] S. M. Salim, M. Ariff, S. C. Cheah, "Wall y + approach for dealing with turbulent flows over a wall mounted cube," Progress in Computational Fluid Dynamics, vol. 10 no. 5/6, pp. 341-351, DOI: 10.1504/PCFD.2010.035368, 2010.
[45] S. M. Salim, S. Cheah, "Wall Y+ strategy for dealing with wall-bounded turbulent flows," Proceedings of the international multiconference of engineers and computer scientists, .
[46] A. Kilavuz, M. Ozgoren, L. A. Kavurmacioglu, T. Durhasan, F. Sarigiguzel, B. Sahin, H. Akilli, E. Sekeroglu, B. Yaniktepe, "Flow characteristics comparison of PIV and numerical prediction results for an unmanned underwater vehicle positioned close to the free surface," Applied Ocean research, vol. 129, article 103399,DOI: 10.1016/j.apor.2022.103399, 2022.
[47] E. Lazar, B. De Blauw, N. Glumac, C. Dutton, G. Elliott, "A practical approach to PIV uncertainty analysis," In27th AIAA aerodynamic measurement technology and ground testing conference,DOI: 10.2514/6.2010-4355, .
[48] I. Grant, E. H. Owens, "Confidence interval estimates in PIV measurements of turbulent flows," Applied Optics, vol. 29 no. 10, pp. 1400-1402, DOI: 10.1364/AO.29.001400, 1990.
[49] S. J. Bendat, A. G. Piersol, Random Data: Analysis and Measurement Procedures, 1986.
[50] G. L. Donohue, D. K. McLaughlin, W. G. Tiederman, "Turbulence measurements with a laser anemometer measuring individual realizations," Physics of Fluids, vol. 15 no. 11, pp. 1920-1926, DOI: 10.1063/1.1693803, 1972.
[51] M. Raffe, C. E. Willert, F. Scarano, C. J. Kähler, S. T. Wereley, J. Kompenhans, "PIV uncertainty and measurement accuracy," In Particle Image Velocimetry: A Practical Guide, pp. 203-241, 2018.
[52] A. Sciacchitano, "Uncertainty quantification in particle image velocimetry," Measurement Science and Technology, vol. 30 no. 9, article 092001,DOI: 10.1088/1361-6501/ab1db8, 2019.
[53] K. Wei, S. S. Chen, "Influence of slenderness ratio on the lateral force of a slender body at high angle of attack," Journal of Physics. Conference Series, vol. 2478 no. 12, article 122024,DOI: 10.1088/1742-6596/2478/12/122024, 2023.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Copyright © 2024 Kai Wei et al. This is an open access article distributed under the Creative Commons Attribution License (the “License”), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License. https://creativecommons.org/licenses/by/4.0/
Details



1 School of Energy and Power Engineering Nanjing University of Science and Technology Nanjing China 210094
2 Shanghai Academy of Spaceflight Technology Shanghai Institute of Spacecraft Equipment No. 251 Huaning Road, Shanghai China 200240
3 Aerospace Jiangnan Group Co., Ltd. Jiangnan Design Institute of Mechanical and Electrical Guiyang China 550009