Computation 12 00140
Computation 12 00140
Computation 12 00140
Turbomachines & Fluid Dynamics Laboratory, School of Production Engineering and Management,
Technical University of Crete, 73100 Chania, Greece; aklothakis@gmail.com
* Correspondence: inikolos@tuc.gr
Abstract: This work investigates the aerodynamic performance of a hypersonic waverider designed
to operate at Mach 7, focusing on optimizing its design through advanced computational methods.
Utilizing the Direct Simulation Monte Carlo (DSMC) method, the three-dimensional flow field
around the specifically designed waverider was simulated to understand the shock wave
interactions and thermal dynamics at an altitude of 90 km. The computational approach included
detailed meshing around the vehicle’s critical leading edges and the use of three-dimensional iso-
surfaces of the Q-criterion to map out the shock and vortex structures accurately. Additional
simulation results demonstrate that the waverider achieved a lift–drag ratio of 2.18, confirming
efficient aerodynamic performance at a zero-degree angle of attack. The study’s findings contribute
to the broader understanding of hypersonic flight dynamics, highlighting the importance of precise
computational modeling in developing vehicles capable of operating effectively in near-space
environments.
1. Introduction
The waverider is a distinct configuration of a hypersonic vehicle engineered to
maintain the shock wave generated by the vehicle attached to its leading edge, offering
Citation: Klothakis, A.; Nikolos, I.K.
numerous advantages [1]: (a) this shock wave effectively separates the flow fields on the
Design and Evaluation of a Hypersonic
upper and lower surfaces of the vehicle, thus preventing high-pressure flow from spilling
Waverider Vehicle Using DSMC.
from the lower to the upper surface at the leading edge; (b) it ensures a uniform flow on
Computation 2024, 12, 140. https://
the lower surface, which is optimal for scramjet engine intake; (c) the design of the
doi.org/10.3390/computation12070140
waverider is derived through reverse engineering from a known flow field around a
Academic Editor: Jeffrey S. Marshall conical shape. This design approach will be elaborated on in later sections of this work.
Received: 17 May 2024 The concept of the waverider was first put forward by Nonweiler in 1963 [2]. At that
Revised: 24 June 2024 time, the available technology was not advanced enough to produce a hypersonic air-
Accepted: 3 July 2024 breathing vehicle. Sandlin and Pessin offered an in-depth explanation of the design
Published: 9 July 2024 methodology for waveriders [1]. A broad analysis of the significant design challenges
associated with hypersonic flight vehicles is discussed in the work of Sziroczak and Smith
[3], where two types of vehicles, hypersonic transports and space launchers, are
Copyright: © 2024 by the authors. examined. Although these are distinct from waveriders, they share many design and
Licensee MDPI, Basel, Switzerland. construction challenges. Zhang et al. employed the three-dimensional Class–Shape
This article is an open access article Function Transformation (CST) method to determine the geometric shapes of Hypersonic
distributed under the terms and Gliding Vehicles (HGVs) [4]. Inviscid modeling methods, relying on modified Newton
conditions of the Creative Commons theory and the Prandtl–Meyer equation, were used to assess the aerodynamics of these
Attribution (CC BY) license vehicles. The use of the NSGA-II multi-objective optimization algorithm helped achieve
(https://creativecommons.org/license designs with high lift–drag ratios and substantial volumetric efficiency. In his Ph.D. thesis,
s/by/4.0/). K. Kontogiannis considered the parameterization and manipulation of waverider
For a practical waverider design, addressing the sharpness of the leading edge is
crucial. Thus, a pragmatic design strategy incorporates blunting techniques to manage
this characteristic. Blunting the leading edge and the junctions where two stream surfaces
converge mitigates heating effects and results in a geometry more feasible for
manufacturing. However, blunting does deviate from the ideal waverider design,
Computation 2024, 12, 140 4 of 24
allowing some pressure leakage from the lower to the upper surface, which in turn
increases drag and diminishes the aerodynamic efficiency of the waverider. Studies [33,34]
have demonstrated that while blunting reduces the heat flux experienced by the vehicle,
it adversely impacts the lift–drag ratio (L/D). Despite this, a blunted waverider design can
still offer superior aerodynamic performance compared to other design alternatives. The
extent and nature of the blunting are crucial and require a balance between reducing
thermal stresses and optimizing aerodynamic efficiency. With various possible
configurations for a waverider, there are numerous approaches to designing a blunt
leading edge.
There are two primary methods for blunting the leading edge of a waverider: one
involves removing material from the edge, while the other entails adding material to it
[35]. Tincher and Burnett have proposed that the addition of material to the leading edge
incurs a lower aerodynamic penalty compared to material removal [35]. This approach is
favored because it allows for maintaining more of the original aerodynamic profile,
potentially minimizing the negative impacts on performance that are associated with
leading edge modifications.
randomly assigned within each cell, with specific initial positions, velocities, and energy
states. These steps are cyclically repeated for each timestep to model the time evolution of
the flow. Macroscopic flow properties like velocity, pressure, density, and temperature
are derived from the weighted averages of these microscopic particle properties. For all
the flow simulations discussed in this study, the DSMC method as detailed by Graeme
Bird [36] and as implemented in the Stochastic Parallel Rarefied-gas Time-accurate
Analyzer (SPARTA) at Sandia National Laboratories [38], was utilized to examine the flow
characteristics.
Figure 3. The geometry of the 7-degree half cone used for the calculation of the initial flow field.
As it was previously mentioned, all the simulations were carried out using the open-
source parallel DSMC code SPARTA (Stochastic Parallel Rarefied-gas Time-accurate
Computation 2024, 12, 140 6 of 24
Analyzer), which was developed at Sandia National Laboratories [38]. After computing
the three-dimensional flow field around the 7-degree cone, the flow streamlines were
extracted at specific parallel planes to delineate the shockwave boundaries. Figure 4 offers
a visualization of these calculated flow field streamlines. The shock boundaries are
discernible by the bending of the flow streamlines in both the side and top planes, and by
a change in the color of the streamwise velocity flow field at the rear plane behind the
cone.
Figure 4. Streamlines of the three-dimensional flow field around the 7-degree cone. Side view (top),
top view (middle), rear view (bottom).
Computation 2024, 12, 140 7 of 24
After obtaining the flow planes from the DSMC simulation, these are imported into
a Computer-Aided Design (CAD) software (SolidWorks 2021) to precisely define the
geometry of the waverider. The design process begins at the rear of the vehicle, crafting
curves that are tangent to the observed circular shock shape, with the first characteristic
cross-section displayed in Figure 5 (bottom) and Figure 6. As seen in Table 1, due to the
low Reynolds number the flow falls within the laminar regime.
To ensure the design remains practical for real-world applications, the leading edges
of the waverider are fashioned to be blunt rather than sharp. To accurately control the
bluntness of the leading edges, a four-point Bezier curve is utilized to sculpt the leading-
edge profile. This method provides the flexibility to finely adjust the curvature and thus
tailor the aerodynamic and thermal properties of the edge. Similarly, additional sections
along the waverider’s longitudinal axis are designed at intervals of every 30 cm. Both the
upper and lower surfaces of the waverider are defined by the shockwave boundaries
generated from the initially prescribed 7-degree cone geometry. This approach ensures
that the final waverider design is optimized within the constraints set by the initial flow
conditions and the aerodynamic needs of hypersonic travel.
Computation 2024, 12, 140 8 of 24
Figure 5. Waverider surface along the flow streamlines, in comparison with the initial cone. Top
view (top), side view (middle), and rear view (bottom).
the design of all the sections, a loft operation with tangency constraints is employed to
generate the surfaces of the waverider. To simplify the design process and leverage the
symmetry of the flow field, only half of the final geometry is initially designed. This half
is then mirrored around the symmetry plane to produce the complete vehicle structure.
Figure 7. Surface lofts along the vehicle profiles (top). Vehicle overview without the nose section
(bottom).
Figure 9. (Top): Waverider sections and the complete geometry. (Bottom): Flowchart of the design
methodology.
conditions did not change. As shown in Figure 11 (bottom) the Knudsen number falls
within the upper end of the slip regime or within the transition regime.
Figure 10. The utilized surface mesh. Lower surface (top) and isometric view (bottom).
Computation 2024, 12, 140 14 of 24
Figure 11. Pressure contours around the vehicle (top), Knudsen number of the flow field based on
the waverider length (bottom).
5. Simulation Results
Due to the extensive computational demands associated with simulating hypersonic
flows, several strategies are implemented to optimize the use of computational resources.
One effective technique is the simulation of only half of the waverider’s geometry, with
the resulting flow field then being symmetrically mirrored across a predefined plane. This
approach significantly reduces the computational load without sacrificing the accuracy of
the results. A key objective of these simulations is to confirm the correct positioning of the
shock wave relative to the vehicle’s leading edges. This verification is conducted by
analyzing the pressure contours and Q-criterion plots, which are displayed in Figure 11
and Figure 12, respectively.
The Q-criterion, calculated using Equation (2), involves Ω the antisymmetric
component of the velocity gradient representing vorticity, and 𝑆 , the symmetric
component representing the rate of strain.
Figures 15 and 16 present the streamwise velocity contours and the temperature field
around the vehicle. These figures are crucial for understanding how the vehicle interacts
with the flow at high speeds. In particular, the simulations reveal important thermal
dynamics and fluid flow characteristics. For the specific flow conditions applied in this
study, the temperature field around the vehicle shows significant thermal variations. A
peak temperature of 1400 Kelvin is observed just in front of the nose section of the
waverider, indicating areas where aerodynamic heating is most intense due to the rapid
flow deceleration. This information is vital for assessing the thermal resistance and
structural integrity of the waverider under operational conditions.
Figure 15. Streamwise velocity contours on the symmetry plane of the waverider.
Computation 2024, 12, 140 17 of 24
Figure 16. (Top): total temperature field around the vehicle. (Bottom): rotational temperature field
(on the symmetry plane of the waverider).
In Figures 17–20, various flow properties are illustrated on a plane parallel to the
waverider, with the surface of the vehicle distinctly highlighted by the variations in the
surface pressure. Specifically, Figure 17 displays the pressure contours on this plane,
providing a clear visualization of how the pressure varies in response to the shape and
aerodynamics of the vehicle. Notably, considering that the free-stream pressure is
established at 0.55 Pa, there is a significant pressure increase observed around the nose of
the waverider, rising to an order of magnitude higher than the ambient conditions.
Additionally, the pressure around the leading edges of the waverider is approximately
four times higher than that of the free-stream pressure. This detailed mapping of pressure
distribution is critical for evaluating the aerodynamic performance and structural stresses
experienced by the waverider during operation.
Computation 2024, 12, 140 18 of 24
Figure 18 contains the velocity contours on the same plane. In this figure, the
attachment of the shock to the leading edge and the nose of the vehicle can be clearly
observed. As can be observed in Figure 19, the temperature around the leading edges is
800 K, which is about four times higher than the free-stream temperature of 196 K (with a
peak of around 1400 K at the nose of the vehicle).
Figure 18. Streamwise velocity contours on a horizontal plane parallel to the vehicle.
In Figure 20, the Q-criterion is meticulously illustrated on the same plane parallel to
the waverider, effectively capturing the entire field of the attached shock wave. This
visualization underscores the precise areas where the shock remains attached to the
surface of the waverider, a critical aspect of its aerodynamic design. Additionally, the
positive values of the Q-criterion in this figure indicate the formation of two distinct
vortices, located symmetrically, one on each side, at the tip of the vehicle’s rear section.
These vortices are pivotal in understanding the flow dynamics at the rear of the waverider,
particularly how they influence the overall stability and control of the vehicle at high
speeds. For a more detailed observation of these vortices, Figure 21 provides an enhanced
view where the vorticity at the rear of the waverider is plotted. In this figure, the two
vortices are distinctly highlighted in red, offering a stark contrast against the rest of the
flow field. This color differentiation not only emphasizes the location and intensity of the
vortices but also aids in analyzing their structure and interaction with the surrounding
flow. The depiction of these vortices in such vivid detail is crucial for engineers and
designers to evaluate the aerodynamic interactions at play and to refine the waverider’s
design for optimized performance.
Given the intricate design of the waverider, it was deemed necessary to employ the
three-dimensional iso-surfaces of the Q-criterion to fully capture the complex spatial
Computation 2024, 12, 140 20 of 24
characteristics of the shocks and the structured vortices surrounding the vehicle. Figure 22
illustrates these three-dimensional Q-criterion contours enveloping the vehicle, with the
contours colored by vorticity magnitude to emphasize the variations in the rotational flow
behavior. The transparent grey iso-surfaces distinctly outline the external shape of the shock
waves, providing a comprehensive view of their three-dimensional form. Notably, parts of
the shock wave close to the compression surface are also visible in the same figure, offering
a detailed look at how the shock interacts with the vehicle’s structure.
In Figure 23, these contours are similarly plotted but are instead colored by velocity
magnitude to highlight the differences in flow speed. As depicted in this figure,
approximately 30 cm downstream from the nose of the vehicle—where the compression
surface becomes steeper—the vortices detach from the main flow and continue
downstream towards the higher pressure regions at the back of the vehicle. This
separation is critical for understanding the aerodynamic behavior of the waverider.
Additionally, as the angle of the compression surface increases towards the vehicle’s rear,
a noticeable pressure rise occurs, and the flow velocity significantly reduces, slowing
down by about 1000 m/s within just a few centimeters. These detailed visualizations and
analyses of flow dynamics are essential for assessing the performance implications of the
vehicle’s design under operational conditions.
Figure 23. Overview of the three-dimensional Q-criterion contours around the vehicle, colored by
velocity magnitude.
Computation 2024, 12, 140 21 of 24
To determine the aerodynamic forces acting on the waverider, the shear stresses and
pressure components across its three-dimensional surface were meticulously calculated.
These calculations facilitate the precise measurement of lift and drag forces per unit
surface area, with the results displayed in Figure 24. The analysis of this figure reveals
that the maximum drag primarily occurs at the tip of the nose, while a significant amount
of lift is generated in the area just below the nose. By integrating the forces measured
across all the surface cells, the total lift and drag for the waverider were computed,
yielding a lift–drag ratio (L/D) of 2.18. According to existing studies, this L/D ratio is
typical for a waverider operating at a 0-degree angle of attack [44], indicating that a value
above 2.0 is indicative of effective aerodynamic performance under the chosen design
conditions.
The Knudsen number contours, based on the waverider length, are depicted in Figure
11 (bottom), while the Mach number contours at the plane of symmetry are depicted in
Figure 25. As demonstrated the Knudsen number around the vehicle is within the
transitional regime, which justifies the use of the DSMC method.
For this study, only a single angle of attack was investigated; however, the authors in
[44] suggest that a slight increase in the angle of attack could potentially enhance the L/D
ratio by two to three times, thereby significantly improving performance. It is also
noteworthy that waveriders are commonly analyzed at altitudes between 25 km and 40
km in the open literature, unlike the 90 km altitude considered in this study. The elevated
altitude impacts the generation of lift due to different fluid dynamics, necessitating a
larger compression surface than what is typically seen in waveriders operating at lower
altitudes. This adaptation is a critical design modification that enhances the vehicle’s
performance in the rarified atmospheric conditions encountered at high altitudes.
Figure 24. Overview of the lift (bottom) and drag (top) per unit surface, exerted on the waverider’s
lower surface.
Computation 2024, 12, 140 22 of 24
Figure 25. Mach number contours around the vehicle at the plane of symmetry.
6. Conclusions
In the comprehensive design and analysis of a Mach 7 waverider, intricate
computational methodologies were employed to accurately simulate the vehicle’s
aerodynamic properties. After applying the osculating cone methodology, for the
definition of the waverider’s geometry, a precise computational mesh was constructed to
capture the complex surface geometry, particularly around the waverider’s critical
leading edges. An advanced simulation technique, the Direct Simulation Monte Carlo
(DSMC) method, was used to resolve the three-dimensional flow field around the vehicle.
This simulation not only detailed the pressure and thermal conditions surrounding the
waverider but also highlighted the structural interactions of the shock waves, validated
through the visualizations of streamwise velocity and temperature fields.
Further refinement of the simulation model involved creating the three-dimensional
iso-surfaces of the Q-criterion to intricately map out the shock and vortex structures
around the waverider. This allowed for an in-depth examination of the flow dynamics,
particularly the attachment of shock waves and the formation of vortices at the critical
sections of the vehicle. Such detailed analyses were essential for confirming the
aerodynamic efficiency of the design. Additionally, the use of layered meshing techniques
facilitated the accurate capture of boundary layer phenomena, crucial for assessing the
vehicle’s performance under the extreme conditions of hypersonic flight.
The culmination of these extensive simulations was the calculation of lift and drag
forces, where the lift–drag ratio was determined to be 2.18, indicative of efficient
aerodynamic performance for a waverider at a zero-degree angle of attack. These findings
are significant, as they not only demonstrate the waverider’s capability under the
stipulated design conditions but also suggest potential improvements for future designs.
For instance, adjustments in the angle of attack could further optimize the vehicle’s
performance. This study’s insights into high-altitude aerodynamics, where different fluid
dynamics necessitate modifications such as larger compression surfaces, pave the way for
the enhanced designs of future hypersonic vehicles operating in near-space environments.
Author Contributions: Conceptualization, A.K. and I.K.N.; methodology, A.K. and I.K.N.; software,
A.K.; validation, A.K. and I.K.N.; resources, A.K.; writing—original draft preparation, A.K.;
writing—review and editing, I.K.N.; visualization, A.K.; supervision, I.K.N.; project administration,
I.K.N.; funding acquisition, I.K.N. All authors have read and agreed to the published version of the
manuscript.
Computation 2024, 12, 140 23 of 24
Funding: This study was funded by the Integrated Air and Missile Defence Centre of Excellence
(IAMD COE) and any intellectual property resulting from the work covered by this will be property
of the IAMD COE. This paper reflects only the IAMD COE policies and its author(s)’ positions and
it is not intended to create any legal obligations nor does it reflect NATO’s policies or positions, or
engage NATO in any way.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author due to the fact that this project was funded by NATO IAMD-COE, and
sharing the data requires their explicit permission.
Acknowledgments: The authors would like to thank the UKTC and EPSRC for computational time
on the UK supercomputing facility ARCHER2 via projects EP/R029326/1 and EP/X035484/1.
Conflicts of Interest: The authors declare no conflicts of interest.
References
1. Sandlin, D.R.; Pessin, D.N. Aerodynamic Analysis of Hypersonic Waverider Aircraft; NASA-CR-192981; NASA: Washington, DC,
USA, 1993.
2. Nonweiler, T.R.F. Delta wings of shape amenable to exact shock wave theory. J. R. Aero. Soc. 1963, 67, 39.
3. Sziroczak, D.; Smith, H. A review of design issues specific to hypersonic flight vehicles. Prog. Aerosp. Sci. 2016, 84, 1–28.
4. Zhang, T.T.; Wang, Z.G.; Huang, W.; Yan, L. Parameterization and optimization of hypersonic-gliding vehicle configurations
during conceptual design. Aerosp. Sci. Technol. 2016, 58, 225–234.
5. Kontogiannis, K. On Developing Efficient Parametric Geometry Models for Waverider-Based Hypersonic Aircraft. Ph.D. Thesis,
University of Southampton, Southampton, UK, 2017.
6. Son, J.; Son, C.; Yee, K. A Novel Direct Optimization Framework for Hypersonic Waverider Inverse Design Methods. Aerospace
2022, 9, 348.
7. Flower, J.W. Configurations for high supersonic speeds derived from simple shockwaves and expansions. R. Aeronaut. Soc. J.
1963, 67, 287–290.
8. Kuchemann, D. The Aerodynamic Design of Aircraft: A Detailed Introduction to the Current Aerodynamic Knowledge and Practical Guide
to the Solution of Aircraft Design Problems; Pergamon Press: Oxford, UK, 1978.
9. Bowcutt, K.G. Multidisciplinary optimization of airbreathing hypersonic vehicles. J. Propuls. Power 2001, 17, 1184–1190.
10. Cockrell, C.E., Jr. Interpretation of waverider performance data using computational fluid dynamics. J. Aircr. 1994, 31, 1095–
1100.
11. Corda, S.; Anderson, J., Jr. Viscous optimized hypersonic waveriders designed from axisymmetric flow fields. In Proceedings
of the 26th Aerospace Sciences Meeting, Reno, NV, USA, 14 January 1988.
12. Nonweiler, T.R.F. Aerodynamic problems of manned space vehicles. R. Aeronaut. Soc. J. 1959, 63, 521–528.
13. Zubin, M.A.; Ostapenko, N.A. Experimental investigation of some singularities of the supersonic flow around V-shaped wings.
Fluid Dyn. 1975, 10, 647–652.
14. Ostapenko, N.A. Aerodynamic characteristics of V-shaped wings with shock waves detached from leading edges at hypersonic
speeds. Fluid Dyn. 1993, 28, 545–552.
15. Maksimov, F.A.; Ostapenko, N.A. V-shaped wings with an opening angle exceeding π at super- and hypersonic flow velocities.
Dokl. Phys. 2016, 61, 412–417.
16. Mazhul, I.I.; Rakhimov, R.D. Numerical investigation of off-design regimes of flow past power-law waveriders based on the
flows behind plane shocks. Fluid Dyn. 2003, 38, 806–814.
17. Mazhul, I.I.; Rakhimov, R.D. Hypersonic power-law shaped waveriders. J. Aircr. 2004, 41, 839–845.
18. Mazhul, I.I. Off-design regimes of flow past waveriders based on isentropic compression flows. Fluid Dyn. 2003, 45, 271–280.
19. Jones, J.G.; Moore, K.C.; Pike, J.; Roe, P.L. A method for designing lifting configurations for high supersonic speeds, using
axisymmetric flow fields. Ing.-Arch. 1968, 37, 56–72.
20. Mangin, B.; Benay, R.; Chanetz, B.; Chpoun, A. Optimization of viscous waveriders derived from axisymmetric power-law blunt
body flows. J. Spacecr. Rocket. 2006, 43, 990–998.
21. Corda, S. Viscous Optimized Hypersonic Waveriders Designed from Flows over Cones and Minimum Drag Bodies. Ph.D.
Thesis, University of Maryland, College Park, MD, USA, 1988.
22. Ding, F.; Liu, J.; Shen, C.-B.; Huang, W. Novel approach for design of a waverider vehicle generated from axisymmetric
supersonic flows past a pointed von Karman ogive. Aerosp. Sci. Technol. 2015, 42, 297–308.
23. Rasmussen, M.L. Lifting-body Configurations Derived from Supersonic Flows Past Inclined Circular and Elliptic Cones. In
Proceedings of the 5th Atmospheric Flight Mechanics Conference for Future Space Systems, AIAA Paper, Boulder, CO, USA,
6–8 August 1979; p. 1665.
24. Rasmussen, M.L. Waverider configurations derived from inclined circular and elliptic cones. J. Spacecr. Rocket. 1980, 17, 537–545.
25. Rasmussen, M.L.; Jischke, M.C.; Daniel, D.C. Experimental forces and moments on cone-derived waveriders for m = 3 to 5. J.
Spacecr. Rocket. 1982, 19, 592–598.
Computation 2024, 12, 140 24 of 24
26. Rasmussen, M.L.; Jischke, M.C.; Daniel, D.C. Experimental surface pressures on cone-derived waveriders for M = 3–5. J. Spacecr.
Rocket. 1983, 20, 539–545.
27. Lin, S.C.; Rasmussen, M.L. Cone-derived Waveriders with Combined Transverse and Longitudinal Curvature. In Proceedings
of the 26th Aerospace Sciences Meeting, AIAA Paper 88–03714, Reno, NV, USA, 11–14 January 1988.
28. Liu, C.Z.; Bai, P.; Chen, B.Y.; Ji, C.Q. Rapid Design and Optimization of Waverider from 3D Flow. In Proceedings of the 16th
AIAA Aviation Technology, Integration, and Operations Conference, AIAA Paper 2016–3288, Washington, DC, USA, 13–17 June
2016.
29. Mclaughlin, T.A. Viscous Optimized Hypersonic Waveriders for Chemical Equilibrium Flow. Master’s Thesis, University of
Maryland, College Park, MD, USA, 1990.
30. Anderson, J.D.; Chang, J.; Mclaughlin, T.A. Hypersonic Waveriders: Effects of Chemically Reacting Flow and Viscous
Interaction. In Proceedings of the 30th Aerospace Sciences Meeting and Exhibit, AIAA Paper 92–0302, Reno, NV, USA, 6–9
January 1992.
31. Li, M.Y.; Zhou, H. Research on the design method of waverider aerodyanmic configuration based on commercial CFD software.
Chin. Q. Mech. 2014, 35, 293–299.
32. Taylor, G.I.; Macoll, J.W. The air pressure on a cone moving at high speeds I. Proc. R. Soc. A 1933, 139, 278–297.
33. Santos, W.F. Bluntness impact on lift-to-drag ratio of hypersonic wedge flow. J. Spacecr. Rocket. 2009, 46, 329–339.
34. Chen, X.Q.; Hou, Z.X.; Liu, J.X.; Gao, X.Z. Bluntness impact on performance of waverider. Comput. Fluids 2011, 48, 30–43.
35. Tincher, D.J.; Burnett, D.W. Hypersonic waverider test vehicle-a logical next step. J. Spacecr. Rocket. 1994, 31, 392–399.
36. Bird, G.A. Molecular Gas Dynamics and the Direct Simulation of Gas Flows, Clarendon Press: Oxford, UK, 1994.
37. Wagner, W. A convergence proof for Bird’s direct simulation Monte Carlo method for the Boltzmann equation. J. Stat. Phys.
1992, 66, 1011–1044.
38. Gallis, M.A.; Torczynski, J.R.; Plimpton, S.J.; Rader, D.J.; Koehler, T. Direct Simulation Monte Carlo: The quest for speed. In
Proceedings of the 29th Rarefied Gas Dynamics (RGD) Symposium, Xi’an, China, 13–18 July 2014.
39. Koura, K.; Matsumoto, H. Variable soft sphere molecular model for air species. Phys. Fluids A Fluid. Dyn. 1992, 4, 1083–1085.
40. Borgnakke, C.; Larsen, P.S. Statistical Collision Model for MonteCarlo Simulation of Polyatomic Gas Mixture. J. Comput. Phys.
1975, 18, 405–420.
41. Gallis, M.A.; Torczynski, J.R. Effect of Collision-Partner Selection Schemes on the Accuracy and Efficiency of the Direct
Simulation Monte Carlo Method. Int. J. Numer. Methods Fluids 2011, 67, 1057–1072. https://doi.org/10.1002/fld.2409.
42. Zhan, J.M.; Li, Y.T.; Wai, W.H.O.; Hu, W.Q. Comparison between the Q criterion and Rortex in the application of an in-stream
structure. Phys. Fluids 2019, 31, 121701.
43. Jeong, J.; Hussain, F. On the identification of a vortex. J. Fluid Mech. 1995, 285, 69–94.
44. Huang, W.; Ma, I.; Wang, Z.; Pourkashanian, D.B.; Luo, S.; Lei, J.A. parametric study on the aerodynamic characteristics of a
hypersonic waverider vehicle. Acta Aeronaut. 2011, 69, 135–140.
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.