Journal Pre-proof Design and aerodynamic performance analysis of a variable-sweep-wing morphing waverider
Pei Dai, Binbin Yan, Wei Huang, Yifei Zhen, Mingang Wang, Shuangxi Liu
PII:
S1270-9638(19)31844-9
DOI:
https://doi.org/10.1016/j.ast.2020.105703
Reference:
AESCTE 105703
To appear in:
Aerospace Science and Technology
Received date:
8 July 2019
Revised date:
10 January 2020
Accepted date:
11 January 2020
Please cite this article as: P. Dai et al., Design and aerodynamic performance analysis of a variable-sweep-wing morphing waverider, Aerosp. Sci. Technol. (2020), 0, 105703, doi: https://doi.org/10.1016/j.ast.2020.105703.
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier.
1
Design and Aerodynamic Performance Analysis of a
2
Variable-sweep-wing Morphing Waverider
3
1. Pei Dai: School of Astronautics, Northwestern Polytechnical University, Xi’an 710072, China;
4 5 6 7 8 9 10 11 12 13
E-mail:
[email protected] 2. Binbin Yan * (corresponding author): School of Astronautics, Northwestern Polytechnical University, Xi’an 710072, China; E-mail:
[email protected] 3. Wei Huang: Science and Technology on Scramjet Laboratory, National University of Defense Technology, Changsha 410073, China; E-mail:
[email protected] 4. Yifei Zhen: a. The 9th Designing, CASIC, Wuhan, Hubei 430040, China; b. School of Aeronautics, Northwestern Polytechnical University, Xi’an 710072, China; E-mail:
[email protected] 5. Mingang Wang: School of Astronautics, Northwestern Polytechnical University, Xi’an 710072, China; E-mail:
[email protected] 6. Shuangxi Liu: School of Astronautics, Northwestern Polytechnical University, Xi’an 710072,
14
China; E-mail:
[email protected]
15
Abstract: The wide speed range and large flight envelope of the hypersonic vehicle require that its
16
aerodynamic configuration still has good aerodynamic performance at low Mach number. Therefore,
17
the variable Mach number waverider is proposed to achieve good flight performance in a wide speed
18
range. In this paper, based on the delta-winged variable Mach number waverider, a
19
variable-sweep-wing morphing waverider is proposed and studied, including four specific
20
sweep-wing configurations, namely loiter, standard, dash and wing-retracted configurations. In the
21
current study, the aerodynamic performances of this variable-sweep-wing morphing waverider with
22
four configurations are investigated under subsonic/supersonic/hypersonic flight conditions. At the 1
23
same time, the numerical approaches employed are validated against the available experimental data
24
in the open literature. The obtained results show that compared with the wing-retracted configuration,
25
the best flight performance of this variable-sweep-wing morphing waverider can be achieved using
26
different configuration for different flight condition. However, within the hypersonic speed range, the
27
aerodynamic performance is improved through morphing but its advantages are not as large as that in
28
the subsonic speed. Besides, effects of wing downwash and shockwave are also analyzed. In
29
conclusion, the variable-sweep-wing morphing waverider improves both low-speed and high-speed
30
aerodynamic performances, and it expands the flight speed range.
31
Keywords: Morphing waverider; aerodynamic performance; variable-sweep-wing; wide-speed-range
32
vehicle
33 34
1. Introduction
35
Traditional aircrafts are designed for specific flight conditions, causing certain mission segments
36
to be compromised. One solution of this problem is to design aircrafts for a wide speed range with due
37
consideration of different flight scenarios. With the development of the aeronautical and astronautical
38
technology, the design of aircrafts aims to the wider velocity range and larger space range, especially
39
for the aerodynamic configuration design of the hypersonic vehicle. For example, a reusable launch
40
vehicle (RLV) operating at different flight regimes from subsonic to hypersonic speeds was proposed
41
in Ref. [1], and aerodynamic control surfaces’ effects were provided as well. Moreover, many novel
42
configurations have been proposed for hypersonic vehicles to achieve good flight performance in a
43
wide speed range. One solution of this problem is design of the variable Mach number waverider.
44
Three design methodologies of the variable Mach number waverider for a wide-speed range are 2
45
cone-derived method, osculating-cone-derived method and the tandem/parallel combination of
46
different waveriders. In Ref. [2], a variable Mach number waverider was proposed based on the
47
cone-derived method. With the variation of Mach number, the vehicle always had the waverider
48
characteristics partly, and its overall performance was robust in the wide-speed range. Ref. [3]
49
introduced a novel design approach of the cone-derived variable Mach number waverider based on
50
the streamline tracing technique. Compared with the traditional conical-derived waverider, the
51
streamlines using to form the vehicle’s lower surface were traced in flow-fields with different Mach
52
numbers. The aerodynamic configuration of the vehicle generated by this novel method was mainly
53
affected by the upper surface’s base curve, which was parameterized in their paper.
54
As for the osculating-cone-derived method, a variable Mach number waverider design method
55
was proposed in Ref. [4]. The osculating cone variable Mach number waverider owned higher
56
lift-to-drag ratio throughout the flight profile when compared with the osculating cone constant Mach
57
number waverider, and it had superior low-speed aerodynamic performance (M=4.0) while
58
maintaining nearly the same high-speed aerodynamic performance. Moreover, two kinds of design
59
methods for the constant swept waverider were discussed in Ref. [5], and both of them had more
60
flexible design curves. Compared with the general osculating cone waverider, the cuspidal waverider
61
had better high-speed aerodynamic performance under any flight condition in the high-speed range.
62
By means of combination, Wang et al. [6] designed a novel wide-speed-range waverider which
63
can level takeoff phase and hypersonic cruising phase by connecting a low-speed waverider and a
64
hypersonic waverider. Li et al. [7] also proposed design schemes of the tandem wide-speed-range
65
waverider by combining two waveriders which were designed for different Mach numbers (M=4 and
66
M=8) in tandem. And the appropriate configuration of the connection section was very important for 3
67
this hypersonic vehicle. Besides, the parallel waverider on a wide speed-range with two wings was
68
proposed in Ref. [8]. These wings enhanced the pitching moment performance and maneuverability of
69
this vehicle. However, the wings increased drag coefficient of the parallel vehicle remarkably, and
70
resulted in the decrease of the lift-to-drag ratio. In addition, in order to improve the lift-to-drag ratio in
71
the landing phase, Liu et al. [9] designed the wide-ranged multistage morphing waverider which was
72
generated in the same conical flow field, and it contained a free-stream surface and different
73
compression-stream surfaces.
74
The wide speed and large flight envelope of the hypersonic vehicle require that its aerodynamic
75
configuration still has good aerodynamic performance at low Mach number. However, the
76
above-mentioned papers designed waveriders in the hypersonic speed and they rarely included the
77
analysis of the low speed aerodynamic performance in the subsonic/transonic speed. As for other
78
researches, the low speed aerodynamic performance of waverider has been examined. In the Langley
79
Research Center, Pegg et al. [10] tested two separate waverider-derived vehicles, and these tests
80
provided measurements of moments and forces about all three axes, control effectiveness, flow field
81
characteristics and the effects of configuration changes. In Ref. [11], the planform generated by the
82
optimization of a dual fuel waverider configuration at a hypersonic cruise speed of Mach 10 was
83
examined for its low speed performance at Mach 0.25. Takama [12] proposed that outer wings were
84
attached to the ideal waverider with analyses of both subsonic and hypersonic aerodynamic
85
performances. The proposed practical waverider successfully improved the low-speed performance
86
without losing the hypersonic performance. In Ref. [13], the low speed aerodynamic performance
87
(from M=0.8 to M=2.5) of vortex lift waveriders with a wide-speed range was investigated, and the
88
low speed lift-to-drag ratio of the cuspidal waverider was higher than that of the general osculating 4
89
cone waverider.
90
Another solution for wide speed range vehicle is though morphing technology. Morphing
91
aircrafts can change different wing shapes or geometries to achieve the optimal flight performance
92
according to various mission scenarios, such as takeoff, loiter, reconnaissance, attack and landing. In
93
Ref. [14], the Naval Research Laboratory developed a morphing waveriders with a constant leading
94
edge and a top surface with a morphing lower stream surface to enable on-design performance across
95
a wide Mach number range from Mach 5 to Mach 10. A multiple joint variable-sweep morphing
96
aircraft was developed in Ref. [15], and it allowed independent choice of inboard and outboard for
97
each wing. Besides, Chakravarthy et al. proposed a time-varying characteristic equation of the
98
influence of different morphing trajectories in Ref. [16]. A longitudinal LPV model for the Z-wing
99
morphing UAV was proposed in Ref. [17]. An adaptive super-twisting sliding mode control of
100
variable sweep morphing aircraft was proposed in Ref. [18]. In order to make this new morphing
101
aircraft be able to perform rapid autonomous morphing and aerodynamic performance optimization
102
under different missions and flight conditions, Xu et al. developed deep neural networks and
103
reinforcement learning techniques as a control strategy in Ref. [19]. A six-degrees of freedom
104
nonlinear time-varying model and sliding mode flight controller were proposed in Ref. [20]to enhance
105
the lateral maneuverability for a tailless telescopic wing morphing aircraft by using additional
106
asymmetric wing telescoping. In Ref. [21], the controller was designed for hypersonic flight vehicle
107
(HFV) considering the actuator hysteresis and the angle of attack (AOA) constraint respectively. But,
108
researches on the morphing aircraft mainly focus on subsonic/transonic speed.
109
The main objective of this paper is to enhance the low speed aerodynamic performance of the
110
variable Mach number waverider aiming for a wide-speed range flight. In the subsonic/transonic 5
111
speed range, extended morphing wings can produce a large lift and lift-to-drag ratio. However, in the
112
hypersonic speed, wings lead to a larger increment of drag coefficient than lift coefficient, which leads
113
to a decrease of the lift-to-drag ratio. Therefore, a wing-retracted configuration is preferred in the
114
hypersonic speed. In conclusion, two variable sweep wings are added to the delta-winged waverider
115
which was designed in Ref. [5], namely the variable-sweep-wing morphing waverider.
116
The rest of this paper is organized as follows: First, the variable-sweep-wing morphing waverider
117
model and aerodynamic parameters are introduced in Section 2. Next, the numerical approach
118
employed in the current study is verified by two models and the aerodynamic performances of four
119
configurations are analyzed in Section 3. In addition, preliminary discussions on the differences
120
among four configurations are proposed in Section 4. Finally, the conclusion section ends this paper in
121
Section 5.
122 123 124
Fig. 1 Roadmap of this paper 2. Design of the variable-sweep-wing morphing waverider 6
125
The high-speed aerodynamic performance advantages of the delta-winged waverider have been
126
analyzed in detail and it improved the aerodynamic characteristics of wide Mach numbers ranging
127
from 4.0 to 8.0 in Ref. [5]. In this paper, the delta-winged waverider is chosen as the forebody to
128
make use of its good aerodynamic performance with increased lift. To enhance the volumetric
129
efficiency, a fuselage is attached to the forebody with the same cross section at the base plane of the
130
delta-winged waverider forebody. Meanwhile, two all-moving elevons and a vertical tail are control
131
surfaces. In Ref. [22], reusable launch vehicle was configured as a winged body vehicle. This vehicle
132
consists of a fuselage, two double delta wings, two vertical tails and two elevons mounted behind the
133
wing. But in this paper, a waverider forebody provides large lift compared with the fuselage and it
134
makes the center of pressure move forward. Therefore, a large pitching moment generated by control
135
surfaces is required to balance this vehicle as desired positive angle of attack and it can be provided
136
by all-moving elevons. And when the deflection of elevon reaches -20° and center of mass is chosen
137
as 0.33 axial length, the balanced angle of attack will be as small as 4° at some flight conditions (such
138
as M=6, standard configuration). Therefore, all-moving elevons are selected in this paper and a
139
hypersonic vehicle with control surfaces is designed. And the geometric model of the designed
140
hypersonic vehicle with morphing wings retracted is shown in Fig. 2.
(a) 3D view
(b) Top view
(c) Main view
7
(d) Side View
(e) Top view with size marking (mm) 141
Fig. 2 Geometric model of the variable-sweep-wing morphing waverider (wing-retracted configuration).
142 143
Based on that, two variable-sweep-wings are attached to the fuselage to provide low speed
144
aerodynamic performance advantages, especially during the subsonic flight. The loiter configuration
145
with a 20°-sweep-angle supercritical airfoil is chosen as the baseline configuration as shown in Fig. 3.
146
The wing leading edge radius of this airfoil is 0.873mm. Two variable-sweep-wings rotate about
147
rotating shafts continuously near the wing root to suit for different operating regimes. During the
148
hypersonic flight, two morphing sweep wings retract in the body to achieve good hypersonic
149
aerodynamic performance with a high lift-to-drag ratio thanks to the delta-winged waverider
150
forebody.
151 152
Fig. 3 Geometric model of the variable-sweep-wing with the SC(2)-0706 supercritical airfoil.
153
In this paper, two variable-sweep-wings can rotate about rotating shaft continuously. Therefore, 8
154
three wing configurations are chosen as specific examples to investigate the optimal aerodynamic
155
performance which suits for different operating regimes. Fig. 4 to Fig. 6 show geometric models of
156
the variable-sweep-wing morphing waverider of three different configurations (sweep angle Λ =20°,
157
40°, 60°). The configuration with morphing wings retracted (shown in Fig. 2) is denoted as
158
wing-retracted configuration. The configurations with sweep angles of 20° (shown in Fig. 4 ), 40°
159
(shown in Fig. 5) and 60° (shown in Fig. 6) are denoted as loiter, standard and dash configurations,
160
respectively.
(a) 3D view 161
(b) Top view
Fig. 4 Geometric model of the loiter configuration.
(a) 3D view 162
(b) Top view
Fig. 5 Geometric model of the standard configuration.
(a) 3D view
(b) Top view
163
Fig. 6 Geometric model of the dash configuration
164
Wing parameters of different configurations are shown in Table 1. However, as the sweep angle
165
increases, the wing span, wing area and aspect ratio all become lower. 9
166
Table 1. Wing parameters of different configurations considered in the current study.
Sweep angle, Λ (°)
20
40
60
Configuration
Loiter
Standard
Dash
Root Chord (m)
0.168
0.228
0.351
Tip Chord (m)
0.168
0.183
0.180
Span (m)
2.0
1.792
1.512
Wing area ( m 2 )
0.240
0.234
0.221
Aspect ratio
16.667
13.723
10.344
167
For comparative purpose, the top view area of the wing-retracted configuration in Fig. 2 (b) is
168
chosen as the reference area, Sref (Sref=1.312 m 2 ), and the axial length is chosen as the reference length,
169
Lref (Lref =1.785m). In addition, in order to compare the aerodynamic performances of the
170
variable-sweep-wing morphing waverider with different configurations, the above reference length
171
and area are also used to calculate the aerodynamic coefficients of the loiter, standard and dash
172
configurations.
173
3. Analysis of aerodynamic performances
174
For evaluation of the aerodynamic performances of the variable-sweep-wing morphing
175
waverider designed in this paper, the software Metcomp CFD++ [23] is used to simulate 3D viscous
176
flow fields of the variable-sweep-wing morphing waverider under different flight conditions. For this
177
study, the three-dimensional Reynolds-Averaged Navier Stokes (RANS) equations and the SST k-Ȧ
178
turbulence model [2][2][5][24][25] are adopted in this numerical simulation. The operational fluid is
179
air, and it is treated as an ideal gas with no reactions modeled.
180
3.1. Code validation 10
181
In the 2nd CFD Drag Prediction workshop, the lift, drag, and pitching moments are calculated
182
for the DLR-F6 configuration at subsonic flow conditions by solving the Reynolds-averaged
183
Navier–Stokes equations on structured as well as on unstructured, hybrid grids[26]. In addition, to
184
verify the effectiveness of the numerical approach employed in this paper in the hypersonic speed
185
(M=8.2), the hemi-spherically blunted cone-cylinder body model are numerically simulated, and the
186
predicted results are compared with the experimental data.
187
a) DLR-F6 at M=0.75
188
The DLR-F6 model was originally built in the 1980’s and subsequently tested in several
189
investigations at the French ONERA S2MA wind tunnel[27] and main dimensions of the DLR-F6 WB
190
configuration are depicted in Fig. 7. The DLR-F6 model was sting mounted in the 1.77 m × 1.75m
191
transonic test section using a z-sting mount attached to the aft portion of the upper fuselage as shown
192
in Fig. 8 .During experiments, pressure distributions were measured on the right wing by 288 pressure
193
orifices located in 8 span-wise wing sections and total forces were measured by a balance.
194 195
Fig. 7 Main dimensions of the DLR-F6 WB configuration.
11
196 197
Fig. 8 DLR-F6 WB configuration in the ONERA S2MA wind tunnel
198
The numerical results were calculated by Uriel Goldberg [28] using CFD++ codes. In Case 1,
199
single point grid sensitivity study of CFD++ was carried out under the flight condition (Mach number
200
is 0.75, angle of attack is 0.144e, lift coefficient is 0.5) as shown in Table 2. This case was run at a
201
specified lift coefficient (CL=0.5). The numbers of cells of coarse mesh, medium mesh and fine mesh
202
were 5.5 million, 7.4 million and 9.6 million respectively. The maximum increments of drag
203
coefficient and pitching moment coefficient compared with fine mesh were 1.07% and 6.45%
204
respectively.
205
Table 2. Single point grid sensitivity study of CFD++.
Coarse mesh
Medium mesh
Fine mesh
(5.5 million)
(7.4 million)
(9.6 million)
Drag coefficient
0.02915
0.02933
0.02902
Pitching moment coefficient
-0.1388
-0.1385
-0.1394
0.44%
1.07%
------
4.30%
6.45%
------
Grid size
Darg coefficient increment compared with Fine mesh
Pitching moment coefficient increment compared with Fine mesh
12
206
And they were also compared with the experimental data to verify the effectiveness of drag polar
207
study in Case 2 (M=0.75, ¢=-3,-2,-1.5,-1,0,1,1.5,2.0), as shown in Fig. 9. The boundary conditions
208
for the freestream included the Mach number M=0.75, the static temperature T0 = 305 K, and the
209
Reynolds number Re = 3 × 106, in accordance to the experimental setup. And the number of cells is
210
7.4 million. The values of the lift, drag and pitching moment coefficients obtained by the numerical
211
approach are a slightly larger than the experimental data, but their trends with the increase of the angle
212
of attack are the same, and differences between numerical approach and experimental data are shown
213
in Table 3. As shown in Fig. 9 (d), the numerical approach shows very good agreement with the
214
experimental data in the drag polar. According to Ref.[29], there are several reasons why the CFD
215
results were not expected to match the experimental data exactly. First, the CFD runs were all
216
specified to be fully turbulent. Also, the CFD runs were all computed in free air, and the sting mount
217
was not modeled. The effects of these differences are difficult to quantify without specific study to
218
identify them.
(a) Lift coefficient
(b) Drag coefficient
13
(c) Pitching moment coefficient
(d) Drag polar
219
Fig. 9 Comparisons of lift, drag and pitching moment coefficients versus angle of attack.
220
Table 3. Increments of lift, drag and pitching moment coefficients compared with experimental data.
221
Angle of attack (°)
-3
-2
-1.5
-1
0
1
1.5
Lift coefficient increment (%)
42.76
19.84
16.09
13.23
10.36
6.86
4.60
Drag coefficient increment (%)
0.60
2.58
3.62
3.85
5.02
5.51
6.46
Pitching moment coefficient increment (%) 10.89
11.34
11.39
12.16
12.67
10.75
7.60
b) hemi-spherically blunted cone-cylinder body model at M=8.2
222
Singh [30] conducted extensive wind tunnel experiments in the hypersonic speed range on a
223
scale model of a hemi-spherically blunted cone-cylinder body model (shown in Fig. 10). According to
224
Ref. [30], the Mach number is 8.2 and the angle of attack varied from -3° up to 10°. In this paper, the
225
hemi-spherically blunted cone-cylinder body model is numerically simulated, and the computed
226
results are compared with the experimental data to verify the effectiveness of the numerical
227
approaches employed in this paper, and they are shown in Fig. 11. The number of cells is 2 million
228
and the flight condition is M=8.2, Re=9.35h104/cm, the static pressure P0 = 951.5Pa, and the static
229
temperature T0 = 89.3 K. In addition, the length and base area (i.e. 17.7 cm and 3.14 cm2) of the
230
hemi-spherically blunted cone-cylinder body model are used as the reference length and area, 14
231
respectively. In Fig. 11, the red dotted line denotes the experimental data derived from Ref. [30], and
232
the black line denotes the predicted data calculated by CFD++. Differences between numerical
233
approach and experimental data are shown in Table 4.
(a) Grids on the boundary layer
(b) Grids on the surface
234
Fig. 10 Schematic diagram of the grid employed in the hemi-spherically blunted cone-cylinder body
235
model.
236 237
Fig. 11 Comparisons of lift and drag coefficients.
238
Table 4. Increments of lift, drag and pitching moment coefficients compared with experimental data.
Angle of attack (°)
-3
0
3
5
7
10
Lift coefficient increment (%)
4.28
----
4.30
20.01
0.34
20.33
Drag coefficient increment (%)
10.28
25.29
14.78
12.29
12.39
1.74
239
Besides, several hypersonic flow cases of engineering and scientific interest have been computed
240
using various turbulence models available in the CFD++ flow solver in Ref. [31]. In all cases very 15
241
good agreement between predictions and experimental data of pressure, heat transfer and skin friction
242
were obtained. From the above discussions, it can be shown that the numerical approaches applied in
243
this paper are with confidence to study the aerodynamic performances of the variable-sweep-wing
244
morphing waverider, and the predicted results are believable in the following parts.
245
3.2. Aerodynamic performance
246
In Ref. [5], the designed Mach numbers of the delta-winged waverider are 3.86/4.4/8.0. In order
247
to investigate the aerodynamic performance of the variable-sweep-wing morphing waverider, different
248
flight conditions need to be selected for the numerical simulation. In Ref. [3]and [5], the cruising
249
altitude was chosen as 25 km. In Ref. [14], the flight altitude was chosen as 30 km. In Ref. [6], the
250
cruising altitude was chosen as 15 km. At the last flight test of X-51A in 2013, flight altitude was
251
19km. Therefore, the selected supersonic/hypersonic Mach numbers are 3.0/6.0/8.0, and its cruising
252
flight condition is set as 20km with Mach 8. Besides, aerodynamic performances can be easily
253
simulated and analyzed in the different flight altitude by changing free stream parameters in future
254
studies. At this cruising flight condition, the static temperature of the atmosphere is 216K, the static
255
pressure is 5475Pa and the dynamic pressure is 245275Pa. To provide an optimal or steady condition
256
for the propulsion system, the RBCC powered hypersonic vehicle generally climbs with the constant
257
dynamic pressure at the ascent stage [4]. Therefore, the altitude of flight Mach 3 is chosen as 7.25km,
258
while the altitude of flight Mach 6 is chosen as 16.35km according to the constant dynamic pressure
259
trajectory.
260
In addition, the low speed aerodynamic performances are also investigated for the
261
variable-sweep-wing morphing waverider, and the subsonic/supersonic flight conditions are Mach 0.8,
262
H=10km and Mach 1.5, H=10km. All these specific selected flight conditions can refer to Table 5. 16
Table 5. Flight conditions.
263
H(km)
Mach Number
Static pressure (kPa)
Dynamic pressure (kPa)
Velocity (m/s)
10
0.8
26.43
11.84
239.62
10
1.5
26.43
41.63
449.29
7.25
3
39.40
245.28
933.15
16.35
6
98.77
245.28
1770.41
20
8
55.29
245.28
2360.55
264
The lift coefficients, drag coefficients and lift-to-drag ratios of four configurations (loiter,
265
standard, dash and wing-retracted configurations ) with the increase of angle of attack (i.e. í2°, 0°, 2°,
266
5° and 10°) under five flight conditions, are all obtained and depicted in Fig. 12, Fig. 13 and Fig. 14
267
respectively. In Fig. 12 to Fig. 14, the loiter configuration ( Λ =20e) is denoted by marker ƺ and
268
dotted line, the standard configuration ( Λ =40e) is denoted by marker ƶ and dashed line, the dash
269
configuration ( Λ =60e) is denoted by marker Ƹ and dash-dotted line, and the wing-retracted
270
configuration ( Λ =90e) is denoted by marker ͪ and solid line.
271 272
In the following paragraphs, α denotes the angle of attack, CL denotes the lift coefficient,
CD denotes the drag coefficient, and L / D denotes the lift-to-drag ratio.
17
CL
CL
273
(a) M=0.8
(b) M=1.5
0.35 0.3 0.25 0.2
0.25 CL( =20deg)
CL( =20deg)
CL( =40deg)
0.2
CL( =60deg) CL( =90deg)
CL( =40deg) CL( =60deg) CL( =90deg)
0.15
0.15 0.1
0.1 0.05
0.05
0 -0.05 -2
0
2
4
6
8
0 -2
10
(°)
0
2
4
6
8
10
(°)
(d) M=6
CL
(c) M=3
(e) M=8 274 275
Fig. 12 Lift coefficients of the variable-sweep-wing morphing waverider in different configurations.
276 18
CD
CD
(a) M=0.8
(b) M=1.5
0.12
0.08 CD( =20deg)
0.1
0.07
CD( =40deg) CD( =60deg)
0.06
CD( =90deg)
0.08
CD( =20deg) CD( =40deg) CD( =60deg) CD( =90deg)
0.05 0.04
0.06
0.03 0.04 0.02 0.02 -2
0
2
4
6
8
0.01 -2
10
(°)
0
2
4
6
8
10
(°)
(d) M=6
CD
(c) M=3
(e) M=8 277 278
Fig. 13 Drag coefficients of the variable-sweep-wing morphing waverider in different configurations.
279
19
L/D
L/D
(a) M=0.8
(b) M=1.5
3.5
3.5
3
3
2.5 2.5
2 1.5
2
1
1.5 L/D( L/D( L/D( L/D(
0.5 0 -0.5 -2
0
2
4
6
=20deg) =40deg) =60deg) =90deg)
8
L/D( L/D( L/D( L/D(
1 0.5 -2
10
(°)
0
2
4
6
=20deg) =40deg) =60deg) =90deg)
8
10
(°)
(d) M=6
L/D
(c) M=3
(e) M=8 280 281
Fig. 14 Lift-to-drag ratios of the variable-sweep-wing morphing waverider in different configurations.
282
As can be seen from Fig. 12 to Fig. 14, under the same flight condition, lift coefficients and drag
283
coefficients of the four configurations increase with the increase of the angle of attack, and their 20
284
lift-to-drag ratios increase first and then decrease with the increase of the angle of attack. Moreover,
285
under the same angle of attack and the same configuration, the lift, drag coefficients both decrease
286
with the increase of the flight Mach number. The following section provides detailed discussions on
287
drag coefficient, lift coefficient and lift-to-drag ratio.
288
a) Drag coefficient
289
Under different flight conditions, relative changes in drag coefficients with the same angle of
290
attack in different sweep configurations are quite the same; and the drag coefficients become lower as
291
the sweep angle increases. Therefore, drag coefficients at the cruising flight condition (Mach=8.0,
292
H=20km) are chosen as an example to compare the aerodynamic characteristics of the four
293
configurations of the variable-sweep-wing morphing waverider, and they are shown in Table 6.
294
Table 6.
Comparison of drag coefficients of four configurations of the variable-sweep-wing morphing waverider at Mach=8, H=20km
295
Angle of attack (°)
-2
0
2
5
10
Loiter configuration ( Λ =20°)
0.019754
0.022438
0.027548
0.039291
0.070960
Standard configuration ( Λ =40°)
0.018025
0.020702
0.025770
0.037483
0.069082
Dash configuration ( Λ =60°)
0.016476
0.019175
0.024200
0.035976
0.067606
Wing-retracted configuration ( Λ =90°)
0.015335
0.018249
0.022986
0.034679
0.061192
296 297
b) Lift coefficient
298
To further analyze the existing differences, lift coefficients of the variable-sweep-wing morphing
299
waverider in four configurations at flight M=0.8, M=1.5, M=3, M=6 and M=8 are depicted in Table 7,
300
Table 8, Table 9, Table 10 and Table 11, respectively. As can be seen from Table 7 and Table 8, under 21
301
subsonic flight condition (M=0.8) and low speed supersonic flight condition (M=1.5), loiter
302
configuration has the largest lift coefficient compared with other configurations at the same positive
303
angle of attack.
304
However, as can be seen from Table 9 to Table 11, in the high-speed range (M=3, M=6, M=8),
305
the lift coefficients of loiter, standard and dash configurations are nearly the same, but they are all
306
larger than the wing-retracted configuration at positive angles of attack. As can be seen from Table 9,
307
under the flight condition Mach 3, the standard configuration has the largest lift coefficient compared
308
with other configurations at positive angles of attack. As can be seen from Table 10, under the flight
309
condition Mach 6, the dash configuration has the largest lift coefficient compared with other
310
configurations at angles of attack from 0° to 5°. However, at the high angle of attack, the lift
311
coefficient of the standard configuration is the largest. As can be seen from Table 11, under the flight
312
condition Mach 8, the dash configuration has the largest lift coefficient compared with other
313
configurations at positive angles of attack. This shows that different configurations of the
314
variable-sweep-wing morphing waverider suit for different flight conditions with the corresponding
315
largest lift coefficient.
316
Table 7.
Comparison of lift coefficients of four configurations of the variable-sweep-wing morphing waverider at M=0.8, H=10km.
317
Angle of attack (°)
-2
0
2
5
10
Loiter configuration ( Λ =20°)
0.016315
0.129002
0.241981
0.386298
0.577483
Standard configuration ( Λ =40°)
-0.008369
0.087742
0.155285
0.312580
0.537427
Dash configuration ( Λ =60°)
-0.016415
0.058482
0.133100
0.250815
0.483442
Wing-retracted configuration ( Λ =90°)
-0.000625
0.049749
0.102599
0.190366
0.365065
22
318
Table 8.
Comparison of lift coefficients of four configurations of the variable-sweep-wing morphing waverider at M=1.5, H=10km
319
320
Angle of attack (°)
-2
0
2
5
10
Loiter configuration ( Λ =20°)
-0.047298
0.044453
0.137971
0.277054
0.503541
Standard configuration ( Λ =40°)
-0.055504
0.036292
0.129868
0.266058
0.482392
Dash configuration ( Λ =60°)
-0.061294
0.019952
0.102222
0.226396
0.442081
Wing-retracted configuration ( Λ =90°)
-0.038259
0.013649
0.069830
0.160637
0.323762
Table 9.
Comparison of lift coefficients of four configurations of the variable-sweep-wing morphing waverider at M=3, H=7.3km
321
322
Angle of attack (°)
-2
0
2
5
10
Loiter configuration ( Λ =20°)
-5.74e-05
0.049551
0.099514
0.175698
0.304030
Standard configuration ( Λ =40°)
-0.000573
0.050192
0.101112
0.178387
0.306489
Dash configuration ( Λ =60°)
-0.001505
0.049701
0.100436
0.175587
0.296456
Wing-retracted configuration ( Λ =90°)
0.005586
0.043188
0.081403
0.141280
0.241506
Table 10.
Comparison of lift coefficients of four configurations of the variable-sweep-wing morphing waverider at M=6, H=16.3km
323
Angle of attack (°)
-2
0
2
5
10
Loiter configuration ( Λ =20°)
0.012974
0.043439
0.073936
0.120931
0.203206
Standard configuration ( Λ =40°)
0.012745
0.043400
0.073990
0.121197
0.204280
Dash configuration ( Λ =60°)
0.012693
0.043893
0.074700
0.122103
0.203642
Wing-retracted configuration ( Λ =90°)
0.016873
0.042163
0.067351
0.106620
0.173450
23
324
Table 11.
Comparison of lift coefficients of four configurations of the variable-sweep-wing morphing waverider at M=8, H=20km
325
Angle of attack (°)
-2
0
2
5
10
Loiter configuration ( Λ =20°)
0.016603
0.042406
0.068370
0.108734
0.181994
Standard configuration ( Λ =40°)
0.016543
0.042323
0.068231
0.108619
0.182294
Dash configuration ( Λ =60°)
0.016480
0.042599
0.068671
0.109377
0.183858
Wing-retracted configuration ( Λ =90°)
0.020275
0.042060
0.063870
0.096006
0.158142
326
As shown in Fig. 15, in order to quantitatively analyze the lift coefficient differences among four
327
configurations of the variable-sweep-wing morphing waverider, the increment or decrement
328
percentages of lift coefficients of the three wing-extended configurations to that of the wing-retracted
329
configuration are calculated for each flight condition. Under the flight condition M=1.5, α =0D , the
330
increment percentage of the loiter configuration is the largest with the value of 225%. Under the flight
331
condition M=0.8, the increment percentage of the loiter configuration exceed 50% of all positive
332
angles of attack. Among all flight conditions, it is in the supersonic speed range (M=1.5) that the
333
increment percentage of the lift coefficient of the standard and dash configurations are the highest,
334
and with further increase of the flight Mach number, the percentage increase of the lift coefficient is
335
gradually reduced. In other words, compared with the wing-retracted configuration, the low speed
336
performances of the three configurations are improved.
24
(a) Loiter configuration
(b) Standard Configuration
(c) Dash configuration 337
Fig. 15 Increment percentage of lift coefficient of the loiter/standard/dash configuration relative to that of the wing-retracted configuration
338 339
c) Lift-to-drag ratio
340
As can be seen from Fig. 14(a), in the subsonic speed range (M=0.8), the lift-to-drag ratio of
341
loiter configuration (L/D=8.43, Į=2°) is always higher than other configurations, but they are close at
342
high angles of attack. The main reason is that at small angles of attack, the drag coefficients of the
343
four configurations are relatively close, but the lift coefficient of loiter configuration is much larger.
344
However, with the increase of the angle of attack, the drag coefficient of loiter configuration highly
345
surpasses that of other configurations, and this makes the lift-to-drag ratios of the four configurations
346
close at high angles of attack. 25
347
In contrast, under the low speed supersonic flight condition (M=1.5), the standard configuration
348
has the largest lift-to-drag ratio of 3.29 when Į=0°, and this is shown clearly in Fig. 14 (b). As can be
349
seen from Fig. 14 (c)-(e), under the high speed supersonic/hypersonic flight condition (M=3, 6, 8),
350
among loiter, standard and dash configurations, the dash configuration has the largest lift-to-drag ratio
351
when Į=0°. The main reason is that among three wing extended configurations, although the lift
352
coefficients of the four configurations are different from each other, but the difference among drag
353
coefficients is much larger than the increment of the lift coefficient. Therefore, a smaller drag
354
coefficient leads to a larger lift-to-drag ratio. It is quite interesting to notice that under some specific
355
flight conditions (M=6, α =0 and M=8, α =0), the wing-retracted configuration has the largest
356
lift-to-drag ratio compared with other three wing-extended configurations. Moreover, it can be found
357
that at the subsonic speed, the increase of the lift-to-drag ratio of loiter configuration relative to the
358
wing-retracted configuration is significantly higher than that at the supersonic/hypersonic speed. This
359
shows that different configurations of the variable-sweep-wing morphing waverider suit for different
360
flight conditions with their different advantages on aerodynamic performance.
361
As shown in Fig. 16, in order to quantitatively analyze the lift-to-drag ratio differences among
362
four configurations of the variable-sweep-wing morphing waverider, the percentages increase or
363
decrease of the lift-to-drag ratio of the three wing-extended configurations versus the wing-retracted
364
configuration are calculated for each flight condition.
26
(a) M=0.8
(b) M=1.5
(c) M=3
(d) M=6
(e) M=8 365 366
Fig. 16 Increment percentage of the lift-to-drag ratio of the loiter/standard/dash configurations relative to that of the wing-retracted configuration
367
As shown in Fig. 16 (a) and (b), at the same subsonic/low-speed supersonic speed, the
368
percentage increases of the lift-to-drag ratio of the three wing-extended configurations versus the 27
369
wing-retracted configuration decrease with the increase of the angle of attack, and this means that the
370
wing-extended configurations’ aerodynamic advantages gradually diminish with the increase of angle
371
of attack.
372
As shown in Fig. 16 (c), at the flight Mach number 3, the percentage increases of the lift-to-drag
373
ratio of the three wing-extended configurations shows a trend of rising first and then falling with the
374
increase of the angle of attack, which means that compared with the three wing-extended
375
configurations, the aerodynamic performance of the wing-retracted configuration is not significantly
376
reduced at the small angle of attack and high angle of attack. As shown in Fig. 16 (d) and (e), at the
377
flight Mach number 6 and 8, the percentage increases of the lift-to-drag ratio of the three
378
wing-extended configurations increase with the increase of the angle of attack. Compared between
379
Fig. 16 (c)-(e), at the same angle of attack, the percentage increase of the lift-to-drag ratio of the dash
380
configuration is the largest, and it decreases with the increase of the flight Mach number. This
381
indicates that the aerodynamic performance advantage of the dash waverider is more pronounced at
382
low hypersonic speed.
383
As shown in Fig. 16 (c)-(e), the black solid line denotes that the percentage increase of the
384
lift-to-drag ratio equals zero and it denotes the wing-retracted configuration. Therefore, it is
385
interesting to notice that under some specific flight conditions, the percentage increase of the
386
lift-to-drag ratio is lower than zero, which means that the aerodynamic performance of the
387
wing-retracted configuration is higher than that of some configurations (especially standard and loiter
388
configurations), and it is more obvious in high speeds and small angles of attack.
389
In conclusion, the variable-sweep-wing morphing waverider proposed in this paper enhances
390
both low speed take-off performance and high-speed cruising performance, and it expands the flight 28
391
speed range. For different flight conditions and different mission scenarios, the best flight
392
performance of this variable-sweep-wing morphing waverider can be achieved using different
393
configurations.
394
d) Pitching moment coefficient
395
The focus of this paper is the investigation of aerodynamic performance of the morphing
396
waverider, and the center of gravity of this vehicle has not been designed. Therefore, the nose at
397
symmetric plane of this vehicle is chosen as the reference center of pitching moment coefficient, as
398
shown in Fig. 17. And in this paper, a negative pitching moment denotes a nose-down pitching
399
moment.
400 401
Fig. 17 Reference point
402
The pitching moment coefficients of four configurations (loiter, standard, dash and
403
wing-retracted configurations ) with the increase of angle of attack (i.e. í2°, 0°, 2°, 5° and 10°) under
404
five flight conditions, are all obtained and depicted in Fig. 18. In Fig. 18, the loiter configuration
405
( Λ =20e) is denoted by marker ƺ and dotted line, the standard configuration ( Λ =40e) is denoted
406
by marker ƶ and dashed line, the dash configuration ( Λ =60e) is denoted by marker ƻ and
407
dash-dotted line, and the wing-retracted configuration ( Λ =90e) is denoted by marker ƺ and solid
408
line.
409
Under the subsonic flight condition (M=0.8) and the low speed supersonic flight condition 29
410
(M=1.5), the differences among different configurations are quite large. And with morphing wings
411
sweeping backwards, a smaller nose down pitching moment is induced. When aircraft changes from
412
loiter configuration to dash configuration, lift coefficient becomes smaller and center of pressure
413
shifts backwards resulting in a larger moment arm of lift. However, the decrease magnitude of lift is
414
bigger than the increase magnitude of moment arm of lift resulting in a smaller pitching moment
415
coefficient. Under the high speed supersonic (M=3) or hypersonic speeds (M=6, M=8), the
416
differences among three wing-extended configurations are very small. Under hypersonic speed, the
417
lift of morphing wings is very small compared with the lift generated by the waverider forebody and
418
the difference of lift among three wing-extended configurations can be neglected.
419
30
420
(a) M=0.8
(b) M=1.5
(c) M=3
(d) M=6
(e) M=8 421 422 423
Fig. 18 Pitching moment coefficients of four configurations versus angle of attack. 4. Preliminary discussion on differences among four configurations To explore the reasons for differences in the aerodynamic performances among four 31
424
configurations of the variable-sweep-wing morphing waverider proposed in this paper, the flow field
425
characteristics of two specific flight conditions (Mach 0.8, Į=0° and Mach 8, Į=0° ) of the four
426
configurations and effects of wing downwash and shock wave are analyzed in detail in this section.
427
4.1. M=0.8, Į=0°
428
According to the analysis in Section 3.2, under the flight condition Mach 0.8, the loiter
429
configuration has the largest lift coefficient, drag coefficient, and lift-to drag ratio compared with
430
other configurations. Besides, the drag coefficients of the four configurations are quite close at low
431
angle of attack, while at high angle of attack, the drag coefficients of the four configurations are quite
432
different.
433
Fig. 19 illustrates the pressure contours at the lower and upper surfaces of four configurations of
434
the variable-sweep-wing morphing waverider, and the pressure coefficient levels in eight contours are
435
the same for an accurate figure-to-figure comparison. For each subfigure, the upside and downside
436
figure denote different configuration for comparison. Moreover, apart from two variable-sweep-wings,
437
the pressure contours of upper and lower surfaces of all the four configurations are nearly the same,
438
which means that different configurations of the variable-sweep-wing lead to differences in pressure
439
distribution over wings and it is the main reason for different aerodynamic performance. Besides, the
440
lower surface pressure coefficients of four configurations are all higher than the corresponding upper
441
surface pressures. In addition, as can be seen from Fig. 19 (a) and (c), the upper surface of the wing of
442
the loiter configuration has a low pressure area drawn in the dark blue, and the lower surface of the
443
wing of it has a high pressure area drawn in the dark red. Therefore, the biggest difference between
444
pressure coefficients of upper and lower surfaces at the wing area of the loiter configuration makes it
445
have the largest lift coefficient. 32
a)
Upper surface, loiter vs standard
b)
configuration
c)
Upper surface, dash vs wing-retracted configuration
Lower surface, loiter vs standard
d)
configuration
Lower surface, dash vs wing-retracted configuration
Fig. 19 Comparisons of pressure coefficient contours of four configurations 446
Pressure coefficient contours on wing’s cross-section in a chord-wise direction (Y=0.46m) with
447
flowfield streamtraces of three wing-extended configurations are shown in Fig. 20. A high-pressure
448
area near the leading edge of the wing and pressure distribution of both wing and tail can be
449
investigated at this cross section. A small sweep angle of the loiter configuration leads to a
450
high-pressure area near the leading edge, and it results in a large drag coefficient.
33
a)
loiter configuration
b)
standard configuration
c)
dash configuration
Fig. 20 Pressure coefficient contours on a wing's cross-section in a chord-wise direction (Y=0.46m) with flowfield streamtraces of three wing-extended configurations (M=0.8, Angle of attack = 0°) 451
Fig. 21 presents pressure coefficient distributions of lower and upper surfaces of three
452
configurations on different wing’s cross sections (Y=0.46m and Y=0.66m). X/b denotes the ratio of x
453
coordinate with semi-span width. High pressure coefficients at the leading edge and following edge
454
can be observed, and the difference between pressure coefficients of upper and lower surfaces of the
455
loiter configuration is the biggest. The lift coefficient generates by the wing can be approximated by
456
calculating the area enclosed by Cp curves of upper and lower surfaces. Therefore, the loiter
457
configuration has the largest lift. As shown in Fig. 21 (a) and Fig. 20 (c), as to the dash configuration,
458
the pressure coefficient of the upper surface is even large than that of the lower surface when X/b is
459
smaller than 0.3, and it leads to a negative lift coefficient near this area. Therefore, the dash
460
configuration has the smallest lift. Meanwhile, by sweeping the wing, a streamline effectively sees a
461
thinner airfoil. As stated above, caused by a large sweep angle, the airfoil of dash configuration is
462
thinner than the loiter configuration at the same position, and the pressure at its lower surface is
463
smaller, so the dash configuration has the smallest pressure component in the drag direction.
34
a)
Y=0.46m
b)
Y=0.66m
Fig. 21 Pressure coefficients of lower and upper surfaces of three wing-extended configurations at different wing’s cross-section in a chord-wise direction (Y=0.46m and Y=0.66m) 464 465
4.2. M=8, Į=0°
466
Fig. 22 denotes pressure coefficient contours on the waverider forebody and surface of the
467
wing-retracted configuration. As shown in Fig. 22 (d)-(f), the good waverider performance of the
468
variable-sweep-wing waverider in Mach 8 can be observed. The shock wave on the forebody prevents
469
the leakage of the high-pressure gas on the lower surface of the waverider, and the waverider exhibits
470
good shock effect. As can be seen from Fig. 22 (c), the lower surface of the forebody has a
471
homogeneous pressure distribution while the upper surface is surrounded by the free flow condition.
35
472
a) 3D view with three planes
d) X=0.16m
b) Upper surface
c) Lower surface
e) X=0.31m
f) X=0.46m
Fig. 22 Pressure coefficient contours on the waverider forebody cross sections and surface of the wing-retracted configuration. 473
Fig. 23, Fig. 24 and Fig. 25 denote pressure coefficient contours on wing cross sections and
474
upper/lower surfaces of the loiter, standard and dash configurations, respectively. The high-pressure
475
area at the wing leading edge of the loiter configuration is more obvious than the other two
476
configurations, and it leads to a large drag coefficient. Compared with Fig. 20, the difference between
477
the pressure coefficients of upper and lower surfaces of the variable-sweep-wing is very small at the
478
hypersonic speed. In other words, in the hypersonic speed range, the aerodynamic performance
479
advantages of the wing-extended configurations are not as large as those in the subsonic speed. For
480
different flight condition and mission scenarios, the best flight performance of this
481
variable-sweep-wing morphing waverider can be achieved using different configuration. 36
a) 3D view with six planes
d) Y=-0.46m
b) Upper surface
c) Lower surface
e) Y=-0.55m
f) Y=-0.66m
Fig. 23 Pressure coefficient contours on wing cross sections and surfaces of the loiter configuration. 482
a) 3D view with six planes
d) Y=-0.46m
b) Upper surface
c) Lower surface
e) Y=-0.55m
f) Y=-0.66m
Fig. 24 Pressure coefficient contours on wing cross sections and surfaces of the standard configuration. 483
37
a) 3D view with six planes
d) Y=-0.46m
a) Upper surface
b) Lower surface
e) Y=-0.55m
f) Y=-0.66m
Fig. 25 Pressure coefficient contours on waverider forebody, wings and surface of the dash configuration. 484 485
4.3. Analysis of wing downwash
486
Pressure coefficient contours on a chord wise wing and tail cross section (Y=0.46m) with
487
flowfield streamtraces of three wing-extended configurations are shown in Fig. 26. This flow
488
establishes a circulatory motion that trails downstream of the wing. These wing-tip vortices
489
downstream of the wing induce a small downward component of air velocity in the neighborhood of
490
the wing itself. This downward component is called downwash in Ref. [32]. As can be seen in Fig. 26,
491
the flowfield near the leading edge of the horizontal tail is nearly the same as far field air flow. And
492
the downward component induced by the wing downwash is very small.
38
a)
loiter configuration
b)
standard configuration
c)
dash configuration
Fig. 26 Pressure coefficient contours on a wing and tail cross section (Y=0.5m) with flowfield streamtraces of three wing-extended configurations (M=0.8, Angle of attack = 10°) 493
The dash configuration is chosen as an example to investigate the downwash effect on the tail,
494
since the distance between wing and tail of the dash configuration is the nearest. Pressure coefficient
495
contours on a chord wise wing and tail cross section (Y=0.5m) of dash configuration with flowfield
496
streamtraces of three flight condition are shown in Fig. 27. As can be seen in Fig. 26, the differences
497
of downwash effects among three supersonic/hypersonic flight conditions are quite small.
a)
M=3
b)
M=6
c)
M=8
Fig. 27 Pressure coefficient contours on a wing and tail cross section (Y=0.5m) of dash configuration with flowfield streamtraces of three flight condition (Angle of attack = 10°) 498
39
499
4.4. Effects of shock wave
500
Under the supersonic flight condition(M=3, AoA=0deg), comparisons of Mach number contours
501
of loiter, standard and dash configurations at seven cross sections are shown in Fig. 28, Fig. 29 and
502
Fig. 30, respectively. For a clear demonstration, the grids with Mach number greater than 2.9 are not
503
drawn in the right subfigure. Attached bow shock on leading edge of the forebody can be observed
504
explicitly as shown in X direction cross sections at Fig. 28 (b) to Fig. 30 (b). The shock wave on the
505
forebody prevents the leakage of the high-pressure gas on the lower surface of the waverider. The
506
Mach contours along the wing span wise direction shows that effect of forebody shock wave
507
gradually decreases with increase of the distance of y direction from wing root.
a) Z=-0.09m
b) X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 28 Comparisons of Mach number contours of loiter configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m and Z=-0.09m) under the flight condition (M=3, AoA=0e) 508
40
a) Z=-0.09m
b) X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 29 Comparisons of Mach number contours of standard configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m and Z=-0.09m) under the flight condition (M=3, AoA=0e) 509 510
As shown in Fig. 30, two variable-sweep-wings are totally under the influence of forebody shock wave.
a) Z=-0.09m
b) X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 30 Comparisons of Mach number contours of dash configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m and Z=-0.09m) under the flight condition (M=3, AoA=0e) 511
Under the supersonic flight condition(M=6,AoA=0°), comparisons of Mach number contours of
512
loiter, standard and dash configurations at seven cross sections are shown in Fig. 31, Fig. 32 and Fig.
513
33, respectively. For a clear demonstration, the grids with Mach number greater than 5.9 are not
514
drawn in the right subfigure. Mach number contours in Fig. 31 to Fig. 33 allow assessing the bow 41
515
shock shape that takes place past the vehicle. Attached bow shock on leading edge of the forebody can
516
be observed explicitly as shown in X direction cross sections at Fig. 31 (b) to Fig. 33 (b).
a) Z=-0.09m
b) X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 31 Comparisons of Mach number contours of loiter configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m and Z=-0.09m) under the flight condition (M=6, AoA=0e) 517
a) Z=-0.09m
b) X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 32 Comparisons of Mach number contours of standard configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m and Z=-0.09m) under the flight condition (M=6, AoA=0e) 518
42
a) Z=-0.09m
b) X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 33 Comparisons of Mach number contours of dash configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m and Z=-0.09m) under the flight condition (M=6, AoA=0e) 519
Under the hypersonic flight condition(M=8,AoA=0°), comparisons of Mach number contours of
520
loiter, standard and dash configurations at seven cross sections are shown in Fig. 34, Fig. 35and Fig.
521
36, respectively. For a clear demonstration, the grids with Mach number greater than 7.9 are not
522
drawn in the right subfigure. Attached bow shock on leading edge of the forebody can be observed
523
explicitly as shown in X direction cross sections at Fig. 34 (b) to Fig. 36 (b).
a)
Z=-0.09m
b)
X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 34 Comparisons of Mach number contours of loiter configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m) under the flight condition (M=8, AoA=0e) 524 43
a)
Z=-0.09m
b)
X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 35 Comparisons of Mach number contours of standard configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m) under the flight condition (M=8, AoA=0e) 525
a)
Z=-0.09m
b)
X=0.1m, 0.2m, 0.3m and Y=0.42, 0.5, 0.6m
Fig. 36 Comparisons of Mach number contours of dash configuration at seven cross sections (X=0.1m, 0.2m, 0.3m, Y=0.42, 0.5, 0.6m) under the flight condition (M=8, AoA=0e) 526
5. Conclusions and future work
527
In this paper, the design methodology and aerodynamic performances of a variable-sweep-wing
528
waverider are proposed and analyzed respectively. Besides, the aerodynamic performances of four
529
different configurations are compared with each other, and we have come to the following
530
conclusions:
531
z
Under the same flight condition, the drag coefficient of the four different sweep 44
532
configurations increases with the increase of the angle of attack, and it becomes lower as the
533
sweep angle increases. Under the same angle of attack and the same sweep configuration,
534
the drag coefficient decreases with the increase of the flight Mach number.
535
z
In terms of the increase in the lift, under the flight conditions (M= 0.8, M=1.5), the loiter
536
configuration has the largest lift coefficient. The optimal configurations with the largest lift
537
coefficient for different flight conditions are the followings, the standard configuration for
538
Mach 3, and the dash configuration for Mach 6 and Mach 8. This shows that different
539
configurations of the variable-sweep-wing morphing waverider suit for different flight
540
conditions with the largest lift coefficient.
541
z
Under the same flight condition, the lift-to-drag ratio of the four different sweep
542
configurations increases first and then decreases with the increase of the angle of attack.
543
Under the flight condition (M=0.8), the lift-to-drag ratio of the loiter configuration
544
(L/D=8.43, Į=2°) is the highest, but they are close at high angles of attack. In contrast,
545
under the flight condition (M=1.5), the standard configuration has the largest lift-to-drag
546
ratio of 3.29. Under the flight condition (M=3,6,8), the dash configuration has the largest
547
lift-to-drag ratio when Į=5°.
548
z
Though analyses of effects of wing downwash and shock wave, the flowfield near the
549
leading edge of the horizontal tail is nearly the same as far field air flow. And the downward
550
component induced by the wing downwash is very small. Attached bow shock on leading
551
edge of the forebody can be observed. The Mach contours along the wing span wise
552
direction shows that effect of forebody shock wave gradually decreases with increase of the
553
distance of y direction from wing root. 45
554
In conclusion, the variable-sweep-wing morphing waverider proposed in this paper improves
555
both low speed take-off and high-speed cruising performances, and it expands the flight speed range.
556
For different flight conditions and mission scenarios, the best flight performance of this
557
variable-sweep-wing morphing waverider can be achieved using different configurations.
558
46
559 560 561 562
Conflicts of interest The authors declare there is no conflict of interest regarding the publication of this paper. Acknowledgments The authors would like to express their thanks for the support from NSAF (Grant No:
563
U1730135).
564
References
565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596
[1] G. Pezzella, and A. Viviani, Aerodynamic performance analysis of a winged re-entry vehicle from hypersonic down to subsonic speed, Aerosp. Sci. Technol. 52(2016) 129-143. [2] S. B. Li, Z. G. Wang, W. Huang, J. Lei, S. R. Xu, Design and investigation on variable Mach number waverider for a wide-speed range. Aerosp. Sci. Technol. 76(2018) 291-302. [3] T.T. Zhang, Z.G. Wang, W. Huang, S.B. Li, A design approach of wide-speed-range vehicles based on the cone-derived theory, Aerosp. Sci. Technol. 71 (2017) 42–51. [4] Z.T. Zhao, W. Huang, S.B. Li, T.T. Zhang, L. Yan, Variable Mach number design approach for a parallel waverider with a wide-speed range based on the osculating cone theory, Acta Astronaut. 147 (2018) 163–174. [5] Z.T. Zhao, W. Huang, B.B. Yan, L. Yan, T.T. Zhang, R. Moradi, Design and high speed aerodynamic performance analysis of vortex lift waverider with a wide-speed range, Acta Astronaut. 151 (2018) 848–863. [6] F.M. Wang, H.H. Ding, M.F. Lei, Aerodynamic characteristics research on wide speed range waverider configuration, Sci. China Ser. E-Tech. Sci. 52 (10) (2009) 2903–2910. [7] S.B. Li, S.B. Luo, W. Huang, Z. G. Wang, Influence of the connection section on the aerodynamic performance of the tandem waverider in a wide-speed range, Aerosp. Sci. Technol. 30 (2013) 50–65. [8] S.B. Li, W. Huang, Z.G. Wang, J. Lei, Design and aerodynamic investigation of a parallel vehicle on a wide-speed range, Sci. China Inf. Sci. 57 (12) (2014) 128201:1-128201:10. [9] Z. Liu, J. Liu, F. Ding, Z. X. Xia, Novel methodology for wide-ranged multistage morphing waverider based on conical theory, Acta Astronaut 140(2017) 362-369. [10] R.J. Pegg, D.E. Hahne, C.E. Cockrell, et al., Low-speed wind tunnel tests of two waverider configuration models, 6th International Aerospace Planes and Hypersonics Technologies Conference, AIAA Paper. (1995) 6093. [11] C. Tarpley, D. J. Pines, A. P. Kothari, Low-speed stability analysis of the dual fuel waverider configuration. Space Plane & Hypersonic Systems & Technology Conference. (2013) [12] Y. Takama, Practical waverider with outer wings for the improvement of low-speed aerodynamic performance. Aiaa International Space Planes & Hypersonic Systems & Technologies Conference. (2011) [13] Zhao Z. T., W. Huang, L. Yan, T. T. Zhang, S. B. Li, F. Wei, Low speed aerodynamic performance analysis of vortex lift waveriders with a wide-speed range, Acta Astronaut. 161(2019) 209-221. 47
597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638
[14] A. P. Austin, E. R. Robert, E. R. Jesse, B. G. Gabriel, Mach Five to Ten Morphing Waverider: Control Point Study,J. Aircr. 56(2)(2019) 493-504. [15] D. Grant, M. Abdulrahim, R. Lind, Flight dynamics of a morphing aircraft utilizing independent multiple-joint wing sweep, in: AIAA Atmospheric Flight Mechanics Conference and Exhibit, (2006) 6505. [16] A. Chakravarthy, D. Grant, R. Lind, Time-varying dynamics of a micro air vehicle with variable-sweep morphing, J. Guid. Control Dyn. 35 (3) (2012) 890–903. [17] T. Yue, L. Wang, J. Ai, Longitudinal linear parameter varying modeling and simulation of morphing aircraft, J. Aircr. 50 (6) (2013) 1673–1681. [18] B. B. Yan, P. Dai, R. F. Liu, M. Z. Xing, S. X. Liu, Adaptive super-twisting sliding mode control of variable sweep morphing aircraft, Aerosp. Sci. Technol. 92 (2019) 198-210. [19] D. Xu, Z. Hui, Y. Q. Liu, G. Chen, Morphing control of a new bionic morphing UAV with deep reinforcement learning, Aerosp. Sci. Technol. 92 (2019) 232-243. [20] T. Yue, X. Zhang, L. Wang, J. Ai, Flight dynamic modeling and control for a telescopic wing morphing aircraft via asymmetric wing morphing, Aerosp. Sci. Technol. (2017) 328-338. [21] Y. Y. Guo, B. Xu, X. X. Hu, X. W. Bu, Y. Zhang. Two controller designs of hypersonic flight vehicle under actuator dynamics and AOA constraint. Aerosp. Sci. Technol. 80 (2018) 11-19. [22] S. B. Tiwari, R. Suresh, C. K. Krishnadasan, Design, Analysis and Qualification of Elevon for Reusable Launch Vehicle. J. I. of Eng (India): Series C. (2017). [23] S. Chakravarthy, O. Peroomian, U. Goldberg, S. Palaniswamy, Parallel personal computer applications of cfd++, a unified-grid, unified physics framework, Parallel Comput Fluid Dyn. (1999) 127-134. [24] W. Huang, W.D. Liu, S.B. Li, Z.X. Xia, et al., Influences of the turbulence model and the slot width on the transverse slot injection flow field in supersonic flows, Acta Astronaut. 73 (2012) 1–9. [25] W. Huang, S.B. Li, L. Yan, Z.G. Wang, Performance evaluation and parametric analysis on cantilevered ramp injector in supersonic flows, Acta Astronaut. 84 (2013) 141–152. [26] O. Brodersen, M. Rakowitz, S. Amant, Airbus, ONERA, and DLR Results from the Second AIAA Drag Prediction Workshop, J. Aircr. 42 (2005) 932-940. [27] R. Rudnik, M. Sitzmann, J. L. Godard, F. Lebrun, Experimental Investigation of the Wing-Body Juncture Flow on the DLR-F6 Configuration in the ONERA S2MA Facility, Aiaa Applied Aerodynamics Conference. (2013). [28] https://aiaa-dpw.larc.nasa.gov/Workshop2/ pdf/53Goldberg_cfd++.pdf [29] K. R. Laflin, S. M. Klausmeyer, T. Zickuhr, J. C. Vassberg, R. A. Wahls, J. H. Morrison, Data summary from second aiaa computational fluid dynamics drag prediction workshop, J. Aircr. 42(5)(2005) 1165-1178. [30] A. Singh, Experimental Study of Slender Vehicles at Hypersonic Speeds, Cranfield University, Cranfield, 1996 (Ph.D.). [31] U. Goldberg, Hypersonic Turbulent Flow Predictions Using CFD++, AIAA/CIRA 13th International Space Planes and Hypersonics Systems and Technologies Conference. (2005). [32] Fundamentals of aerodynamics(fifth edition). John Anderson Jr.
48
Declaration of Competing Interest
Ǥ