You are on page 1of 8

Aerospace Science and Technology 15 (2011) 129136

Contents lists available at ScienceDirect


Aerospace Science and Technology
www.elsevier.com/locate/aescte
Numerical investigation of unsteady aerodynamic characteristics of a pitching
missile
Yang Lizhi
a,
, Wang Minghai
a
, Gao Zhenghong
b
a
Xian Research Inst. of Hi-Tech, Hongqing Town, Xian 710025, China
b
Department of Aircraft Engineering, Northwestern Polytechnical University, Xian 710072, China
a r t i c l e i n f o a b s t r a c t
Article history:
Received 5 July 2009
Received in revised form 2 April 2010
Accepted 4 July 2010
Available online 16 July 2010
Keywords:
Unsteady ow
Computational method
Hysteresis
Three-dimensional Reynolds averaged NavierStokes numerical simulations were carried out to predict
the aerodynamic loads of a pitching winged missile based on the nite volume method. The Baldwin
Lomax eddy viscosity model with the modications suggested by Degani and Schiff was used here. The
computational results of the aerodynamic loads of a slender revolution body are also given, and agreed
well with the experiment data. The unsteady aerodynamics about the winged missile, which is oscillated
in pitching with different frequencies, with different center of rotation and at different range of angle of
attack is shown in the paper. It is found that from the computational results, the unsteady hysteresis of
the lift, drag and pitch moment coecients vary not only with the frequency, but also with the position
about which the missile performs the oscillatory motions. Although the angle of attack range at which
the missile oscillate in pitching are different, the derivative of lift coecients with respect to the angle
of attack is the same for the same phase of oscillating under the same reduced frequency condition.
2010 Elsevier Masson SAS. All rights reserved.
1. Introduction
Over the last several decades unsteady stall phenomena have
been topics of intensive research. Dynamic lift utilizes the hys-
teresis effects of airfoils or wings pitching up at rapid rates to
delay the onset of stall. It is well known that in two-dimensional
unsteady ow separation does occurs at greater incidence angles
than in the steady case, attainable lift coecients being signi-
cantly higher. In addition, leading-edge vortices form that aid in
the development of lift. Several researchers have shown the effects
of dynamic lift (or dynamic stall) on airfoils, both with experi-
mental and numerical studies [24,18,17,8,6,3,5]. In fact, excellent
review articles on dynamic stall of airfoils have been written by
Ekaterinaris and Platzer [14], as well as Carr [7]. Experimental
and numerical studies have also been conducted on wings [20,
9,23], especially delta wings under dynamic conditions are well
documented, see Refs. [21,15,16,2], for example. Under dynamic
pitching conditions, where the delta wing is pitched upwards, the
location of the vortex burst is farther towards the trailing edge
compared to the same angle of attack under static conditions. This
produces a phase lag in the burst location, allowing transient val-
ues of lift to exceed those obtained during static testing. Similarly,
pitching down the delta wing moves the vortex burst location for-
ward, this produces a phase lead and a reduction in lift, compared
*
Corresponding author. Tel.: +86 029 83208080.
E-mail address: lizhiyang1972@yahoo.com.cn (L. Yang).
to the similar static angle of attack. This introduces the notion of
a hysteresis effect, or time delay, where there is a difference in
the measured C
L
values if the angle of attack is increasing or de-
creasing. Modern aircraft use either slender delta wings or similar
devices (e.g., leading-edge extensions) and take advantage of this
hysteresis effect of the delta wing, in an attempt to increase the
performance envelope.
Some work has been done on studying the dynamic lifting
characteristics of full aircraft congurations. Cummings et al. [10]
showed the dynamic lifting capabilities of a UCAV conguration
both with experimental and numerical studies, for better under-
standing the impact of vortex lift and vortex breakdown, coupled
with dynamic lift, for these congurations.
This study is an effort aimed at furthering the knowledge and
understanding the characteristics of the dynamic aerodynamics for
a generic cruise missile. The mission of the missile is to hit a xed
target while minimizing the exposure to anti-air defenses. This
means that the ight altitude should be as low as possible, but
the impact must be achieved by a vertical dive. This leads to the
generic trajectory shape where the missile initially ies straight
and level at the minimum altitude. When it approaches the tar-
get, it must climb (nose up) to gain enough height for the nal
dive (nose down). This up-and-down terminal maneuver is known
as the bunt. Although the angle of attack is not high in the bunt
process, the rates at which the missile pitches up and down may
be very rapid [25]. And very little work has been done on the
studying the dynamic characteristics of the aerodynamic loads of a
cruise missile in the bunt maneuver. In this paper the characteris-
1270-9638/$ see front matter 2010 Elsevier Masson SAS. All rights reserved.
doi:10.1016/j.ast.2010.07.001
130 L. Yang et al. / Aerospace Science and Technology 15 (2011) 129136
tics of the unsteady aerodynamics about a winged missile, which
is oscillated in pitching with different frequencies, with different
center of rotation and at different range of angle of attack, is stud-
ied numerically.
2. Computational method
It is known that the cost of performing wind tunnel or ight
experiments in unsteady ows is very high. Numerical simula-
tion of unsteady ows is becoming a powerful tool as a result
of continuous improvements in numerical methods as well as the
development of speed and storage capability of computers. Past
research efforts concerning the development of CFD methods for
simulating dynamic stall have been reviewed in Ref. [5]. In re-
cent years the progress has been made in developing and applying
high-resolution methods to unsteady aerodynamic ows see, for
example, Drikakis et al. [12] and Hahn and Drikakis [19]. Although
great progress has been made for simulation methods like Di-
rect Numerical Simulation (DNS) and Large-Eddy Simulation (LES),
these two methods are currently outside of the realm of indus-
trial applications for Reynolds numbers typically found in external
aerodynamics and will remain so for quite some time [6,13]. For
the time being, unsteady Reynolds-averaged NavierStokes simula-
tion (URANS) seem to be the only feasible way for simulating ows
around a pitching missile at high Reynolds numbers.
2.1. Governing equations
The conservation form of the 3-D dimensionless, unsteady,
compressible, Reynolds averaged NavierStokes equations in Carte-
sian coordinates is as follows
Q
t
+
F
x
+
G
y
+
H
z
=
1
Re

x
+
G

y
+
H

(1)
here
Q =

u
v
w
E

, F =

u
u
2
+ p
uv
uw
Hu

G =

v
uv
v
2
+ p
vw
Hv

, H =

w
uw
vw
w
2
+ p
Hw

where p, , (u, v, w), E and H are pressure, density, the Carte-


sian velocity components, total specic energy and total specic
enthalpy, respectively, and H = E + p/, Re is Reynolds number,
F
v
, G
v
, H
v
are viscous terms, and the details can be found, for in-
stance, in Ref. [1].
2.2. Numerical implementation
Considering a control volume bounded by the closed surface
, we may rst write the integral form of Eq. (1) as:

QdVol +

N ndS
1
Re

N
v
ndS =0 (2)
with
N =Fi +Gj +Hk
N
v
=F
v
i +G
v
j +H
v
k
Fig. 1. Gas boundary of imagination for a cross section.
Then we apply Eq. (2) to the given computational mesh system.
To solve these equations, the nite volume scheme with explicit
multi-stage RungeKutta time stepping formulated by Jameson et
al. [22] is used here. In order to accelerate the computations, the
implicit residual smoothing technique with local variable parame-
ters [16] is applied in the present paper.
2.3. Boundary condition
(1) At the solid surface boundary: no-slip and no thermal conduc-
tion
u
w
= v
w
= w
w
=0,

T
n

w
=0
(2) At the outside boundary: the characteristic variable boundary
conditions [16] are used here.
2.4. Turbulent model
The BaldwinLomax eddy viscosity model with the modica-
tions suggested by Degani and Schiff [11] was used here. The B-L
model [4] is simplistic and it is known that the model can perform
well for attached or little separated ows, but it leads to large
inaccuracies for high angle of attack and largely separated ows.
Previous studies have examined the effects of turbulence models
on moving boundaries see, for example, Barakos and Drikakis [6].
For unsteady ow computation, the CPU time per step doubles
when a two-equation model is employed instead of an algebraic
one on oscillating and ramping aerofoils in the study [6]. Here the
B-L model was used for two reasons, one is that the angle of at-
tack range at which the missile oscillate in pitching and the mean
angle of attack are small in the paper, the other is that it will not
lead to long computing times.
2.5. Grid generation
An OH grid is generated by combining a series of two-
dimensional meshes. Each mesh is generated previously by confor-
mal mapping technique in planes normal to the longitudinal axis
of the missile. In two-dimensional grid generation processes, the
leading edge of the wing is stretched outward to be a boundary of
imagination, called gas boundary, shown in Fig. 1, in order to get a
high quality of the grid. The gas boundary has a thickness of zero
in the computational process.
3. The results and analysis
3.1. The body of revolution
To validate the computational results, the surface pressure of
a body of revolution at different angles of attack is studied rst,
and the results are compared with the experimental data [26].
The body had an ogival nose 3 diameters long (3D), tangent to a
cylindrical afterbody 7.7 diameter long (7.7D). The dimensions and
details of the model are shown in Ref. [26]. The origin of Cartesian
coordinates is at the nose vertex of the body, and x axis coincident
with the body axis, and =0

is in the vertical plane of symme-


try on the windward side. Results are presented for Mach number
L. Yang et al. / Aerospace Science and Technology 15 (2011) 129136 131
Table 1
Difference between the mesh independence calculations.
C
A
C
N
86 120 90 cells 0.28% 2.51%
86 200 90 cells 0.20% 0.26%
M

= 0.3 at Reynolds number Re = 0.44 10


6
, as same as the
test condition in Ref. [26] for comparison. In Ref. [26] the model
was sting supported in the test section, and the sting deection
due to aerodynamic loads is estimated to be less than 0.1

. No
corrections were applied to the data to account for this deection.
Wind-tunnel wall corrections were considered negligible because
the model was small compared to the test section.
A grid independence study was performed in order to ensure
that the mesh was rened suciently. An OH grid of 86 120
90 in the axial, circumferential, and radial directions respectively,
was used as the coarsest mesh. The grid was generated for a half-
body. The distance from the wall to the rst cell center was less
than 2 10
6
D. In addition, the grid was clustered in the vicinity
of the body nose in attempt to accurately capture solution varia-
tions in that region. The second grid was generated by adding 40
nodes in the circumferential direction to the coarsest mesh. Finally,
the nest mesh used for the study had 200 nodes in the circumfer-
ential direction. No additional grid points were added in the axial
direction for this mesh because of its accuracy, as well as our ex-
perience in using it for aerodynamic loads computation in similar
case before this study. The grid independence study was performed
with the body of revolution at =15

. In Table 1, the magnitude


of the calculated axial force coecient, C
A
, and the normal force
coecient, C
N
, for the coarsest and nest mesh are given relative
to that of the intermediate grid.
The relative magnitude of the calculated normal force coef-
cient between the coarse mesh and the intermediate mesh is
around 2.5%, which hints at some grid dependence effect. On the
other hand, comparison of the normal force coecients obtained
using the nest mesh with that of the intermediate mesh indi-
cate that increasing the number of cells to 1.5 million has only a
small impact on the results. Therefore, it was decided to proceed
with the computations using the intermediate mesh composed of
1.2 million cells.
The circumferential pressure distribution computed at several
longitudinal positions on the ogival nose and cylindrical afterbody
at angle of attack of 10

, 15

, and 20

are shown in Fig. 2. The


results of experiment [26] have been included in this gure for
comparison. Fig. 2 shows a good agreement between C
p
obtained
from the present study and that obtained experimentally at lon-
gitudinal positions x/D = 0.5, x/D = 2.0, x/D = 6.0, while at the
position x/D =10.0, the computed results underestimates the ex-
perimental data. This difference could be attributed to the effect of
turbulence at the section near the aft of the body and the turbu-
lence modeling cannot simulate the massive separated ow very
well.
In order to validate the computational results, comparisons be-
tween experiments and numerical results for the normal force
distributions (local normal force coecient C
n
) along the body
length (x/D) at three different angles of attack are also presented
in Fig. 3. It was shown that the CFD matched accurately the ex-
periment results on the nose of the body of revolution (before
x/d =3.0), but the magnitude of the normal force is not captured
as well on the afterbody. Differences between the numerical re-
sults and experimental data in Fig. 3 are subsequently reected on
the whole normal force coecients (C
N
) of the body, which was
shown in Fig. 4.
Fig. 2. Circumferential pressure distributions of cross section (M

=0.3).
132 L. Yang et al. / Aerospace Science and Technology 15 (2011) 129136
Fig. 2. (continued)
Fig. 3. Normal-force distributions along the body length.
Fig. 4. Normal-force coecients of the body vs .
Table 2
Difference between the mesh independence calculations.
C
A
C
N
86 160 90 cells 0.31% 1.30%
86 240 90 cells 0.22% 0.19%
Fig. 5. The surface grid of the missile.
Fig. 6. Crosswise section grid at the wing.
Fig. 7. Crosswise section grid at the tail.
L. Yang et al. / Aerospace Science and Technology 15 (2011) 129136 133
Fig. 8. Lift, drag and pitching moment coecient vs for different reduced fre-
quency.
3.2. The cruise missile
In this case, a generic cruise missile oscillates in pitching with
(t) =
m

m
cos()
where
m
is the mean angle of attack, which was varied to ob-
tain results for three ranges of angle of attack, 0

, 0

, and 0

, that is
m
equal to 0.5

, 1.5

, and 2.5

,
respectively. is the phase angle of oscillation
=t, =2 f , f =

r
k
Here r
k
is the reduced frequency. This pitching function was used
since it produces a motion without any discontinuities in acceler-
ation at the beginning and end of the motion, thus being easier to
implement in a CFD programme.
The missile oscillates in pitching about three longitudinal loca-
tions on the conguration, x/L =1.6, x/L =1.73 (about 1/4 chord)
and x/L =1.912, here L is the half span of the missile (shown in
Fig. 5). Three values of reduced frequency are used: 0.336, 0.5 and
0.628. These three reduced frequency are chosen for two reasons,
one is that the three number are the magnitude of general reduced
frequency countered in the real ight, the other is that compar-
isons between the results of this study and the study before can
be made. Here the free stream Mach number and Reynolds num-
ber are 0.6 and 1.61 10
6
, respectively.
A grid independence study was also performed with the missile
at =1

. For this study, an OH grid of 8616090 in the axial,


circumferential, and radial directions respectively, was used as the
coarsest mesh. The grid was generated for half of the conguration.
The minimum spacing normal to the wall surface was 1.0 10
6
.
The intermediate grid was generated by adding 40 nodes in the cir-
cumferential direction to the coarsest mesh. The nest mesh used
for the study had 240 nodes in the circumferential direction. In
the axial direction, 76, 86, and 96 grid points were used for the
grid system, respectively, and 86 grid points were taken as a good
compromise between the coarse and ne grid points, and still has
adequate delity in the axial direction. The magnitude of the cal-
culated axial force coecient, C
A
, and the normal force coecient,
C
N
, for the coarsest and nest mesh are given relative to that of
the intermediate grid, shown in Table 2.
Comparison of the aerodynamic coecients obtained using the
nest mesh with that of the intermediate mesh indicate that in-
creasing the number of cells has only a small impact on the results.
Therefore, the intermediate mesh was used in the study. The ge-
ometry and the surface grid of the missile are shown in Fig. 5, and
the corresponding crosswise section grid at the wing and tail of
the missile are shown in Figs. 6 and 7, respectively.
Fig. 8 shows the variation of lift, drag, and pitching moment
coecients with for the three values of reduced frequency,
r
k
= 0.336, 0.5 and 0.628. The missile oscillates in pitching about
the longitudinal locations on the conguration, x/L = 1.73,
m
equal to 1.5

, and the range of angle of attack is 0

.
The static aerodynamic coecients results are also presented for
reference. Notice the hysteresis for 0

, with increased lift


(relative to the static case) being obtained during the pitch-up mo-
tion. During the pitch-down motion there is decreased lift for the
remainder of the cycle. In addition, the hysteresis effect increases
as the reduce frequency increases. The pressure coecient distri-
butions over leeward side of the wing of the missile for =1

in
steady case and pitch up case (r
k
= 0.628) are shown in Fig. 9(a)
and (b), respectively.
Fig. 10 shows the variation of lift, drag, and pitching moment
coecients with about three longitudinal locations on the con-
guration, x/L = 1.6, x/L = 1.73 and x/L = 1.912. The reduced
frequency r
k
= 0.628. Notice that each of these cases has some-
what different characteristics at the same reduced frequency. The
results for C
D
and C
M
only show slight differences when compared
with C
L
results.
The lift coecients of the missile which oscillates in pitching
at three ranges of angle of attack 0

, 0

, and
0

at two cases of reduced frequency r


k
= 0.336 and
r
k
=0.628 are presented in Figs. 11 and 12, respectively. The mis-
sile oscillates in pitching about the longitudinal locations on the
conguration x/L =1.73. Notice that while the hysteresis effect in-
creases as the angle of attack range increases, the curves of the lift
coecient vs seems similar for the three ranges of angle of at-
tack. Therefore the derivative of lift coecients with respect to the
angle of attack, C

L
, was calculated, shown in Fig. 13. It shows that
while the angle of attack ranges at which the missile oscillates in
134 L. Yang et al. / Aerospace Science and Technology 15 (2011) 129136
Fig. 9. Pressure contours over leeward side of the wing.
pitching are different, the derivative C

L
is the same for the same
phase of oscillating under the same reduced frequency condi-
tion. The variations of the derivative C

L
vs are different from
each other for different reduced frequency 0.336 and 0.628.
Here the derivative C

L
is calculated for the rst cycle, since the
lift coecients for the second cycle are nearly the same as those
in the rst cycle. So the phase of oscillation is between 0 and
2 in Fig. 13. The variation of the derivative C

L
vs in pitch-
up and pitch-down are shown in Fig. 13(a) and (b), respectively.
For comparison purpose, the derivatives C

L
where is near 0,
and 2 are not given because the value of the derivative there is
very big or even innite comparing with the most value of the
derivatives.
Considering the reason why the derivative C

L
is the same
for the same phase of oscillating under the same reduced fre-
quency condition at different ranges of angle of attack shown in
Fig. 13. In this case, take the lift coecients as the function of
t, x, , ,
...
, . . . . That is
C
L
=C
L
(t, x, , ,
...
, . . .)
here x indicates some information of oweld, for example, the
location of ow separation. The missile oscillates in pitching with
(t) =
m

m
cos()
the derivative of the angle of attack with respect to time
(t) =
m
sin(t)
the second order derivative of the angle of attack with respect to
time
(t) =
m

2
cos(t), . . .
That is
C
L
=C
L
(t, x, ,
m
)
Here take x as the same information, because the angles of attack
for the missile which oscillates in pitching at the three ranges of
angle of attack are small, and the information of the oweld dy-
namic system are similar.
Therefore
C
L
=C
L
(t, ,
m
)
The derivative
C

L
=
dC
L
d
=
dC
L
/dt
d/dt
=
dC
L
/dt

m
sin(t)
Here the function (dC
L
/dt)/
m
must be the same for the three
ranges of angle of attack because of the same derivative C

L
for the
same phase of oscillating , that is (t).
The variation of (dC
L
/dt)/(2
m
) vs time is shown in Fig. 14.
It shows that the function (dC
L
/dt)/
m
is the same for the three
ranges of angle of attack. From the above reasoning, we may reach
a conclusion that the derivative C

L
is the same for the same phase
of oscillation (t) when the oweld going into the same dynamic
systems.
4. Conclusion
Three-dimensional Reynolds averaged NavierStokes numerical
simulations were carried out to predict the oweld characteristics
of a pitching winged missile based on the nite volume method.
The Baldwin-Lomax eddy viscosity model with the modications
suggested by Degani and Schiff was used here. In order to validate
the computational results, the aerodynamic loads of a slender rev-
olution body investigated numerically are also given, and agreed
well with the experiment data.
The characteristics of the unsteady aerodynamics about a
winged missile, which is oscillated in pitching with different fre-
quencies, with different center of rotation and at different range of
angle of attack are shown in this paper. It is found that from the
computational results, the unsteady hysteresis of the lift, drag and
pitch moment coecients vary not only with the frequency, but
also with the position of the oscillating axis. Although the angle of
attack range at which the missile oscillate in pitching are different,
the derivative of lift coecients with respect to the angle of attack
is the same for the same phase of oscillating under the same re-
duced frequency condition when the oweld of the missile going
into the same dynamic systems.
L. Yang et al. / Aerospace Science and Technology 15 (2011) 129136 135
Fig. 10. Lift, drag and pitching moment coecient vs about different longitudinal
locations.
Fig. 11. Lift coecient for different ranges of angle of attack (r
k
=0.336).
Fig. 12. Lift coecient for different ranges of angle of attack (r
k
=0.628).
Fig. 13. The derivative C

L
vs in pitching-up and pitching-down.
Fig. 14. The function (dC
L
/dt)/(2
m
) vs time.
136 L. Yang et al. / Aerospace Science and Technology 15 (2011) 129136
References
[1] J.D. Anderson, Governing equations of uid dynamics, in: J.F. Wendt (Ed.), In-
troduction to Computational Fluid Dynamics, Springer-Verlag, 1992, pp. 1551,
Chapter 2.
[2] M.T. Arthur, F. Brandsma, N. Ceresola, W. Kordulla, Time accurate Euler calcula-
tions of vortical ow on a delta wing in pitching motion, AIAA Paper 99-3110,
1999.
[3] K.J. Badcock, F. Cantariti, I. Hawkins, M. Woodgate, L. Dubuc, B.E. Richards, Sim-
ulation of unsteady turbulent ows around moving aerofoils using the pseudo-
time method, International Journal for Numerical Methods in Fluids 32 (5)
(2000) 585604.
[4] B.S. Baldwin, H. Lomax, Thin layer approximation and algebraic model for sep-
arated turbulent ows, AIAA Paper 78-257, 1978.
[5] G. Barakos, D. Drikakis, An implicit unfactored method for unsteady turbu-
lent compressible ows with moving boundaries, Computers and Fluids 28 (8)
(1999) 899922.
[6] G. Barakos, D. Drikakis, Computational study of unsteady turbulent ows
around oscillating and ramping aerofoils, International Journal for Numerical
Methods in Fluids 42 (2) (2003) 163186.
[7] L.W. Carr, Progress in analysis and prediction of dynamic stall, Journal of Air-
craft 25 (1) (1988) 617.
[8] L.W. Carr, M.S. Chandrasekhara, M.C. Wilder, Effect of compressibility on sup-
pression of dynamic stall using a slotted airfoil, Journal of Aircraft 38 (2) (2001)
296309.
[9] F.N. Coton, R.A. Galbraith, An experimental study of dynamic stall on a nite
wing, Aeronautical Journal 103 (May 1999) 229236.
[10] R.M. Cummings, S.A. Morton, S.G. Siegel, S. Bosscher, Numerical prediction and
wind tunnel experiment for a pitching unmanned combat air vehical, AIAA
Paper 2003-0417, 2003.
[11] D. Degani, L.B. Schiff, Computation of turbulent supersonic ows around
pointed bodies having crossow separation, Journal of Computational
Physics 66 (1986) 173196.
[12] D. Drikakis, M. Hahn, A. Mosedale, B. Thornber, Large eddy simulation us-
ing high-resolution and high-order methods, Philosophical Transactions of the
Royal Society A 367 (2009) 29852997.
[13] P.A. Durbin, A perspective on recent developments in RANS modeling, in:
W. Rodi, N. Fueyo (Eds.), Engineering Turbulence Modeling and Experiments,
vol. 5, Elsevier, Amsterdam, 2002, pp. 316.
[14] J.A. Ekaterinaris, M.F. Platzer, Computational prediction of airfoil dynamic stall,
Progress in the Aerospace Science 33 (1997) 759846.
[15] L.E. Ericsson, Complex angular amplitude and frequency effects on delta wing
unsteady aerodynamics, AIAA Paper 99-4009, 1999.
[16] Z. Gao, Research on the hysteresis properties of unsteady aerodynamics
about the oscillating wings, Applied Mathematics and Mechanics (English Edi-
tion) 20 (8) (1999) 895907.
[17] D. Greenblatt, D. Neuberger, I. Wygnanski, Dynamic stall control by intermit-
tent periodic excitation, Journal of Aircraft 38 (1) (2001) 188190.
[18] E. Guilmineau, P. Queutey, Numerical study of dynamic stall on several airfoil
sections, AIAA Journal 37 (1) (1999) 128130.
[19] M. Hahn, D. Drikakis, Implicit large-eddy simulation of swept-wing ow using
high-resolution methods, AIAA Journal 47 (3) (2009) 618629.
[20] J. Henkner, Phenomena of dynamic stall on swept wings, in: 22nd International
Congress of the Aeronautical Sciences, ICAS Paper 2000-2.9.2, Aug.Sep. 2000.
[21] I. Heron, R.Y. Myose, Delta wing vortex burst behavior under dynamic
freestream, Part 1 Fast pitch-up during deceleration, AIAA Paper 2007-6725,
Aug. 2007.
[22] A. Jameson, W. Schmidt, E. Turkel, Numerical solutions of the Euler equations
by nite volume methods using RungeKutta time stepping scheme, AIAA Pa-
per 81-1259, 1981.
[23] P.E. Morgan, M.R. Visbal, Simulation of unsteady three-dimensional separation
on a pitching wing, AIAA Paper 2001-2709, June 2001.
[24] H. Ranke, Unsteady separation in two-dimensional turbulent ows, in: 19th
Congress of the International Council of the Aeronautical Sciences, ICAS Paper
94-4.5.1, Sep. 1994.
[25] S. Subchan, R. Zbikowski, J.R. Cleminson, Optimal trajectory for the terminal
bunt problem: an analysis by the indirect method, AIAA Paper 2003-5791,
2003.
[26] B.E. Tinling, C.Q. Allen, An investigation of the normal-force and vortex-wake
characteristics of an ogive-cylinder, NASA-TND-1297.

You might also like