Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

JOURNAL OF SPACECRAFT AND ROCKETS

Flight Simulation and Drag Prediction for a Pitching-Accelerating


Hypersonic Reentry Vehicle

Mostafa Khalil,∗ Mohamed S. Abdel-Gawad,† and Mahmoud Y. M. Ahmed‡


Military Technical College, Cairo 11766, Egypt
https://doi.org/10.2514/1.A35441
A nonwinged reentry vehicle, such as the upper stage of a ballistic missile, endures harsh aerodynamic loads during
the phase of reentry. This phase is characterized by sophisticated flight phenomena including hypersonic speed,
continuous variation of surrounding atmospheric conditions, and pitching transient vehicle motion. During the flight
in this descent phase, drag is considered the dominant aerodynamic force where accurate prediction is crucial as far as
reliable trajectory prediction and mission planning are concerned. This study aims to investigate drag on a generic bi-
cone nonwinged reentry vehicle resembling an upper stage of a ballistic missile during descent. This is done by using
both trajectory prediction and numerical simulation of vehicle–air interaction. To obtain the vehicle dynamics and
instantaneous flight conditions through descending phase, a three-degree-of-freedom pitch flight trajectory model is
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

utilized starting from point of separation. It was concluded that the vehicle experiences small-amplitude pitching
motion while descending from about 90 to 30 km and accelerating in a hypersonic regime (Mach 5.7–7). Drag is then
calculated via time-dependent numerical simulations. The results show the flowfield pattern evolution in addition to
the temporal variation of drag on the vehicle during the flight phase in concern.

Nomenclature vehicle. While lift plays a minor role for such unwinged nonlifting
CD = drag coefficient vehicle, drag appears to be the dominant aerodynamic force acting
CαL = lift slope coefficients on the vehicle. An accurate estimation of drag on a reentry vehicle
Cqm = change in the pitching moment coefficients with respect should account for a multitude of complexities. The complexities of
to the pitching angular velocity flow around the descending reentry vehicles are owed to the continu-
Cαm = change in the pitching moment coefficients with respect ous and sharp variation in atmospheric conditions. Air pressure and
to the angle of attack density increase while temperature profile varies from one layer to
Dbp = base pressure drag, N another. The flow changes from free molecular, to slip flow, and
eventually to continuum flow depending on Knudsen number [1,2].
Dff = forebody friction drag, N
In addition, the freestream speed increases as the vehicle accelerate
Dfp = forebody pressure drag, N
during descent while performing pitching unsteadiness. Finally, the
g = gravitational acceleration, m∕s2 flow around the vehicle is hypersonic, invoking extreme temper-
Iyy = vehicle mass moment of inertia, kg ⋅ m2 atures and heat flux, strong shock waves, and shock–boundary-layer
L = vehicle length, m interaction.
Lref = vehicle reference length, m The focus of previous studies on hypersonic vehicles is more
m = vehicle mass, kg oriented toward steady flight conditions. For instance, the compre-
q = pitching angular velocity, rad/s hensive database by Weiland [3] includes steady aerodynamic char-
q∞ = freestream dynamic pressure, N∕m2 acteristics of a wide variety of reentry vehicle configurations. On the
Sref = vehicle reference area, m2 other hand, studies dedicated to understanding the transient aerody-
V = vehicle total velocity relative to the air stream, m/s namic features around complete missile and projectile configurations
α = angle of attack, rad at supersonic and low hypersonic speeds are available in the open
γ = flight-path angle, rad literature. In the majority of these studies, the focus was made on one
γ_ = flight path rate, rad/s form of unsteadiness, and numerical simulations were adopted.
θ = pitch angle, rad Numerical simulation tools based on solving Navier–Stokes equa-
θ = pitch angular acceleration, rad∕s2 tions assume that the flow around the vehicle is a continuum flow of
real gas, an assumption that applies for flight altitudes below approx-
imately 90 km in the standard atmosphere [4].
I. Introduction For instance, pure pitching of hypersonic vehicles of various

A N UPPER stage of a ballistic missile, which is designed to


deliver its payloads at long ranges, is a reentry vehicle by
definition. During the descent phase toward a target, this hypersonic
shapes was studied in [5–14], whereas phugoid, flapping, and rolling
unsteadiness were separately examined in [14–16], respectively. In
contrast, the impact of pure longitudinal acceleration on hypersonic
vehicle needs to accurately follow its predefined trajectory using vehicle characteristics in Mach 1–5 freestream was addressed
onboard guidance, navigation, and control (GNC) devices. One numerically by Gledhill et al. [17]. Yawing–spinning motion of
fundamental input to GNC is the aerodynamic characteristics of the hypersonic projectiles with Mach numbers of 2.89 and 4.49 at a
constant altitude was numerically studied by Cayzac et al. [18]. In
Received 5 April 2022; revision received 26 July 2022; accepted for contrast, pitching–yawing for spinning finned projectile was exam-
publication 15 August 2022; published online 7 September 2022. Copyright ined by Sahu [19] at Mach 3 and a spin rate of 2500 rad∕s. Liu et al.
© 2022 by Mahmoud Ahmed, The Military Technical College. Published by [20] investigated numerically the static and dynamic stability deriv-
the American Institute of Aeronautics and Astronautics, Inc., with permission. atives of hypersonic waverider performing both pitching oscillation
All requests for copying and permission to reprint should be submitted to CCC and longitudinal acceleration at constant flight altitude and Mach.
at www.copyright.com; employ the eISSN 1533-6794 to initiate your request. To the authors’ best knowledge, unsteady aerodynamic charac-
See also AIAA Rights and Permissions www.aiaa.org/randp.
*Lecturer of Aerodynamics and Flight Mechanics, Aerospace Engineering
teristics of an unwinged hypersonic vehicle in a real descent
Department. trajectory involving variable freestream conditions, longitudinal

Researcher, Aerospace Engineering Department. acceleration, and pitching were not discussed in the open literature.
‡ This shortage is the main motivation for running research conducted
Associate Professor of Aerodynamics, Aerospace Engineering Depart-
ment; mahmoud.yehia@mtc.edu.eg. Member AIAA. by the authors using numerical simulation taking into consideration
Article in Advance / 1
2 Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED

Z
0.157D
0.03D x

0.007D D
CG

V
q
Z a
1.37D
2.3D
2.5D g
Fig. 1 Configuration and dimensions of the case study model.
O X
all flight complexities stated above. The objective is to understand the
aerodynamics and estimate the drag acting on a hypersonic reentry
vehicle that has the form of a bi-cone representing the upper stage of a Fig. 2 Three-degree-of-freedom flight model in the pitch plane.
long-range ballistic missile. After separation, the vehicle flies freely
performing a combined pitching oscillation and longitudinal accel- 150000 9
eration during descent in a dense atmosphere with flight Mach
numbers varying from 5.7 to 7. The study is conducted by numeri-

Altitude [m]

Mach No.
100000 6
cally simulating the transient continuum flow around the vehicle.
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

Before numerical simulations, vehicle flight performance and the


corresponding freestream conditions are obtained by solving the 50000 3
trajectory problem from the point of separation till the ground impact. Altitude
The remainder of the paper is organized as follows. First, features Mach No.
0 0
of the case-study vehicle are illustrated, and the research methodol- 0 40 80 120 160 200 240 280 320

ogy is explained in the next section. Then, the main findings of the Flight time [s]
study are illustrated and discussed. Finally, the paper wraps up with Fig. 3 Vehicle flight performance from separation to ground impact.
the key conclusions.

-20 1.5
Flight path, Pitch angles [deg]

II. Case Study


-25 1
The hypersonic vehicle in concern is a nonwinged reentry vehicle

Pitch rate [deg/s]


-30 0.5
representing the upper stage of a typical ballistic missile [21]. It has
the form of a bi-cone with a multistep tip (Fig. 1). The carrier missile -35 0
is a multistage ballistic missile that is launched vertically to hit remote -40 -0.5
ground stationary targets. After separation of the booster and sus- -45 -1
tainer stages of the missile, the unpowered upper stage continues its -50 -1.5
flight separately till the ground impact. The main vehicle specifica- Gama Theta Theta_dot
-55 -2
tions and flight conditions at the point of separation are listed in 228 238 248 258 268 278 288 298 308
Table 1. Flight time [s]

Fig. 4 Vehicle flight dynamics through the descent in the dense


atmosphere.
III. Flight Trajectory Simulation
The hypersonic vehicle separates during the missile ascent and
continues a ballistic trajectory up to the summit and then descends   _ 
θd
toward the target. Separation conditions are adopted as initial flight I yy θ  q∞ Sref Lref Cαm α  Cqm (3)
conditions for trajectory prediction. A computer code is developed V
based on the (three-degree-of-freedom pitch) equations to estimate
the full unpowered trajectory of the vehicle from the separation point where V is the vehicle’s total velocity relative to the air stream, and m,
to the ground impact. In the model, as illustrated in Fig. 2, (X, Z) is the I yy , Sref , and Lref are the vehicle mass, mass moment of inertia,
inertial frame of the pitch plane located at the launch site, and (x, z) is reference area, and reference length, respectively. Gravitational
the body frame located at the vehicle center of gravity (CG). The acceleration and freestream dynamic pressure are g and q∞ , respec-
Earth is approximated to be flat and nonrotating. The equations tively. CD and CαL are the drag and lift slope coefficients acting on the
describing the unpowered vehicle motion in pitch plane affected by vehicle, respectively. Vehicle angle of attack, pitch angle, and flight-
aerodynamic drag, lift, pitching moment, and gravity are stated as path angle are α, β, and γ, respectively, whereas its flight path rate and
follows [22]: pitch angular acceleration are γ_ and θ,
 respectively. The change in the
pitching moment coefficient with respect to the pitching velocity, q is
mV_  −mg sin γ − q∞ Sref CD (1) Cqm whereas Cαm is the change in the pitching moment coefficients
with respect to the angle of attack. Aerodynamic coefficients and
mV γ_  −mg cos γ  q∞ Sref CαL α (2) derivatives input to the trajectory model above are calculated using an
engineering technique based on analytical techniques, tabulated data,
and empirical formulas developed by the research group [23] that
Table 1 Case study main specifications were confirmed to yield reliable results [24].
As shown in Fig. 3, the vehicle flight performance including the
Parameter Value flight altitude and velocity as a function of flight time are illustrated.
Fineness ratio 2.5 The flight time zero denotes the instant of the vehicle separation.
Height of separation, km 39.3 According to the flight simulation results, the vehicle takes about 81 s
Flight path angle at separation, deg 40 through the dense atmosphere (85–90 km altitude [25]) till the impact
Location of the center of gravity 1.13
(in calibers, measured from nose tip)
with the ground. The evolution of vehicle pitch and flight path angles
Flight speed at separation, m/s 2020 and the corresponding pitch rate is illustrated in Fig. 4 over the entire
flight time.
Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED 3

2050 4 6.5 640


Velocity AOA
Velocity [m/s]

Mach No.
AOA [deg]
1950 0 6.4 540

P . [Pa]
1850 -4 6.3 440
M P_inf
6.2 340
1750 -8 273 273.5 274 274.5 275 275.5 276
228 233 238 243 248 253 258 263 268 273 278
Flight time [s]
a) Flight time [s]
a) Freestream mach and pressure
1E+7 7.2
2025 1
Reynolds No.

1E+6 6.9

Velocity [m/s]
Mach No.

AOA [deg]
2020 0
1E+5 6.6
1E+4 6.3 2015 -1

1E+3 6 2010 -2
Re M Velocity AOA
1E+2 5.7 2005 -3
228 233 238 243 248 253 258 263 268 273 278
273 273.5 274 274.5 275 275.5 276
b) Flight time [s] Flight time [s]
Flight path, Pitch angles [deg]

-20 1 b) Vehicle velocity and angle of attack (AOA)


-25 0.5 Fig. 6 Temporal variation of vehicle kinematic and freestream condi-

Pitch rate [deg/s]


tions for the main case.
-30 0
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

-35 -0.5
simulated to economize the simulation budget. During the 3 s period
-40 -1 starting at the altitude of 38 km, the vehicle performs one-quarter of
Gama Theta theta_dot
-45 -1.5
the pitching cycle along with longitudinal acceleration. This period is
228 233 238 243 248 253 258 263 268 273 278 marked by the dashed frame in Fig. 5 and better illustrated separately
c) Flight time [s] in Fig. 6. Details of these three cases are listed in Table 3.
Fig. 5 Results of flight trajectory simulation through the concerned It is worth mentioning here that the results of case 1 were briefly
period of study. reported in [26] and are presented here for the sake of completeness.
In addition to these three transient simulation cases, a set of steady
simulations (case 4 to case 12) are planned for the sake of comparison
The dashed frame in Fig. 3 highlights the period of the trajectory of and analysis.
interest in the present study. This period is characterized by the Table 4 explains the setup for steady simulation cases. In all 12
accelerating descent in the dense atmosphere; i.e., the vehicle veloc- cases, operating conditions imply continuum flow as the maximum
ity increases while flying toward the ground. This period spans 51 s value of the Knudsen number does not exceed 0.01 (case 4).
starting from the altitude of 87 km to the altitude of 29.4 km. During
the flight in the period in concern, the temporal variation in vehicle B. Computational Domain Definition
kinematics, freestream dimensionless parameters, and attitude analy-
sis are shown in Figs. 5a–5c, respectively. Trajectory simulation For zero-incidence cases, an axisymmetric 2D computational do-
results are used as inputs for the following computational fluid main is constructed. Domain configuration, extents, and boundary
dynamics (CFD) simulations. definitions are illustrated in Fig. 7a. The domain is divided into nine
blocks at each discontinuity on the vehicle contour to generate a
structured grid. The discretized domain in the vicinity of the vehicle is
shown in Fig. 7b.
IV. Setup for Numerical Simulations and Drag
The spatial resolution quality is assessed through grid sensitivity
Calculation analysis. Four grids with different resolutions are generated to sim-
A. Design of Numerical Simulation Cases ulate the flow around the vehicle at a representative case, namely,
As inferred from the trajectory results above, the unsteadiness Mach 6 freestream flow at standard sea-level conditions. The total
encountered by the vehicle is a combined longitudinal acceleration drag coefficient is taken as a measure of solution quality; the results
and pitching while simultaneously descending. Based on Fig. 4, are shown in Table 5. Based on these results, grid 3 with 91,100 cells
features of the vehicle flight during the descent phase in concern is adopted.
can be summarized in Table 2. For nonzero incidence cases, a half-full three-dimensional com-
Numerical simulation is planned to understand the aerodynamic putational domain is constructed of two interfaced semispherical
behavior of the vehicle under acceleration, and pitching while subdomains inside each other such that their centers coincide at the
descending. For the sake of a better understanding of physics, two vehicle gravity center. Diameters of the inner and outer domains
simpler subcases are planned. In the first subcase (case 1), the vehicle are 2.6 and 6.5 times the vehicle length, respectively. The interface
is assumed to perform pure longitudinal acceleration while descend- between the two domains permits the sliding of the inner grid with
ing for the entire period in concern (51 s). In the other subcase respect to the outer one to perform the pitching motion. Figure 8a
(case 2), the vehicle is assumed to perform pure pitching while flying shows the shape and boundary conditions for the computational
in cruise conditions (i.e., constant altitude and speed) for the whole domain. The inner domain is divided into 18 blocks at each dis-
pitching period of 17 s. For the main case in which a combined continuity on the vehicle contour, whereas the outer domain is
acceleration–pitching is performed (case 3), a representative period is divided into 5 blocks to generate a structured grid. First cell heights
and cell counts normal to the solid wall are taken identical to those in
Table 2 Key flight features of the vehicle grid 3 used in the 2D computational domain, the quality of which is
for the interval in concern assured through grid sensitivity checks above. Eventually, inner and
outer domains included 2.85 and 0.77 million cells, respectively.
Flight feature Value Figure 8b shows the discretized domain as well as a zoom-in on the
Average longitudinal acceleration 5 m∕s2 vehicle.
Average pitching period 17 s
Pitching amplitude 3° C. Solver Setup and Boundary Conditions
Time of pitching onset from separation 262 s
Altitude of pitching onset 87 km A commercial CFD solver is used [27], in which the transient
density-based, implicit, double-precision second-order accurate
4 Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED

Table 3 Design of transient numerical simulation cases


Operating conditions
Case Case definition Physical time, s Mach No. Reynolds No. ×10−6 Altitude, km Incidence angle, deg
1 Pure linear acceleration while descent 51 5.7–7 0.55–1.6 87–30 0
2 Pure pitching at cruise conditions 17 6 0.03 60 −3 to 3
3 Combined acceleration–pitching while descent 3 6.2–6.5 4.7–8.7 38–34 0–3

Table 4 Design of steady numerical simulation cases Table 5 Grids characteristics for grid independence check
Operating conditions Grid ID No. of cells First cell height, mm CD % Change
Mach Reynolds Altitude, Incidence angle, 1 2,900 23 0.1384 ——
Case No. No. ×10−6 km deg 2 13,200 0.5 0.1236 10.69
4 6.5 0.01 70 0 3 91,100 0.15 0.1226 0.81
5 6 0.03 60 4 360,200 0.007 0.1225 0.08
6 5.7 0.07 52
7 6.5 1.02 33
8 0
D. Time-Step Sensitivity Analysis
9 1
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

10 6 0.03 60 2 The temporal resolution of the transient simulations is decided


11 3 based on time sensitivity checks in which acceleration and pitching
12 4 are treated separately since each has a separate UDF. For the pure
acceleration case (case 1), three time-step sizes, namely, 0.1, 0.01,
and 0.001, of a second are examined, and temporal variation of
technique is implemented. Air is treated as an ideal gas, while the vehicle total drag coefficient is adopted as a measure of solution
quality. Based on the results shown in Fig. 9a, a time-step size smaller
laminar–turbulent transitional flow is handled using the intermittency
than 0.01 s yields an insignificant effect on the solution quality. So, a
transition shear-stress transport (SST) k-ω turbulence model avail-
time step of 0.01 s is utilized through the simulation of case 1.
able in the solver. The vehicle surface is treated as an adiabatic no-slip
For the pure pitching case (case 2), four different time-step sizes,
wall. At the inlet boundary, freestream Mach and static pressure and
namely, 0.1, 0.04, 0.02, and 0.01, of the pitching period (17 s) are
temperature are defined. At the exit, the freestream static pressure and examined. Figure 9b shows the results of time-step sensitivity analy-
total temperature are specified. The boundary conditions for the sis for pure pitching based on the proposed approach. A time-step size
freestream flow vary from one case to the other according to Tables 3 of 0.01 of the period (i.e., 0.17 s) is chosen for case 2. For the main
and 4. For all steady cases, freestream conditions are directly defined, case (case 3), the smaller time-step size, namely, 0.01 s is adopted for
whereas for transient cases, user-defined functions are used. case 3 simulation.
For case 1, a user-defined function (UDF) is developed to account
for the temporal variation in atmospheric conditions and Mach
number during accelerating descent based on trajectory results
(Fig. 5). For case 2, another UDF is developed to account for the V. Results and Discussion
pitching motion of the inner domain with respect to the fixed outer A. Validation of the Solver
one. This is sought to represent the real motion of the vehicle with Wide-ranged steady aerodynamic characteristics experimental
respect to the incoming flow and thus express more accurately the database of various reentry vehicle configurations was provided by
associated physics. For the main case, case 3, the two UDFs are Weiland [3]. To conduct the validation of the CFD solver, a case of a
implemented to simultaneously account for longitudinal accelera- slender bi-cone with a hemispherical tip configuration is selected.
tion, pitching, and descent. Figure 10 shows the configuration of the model. The computational
domain and solver setup have the same aspects explained previously.
Three simulation cases are undertaken at three different conditions to
compare with the available wind tunnel results.
Inlet (pressure far-field)
Table 6 holds a comparison between the experimental and calcu-
Outlet (pressure outlet)
lated drag coefficients at the given values of Mach number and
Reynolds number (referred to D  4.159 m). The small deviation
is taken as an indication of solver validity in solving the flow problem
2.5 L
Model (wall)
in concern.
Axis
B. Results for Case 1: Pure Longitudinal Acceleration While Descent
at Zero Incidence
The flowfield features around the vehicle at a representative time
0.2 L L 5L
instant, at 166 s from separation, are illustrated in Fig. 11. An over-
a) Shape, extents, and boundary definitions, L is the model length view of the domain is shown as pressure and Mach contours to the left
and right, respectively (Fig. 11a). A close-up view of Mach contours
around the model is shown in Fig. 11b, while a zoom-in view of the
vehicle’s nose tip is shown in Fig. 11c with Mach and pressure
contours at the top and bottom halves of the view, respectively. The
shock wave generated ahead of the body is oblique and straight with a
small curvature at the tip. Since the nose tip is rounded, the shock
wave is detached at a small distance ahead of the tip. A weak
expansion wave can be shown at the shoulder of the first conical
section downstream of the tip. In addition, the thick boundary layer
b) Mesh of the discretized domain characterizing the hypersonic flow is evident. As the boundary layer
Fig. 7 Features of the axisymmetric computational domain. separates at the vehicle base, a shear layer is created that propagates
Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED 5

Pressure far-field

Interface
Pressure outlet

Wall (model)

Outer domain

Inner domain

a) Shape, extents, and boundary definitions


Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

b) Discretized domain features


Fig. 8 Features of the axisymmetric computational domain.

0.3104 downstream and engulfs a large zone of recirculating air at the base of
Drag coefficient

0.3102 the vehicle.


0.3100 The evolution of drag and its coefficient as the vehicle descends is
0.3098 Δt=(1/10) sec
shown in Fig. 12. Drag on the vehicle slightly increases with time as
0.3096 Δt=(1/100) sec
Δt=(1/1000) sec the vehicle descends. Below about 50 km of altitude, drag rise
0.3094
1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 becomes clearly significant as the rate of drag rise attains increasingly
Simulation time [s] high values. This is definitely owed to the continuous rise in free-
a) Pure acceleration case stream dynamic pressure and the subsequent increase in surface
pressure. Meanwhile, the freestream density and, more significantly,
0.08 T/10 velocity increase with rates higher than that of the drag as the vehicle
T/25
AOA [rad]

0.04 T/50 descends. Eventually, the drag coefficient decreases as the vehicle
0 descends at a steeper rate than that typical for this range of Mach
-0.04 numbers (due to the role of density rise).
-0.08 To analyze the drag behavior more deeply, drag components,
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
Simulation time [s] namely, forebody pressure drag (Dfp ), forebody friction drag (Dff ),
b) Pure pitching case and base pressure drag (Dbp ) are examined at four different time
Fig. 9 Time-step sensitivity analysis results.
instances during descent. These components are extracted from
transient simulation results and are listed in Table 7. For the sake of
comparison, the corresponding components based on steady simu-
z lations (cases 4–7, respectively) are listed as well.
2D
1.65D

8° Table 6 Validation case results


20°
D Model drag coefficient
x Mach No. Reynolds No. ×10−6 CFD Experimental [3] Deviation %
0.2D 2.53 249 0.352 0.34 3.5
4.0 393.8 0.341 0.35 2.5
5.96 586.8 0.224 0.23 2.7
Fig. 10 Configuration of the validation case.
6 Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED

Shear layer

Recirculating air zone


Wake flow

a) Domain full view, pressure contours (left), mach contours (right)

Shock wave

Zone-A
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

Boundary layer

b) Mach contours, close-up view

Shock stand-off
Zone-B

Mac h Number Expansion wave


1.35e-4 0.854 1.71 2.55 3.42 4.27 5.13 5.98 6.83 7.69 8.54

c) Zone-A (left), zone-B (right), mach contours (top), pressure contours (bottom)
Fig. 11 Flowfield features at 166 s from separation.

0.9 2800 the dominant component that represents about 32% of the total
Cd Drag [N] drag at the altitude of 77.8 km and increases to reach 65% of the
Drag coefficient

0.7 2100
total drag value as the vehicle reaches the height of 32.5 km.
Drag [N]

0.5 1400 Similarly, the pressure drag on the base (Dbp ) increases as the
vehicle descends to form about 25% of the total drag value. On
0.3 700 the other hand, the forebody friction drag (Dff ) forms about 60%
of the total drag at the altitude of 77.8 km and reach 11% of the
0.1 0
0 4 8 12 16 20 24 28 32 36 40 44 48 52 total drag value when the vehicle is 32.5 km above the ground.
Flight time from separation [s] Compared with steady calculations, the instantaneous drag val-
Fig. 12 Temporal variation of drag and drag coefficient during accel- ues are slightly higher (2–10%) than their steady counterparts at
erating descent. the same freestream conditions. This difference may be explai-
ned by the kinematics of the vehicle during acceleration. The
slight difference may be owed to the relatively small value of
As inferred from Table 7, the rise in drag during the longitudi- longitudinal acceleration, which may make the transient motion
nal acceleration is a general trend for all components along with of the accelerating vehicle a close similarity to a quasi-steady
the whole interval. The pressure drag on the forebody (Dfp ) is one.

Table 7 Evolution of drag components during descent

Dfp , N Dff , N Dbp , N Total drag, N


Flight time from separation, s Transient Steady Transient Steady Transient Steady Transient Steady
166 2.06 2.57 4.38 4.59 0.57 0.55 7.02 7.715
176 8.33 7.99 8.38 7.66 2.49 2.21 19.19 17.86
186 32.46 31.35 16.2 14.42 12.43 11.47 61.1 57.24
196 65.35 63.55 23.14 19.73 28.05 22.93 116.54 106.22
206 1006.72 971.1 163.34 176.7 381.75 359.71 1551.81 1507.5
Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED 7

C. Results for Case 2: Pure Pitching at Cruise Conditions the conical section shoulder are stronger on the windward side than
Figure 13 shows the evolution of key flowfield features, Mach and on the leeward side.
pressure contours, at four different positions of the cycle, namely, The evolution of vehicle drag coefficient during two complete
quarter, half, three-quarters, and full pitching period. Mach contours cycles of a regular pitching oscillation with a period of 17 s is
show the structure of the boundary layer around the vehicle while illustrated in Fig. 14. The dashed line represents the variation of
pitching up. incidence angle during the cycle, whereas the markers indicate the
The boundary layer is thinner on the upper meridian of the vehicle static values of vehicle drag coefficient at the respective instanta-
while pitching down and thicker in case of pitching up. It is noticed neous incidence angles (cases 8–12).
that the boundary layer is always attached and does not separate on The drag performs two cycles per pitching cycle because of the
the leeward side during pitching. This may be owed to the low symmetry of the vehicle. Starting at zero incidence, drag attains a
pitching amplitude and the establishment of a laminar boundary layer minimum value that increases rapidly as the vehicle pitches down,
along the vehicle surface. For pressure contours, the flowfield pres- reaching a maximum value at the lower limit of incidence. During
sure distribution around the vehicle at different states during the cycle pitching up, drag drops to a minimum that is slightly higher than that
is also shown. Pressure contours around the vehicle tip indicate the at the start. A clear drop in the maximum drag at the end of pitching-
change in shock wave location and strength on both sides of the up is also evident. This drop may be explained by the fact that the
vehicle at incidence. Both the shock wave and the expansion fan at peak positive value of incidence is less than the negative one since the
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

Mach Number Mach Number

0.00 0.89 1.78 2.68 3.57 4.46 5.35 6.24 7.14 0.00 0.89 1.78 2.68 3.57 4.46 5.35 6.24 7.14

Pressure [Pa] Pressure [Pa]


0.69 109.74 218.79 327.84 436.89 545.94 654.99 764.04 873.09 0.29 109.35 218.41 327.46 436.52 545.57 654.63 763.68 872.74

a) at t = T/4 c) at t = 3T/4

Mach Number Mach Number


0.00 0.89 1.78 2.68 3.57 4.46 5.35 6.24 7.14 0.00 0.89 1.78 2.68 3.57 4.46 5.35 6.24 7.14

Pressure [Pa] Pressure [Pa]


0.39 108.81 217.22 325.64 434.05 542.46 650.88 759.29 867.70 0.39 108.81 217.22 325.64 434.05 542.46 650.88 759.29 867.70

b) at t = T/2 d) at t=T
Fig. 13 Evolution of Mach contours around the whole vehicle (top), and pressure contours at the nose tip (bottom) during a complete pitching cycle.
8 Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED

0.17 5 1400
4 Drag (acc.)
0.165 3 1300
Drag (pitch-acc.)
Drag Coeff.

2 1200

AOA [deg]
0.16 1

Drag [N]
0 1100
0.155 -1 1000
-2
0.15 -3 900
-4
800
0.145 -5
262 267 272 277 282 287 292 297 700
Flight time [s] 273 273.5 274 274.5 275 275.5 276
Steady drag Unsteady drag AOA Flight time [s]

Fig. 14 Temporal evolution of drag coefficient during two complete Fig. 16 Evolution of drag on under pitch-accelerating descent.
pitching cycles.
D. Results for Case 3: Combined Acceleration–Pitching While
Descent
0.1
The variation of drag acting on the vehicle as it performs pitching
0.09 during accelerating descent for the 3 s in concern is addressed in
0.08 Fig. 16. For the sake of comparison, the variation of drag during the
Drag coefficient

0.07
same period without pitching is plotted with a dashed line.
As expected, drag acting on the vehicle increases as it descends
0.06 during the period in concern. Pitching has the role of slightly reducing
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

0.05 the instantaneous drag at zero incidence and increasing it at the peak
0.04
pitch-up angle. Overall, it can be inferred that the pitching motion has
the impact of increasing the rate by which drag on the vehicle
0.03
increases with time. This trend can be viewed as a superposition of
0.02 two factors; namely, drag rises as the freestream total pressure
T/4, -4 deg T/2, 0 deg 3T/4, 4 deg T, 0 deg
increases, and drag rises as the vehicle pitches up from 0 to 4 deg.
Time in periods, incidence angle
FP-U FF-U BP-U FP-S FF-S BP-S
VI. Conclusions
Fig. 15 Drag components variation during one complete cycle.
The features and evolution of the flowfield around a hypersonic
reentry vehicle during the descent phase were presented in this study.
The vehicle, which is the upper stage of a ballistic missile, experi-
pitching cycle is not symmetric as indicated in Fig. 5a. This drag peak ences both longitudinal acceleration and pitching while descending
is higher than the static drag value at the same incidence angle. With toward the ground. The trajectory from the separation point is esti-
respect to vehicle kinematics, a minor lag of about 0.17 s (1% of the mated using a trajectory prediction code. The flow around the vehicle
cycle period) is addressed in the drag values of the pitching vehicle in is simulated using a commercial CFD tool in three cases: pure
this cycle. This is possibly due to the low amplitude and frequency of longitudinal acceleration during descent, pure pitching at cruise
pitching. conditions, and combined pitching–accelerating while descent. Sim-
To explore the drag components acting on the vehicle and their ulations account for the variation of freestream properties during
individual evolution with time during this cycle of pitching, Fig. 15 descent in a continuum standard atmosphere. The study findings
shows the variation at four different instances during one cycle of can be briefed in that, for the case in concern, as the vehicle descends,
drag components, namely, forebody pressure (FP), forebody friction drag increases almost exponentially with time as a result of the rise in
(FF), and base pressure (BP) based on transient simulations (U). freestream total pressure while the drag coefficient, on the other hand,
Corresponding steady drag components values (S) are also shown for decreases monotonically as the vehicle descends. In addition, upon
comparison. comparing the steady and transient simulation results, the phase shift
The three components of drag are minimum when the vehicle and results difference, which is counted for the transient cases, is
is instantaneously at zero incidence and maximum at the peak inci- owed to the kinematics of the vehicle itself either at the longitudinal
dence angle. Clearly, the dominant (and most incidence-sensitive) acceleration or at pitching oscillation unsteadiness.
component is the forebody pressure drag while base pressure drag is For the case in concern, since the kinematics of the vehicle
the minor contributor and the least sensitive to incidence. In addition, (acceleration and pitching frequency) are small, only marginal
drag on the vehicle base is almost unresponsive to the attitude of the differences can be addressed between steady and transient simula-
vehicle during pitching because of the low pressure. The base drag tions. Also, the pressure drag force on the forebody of the vehicle is
the dominant component of the acting drag force, while the total
variation during pitching is about 6 drag counts (a drag count 
drag force dramatically increases while the freestream is completely
0.0001), which signifies less than 2% of its average value. Since its
turbulent.
transient and steady values are almost equal, it can be treated as
Simulation of the entire descent phase considering both longi-
insensitive to the motion of the vehicle. On the other hand, the
tudinal acceleration and lateral oscillation is considered for future
forebody surface friction drag shows moderate variation during work. Aerodynamic heating effects on the vehicle during descent can
pitching; it is higher at incidence than is it at no incidence. The also be addressed.
variation of friction drag reaches about 19 drag counts that represent
5% of its average value. Similar to base drag, friction drag on the
forebody is less delicate to vehicle kinematics since its transient and References
static values are almost equal (about two drag counts average differ- [1] Anderson, J. D., Jr., Fundamentals of Aerodynamics, McGraw–Hill
ence). Finally, the vehicle forebody pressure drag is the most sen- Education, New York, 2010, p. 26.
sitive component to variation in vehicle state during pitching. The [2] Shen, C., Rarefied Gas Dynamics: Fundamentals, Simulations and
variation of this component reaches within 138 drag counts, repre- Micro Flows, Springer Science and Business Media, New York, 2006,
Chap. 1.
senting about 16% of its average value. The marginal difference
[3] Weiland, C., Aerodynamic Data of Space Vehicles, Springer Science
between static and transient values of forebody pressure drag may and Business Media, New York, 2014, pp. 132–137.
be owed to vehicle dynamics. It can be also inferred that the drop in [4] Celenligil, M. C., Moss, J. N., and Bird, G. A., “Direct Simulation of
total drag amplitude with oscillation is primarily owed to pressure Three-Dimensional Flow About the AFE Vehicle at High Altitudes,”
drag on the vehicle forebody. NASA TM 101492, 1989.
Article in Advance / KHALIL, ABDEL-GAWAD, AND AHMED 9

[5] Hemdan, H. T., “Similarity Solutions for Oscillating Pointed-Nose Science China Physics, Mechanics & Astronomy, Vol. 57, No. 12, 2014,
Slender Axisymmetric Bodies—Part II: Curved Bodies,” Acta Astro- pp. 2194–2204.
nautica, Vol. 49, No. 11, 2001, pp. 611–626. https://doi.org/10.1007/s11433-014-5603-1
https://doi.org/10.1016/S0094-5765(01)00084-4 [17] Gledhill, I. M. A., Forsberg, K., Eliasson, P., Baloyi, J., and Nordström,
[6] Walchner, O., and Clay, J., “Hypersonic Stability Derivatives of Blunted J., “Investigation of Acceleration Effects on Missile Aerodynamics
Slender Cones,” AIAA Journal, Vol. 3, No. 4, 1965, pp. 752–754. Using Computational Fluid Dynamics,” Aerospace Science and Tech-
https://doi.org/10.2514/3.2966 nology, Vol. 13, Nos. 4–5, 2009, pp. 197–203.
[7] Orlik-Rückemann, K. J., “Stability Derivatives of Sharp Wedges in https://doi.org/10.1016/j.ast.2009.04.008
Viscous Hypersonic Flow,” AIAA Journal, Vol. 4, No. 6, 1966, [18] Cayzac, R., Carette, E., and Thepot, R., “Recent Computations and
pp. 1001–1007. Validations of Projectile Unsteady Aerodynamics,” Proceedings of
https://doi.org/10.2514/3.3594 22nd International Symposium on Ballistics, Nov. 2005.
[8] Orlik-Rückemann, K. J., “Oscillating Slender Cone in Viscous Hyper- [19] Sahu, J., “Time-Accurate Numerical Prediction of Free-Flight Aerody-
sonic Flow,” AIAA Journal, Vol. 10, No. 9, 1972, pp. 1139–1140. namics of a Finned Projectile,” Journal of Spacecraft and Rockets,
https://doi.org/10.2514/3.50335 Vol. 45, No. 5, 2008, pp. 946–954.
[9] Lyons, W., Brady, J., and Levensteins, Z., “Hypersonic Drag, Stability, https://doi.org/10.2514/1.34723
and Wake Data for Cones and Spheres,” AIAA Journal, Vol. 2, No. 11, [20] Liu, X., Liu, W., and Zhao, Y., “Navier–Stokes Predictions of Dynamic
1964, pp. 1948–1955. Stability Derivatives for Air-Breathing Hypersonic Vehicle,” Acta
https://doi.org/10.2514/3.2704 Astronautica, Vol. 118, Jan.–Feb. 2016, pp. 262–285.
[10] Ericsson, L. E., “Universal Scaling Laws for Nose Bluntness Effects on https://doi.org/10.1016/j.actaastro.2015.10.015
Hypersonic Unsteady Aerodynamics,” AIAA Journal, Vol. 7, No. 12, [21] “DSSC Report,” Defense Systems Studies Centre, DSSC-TR-2025B,
1969, pp. 2222–2227. Cairo, Egypt, 2015.
https://doi.org/10.2514/3.5519 [22] Chin, S. S., Missile Configuration Design, McGraw–Hill, New York, 1961,
[11] Adams, J., and Griffith, B., “Hypersonic Viscous Static Stability of a
Downloaded by Mahmoud Ahmed on September 16, 2022 | http://arc.aiaa.org | DOI: 10.2514/1.A35441

pp. 94–95.
Sharp 5-Deg Cone at Incidence,” AIAA Journal, Vol. 14, No. 8, 1976, [23] Abdel-Hamid, O. E., “MAC User Guide,” Generalized MAC edn.,
pp. 1062–1068, Technical Research Centre, EAF-TRC, Cairo, Egypt, 2004.
https://doi.org/10.2514/3.7186 [24] El-Mahdy, L. A., Ahmed, M. Y. M., Mahmoud, O. K., and Abdel-
[12] Khalid, M., and East, R., “High Mach Number Dynamic Stability of Hamid, O. E., “A Comparative Study of Prediction Techniques for
Pointed Cones at Small Angles of Attack,” AIAA Journal, Vol. 18, Supersonic Missile Aerodynamic Coefficients,” Journal of Mechanical
No. 10, 1980, pp. 1263–1265. Engineering, Vol. 14, No. 1, 2017, pp. 35–60, https://ir.uitm.edu.my/id/
https://doi.org/10.2514/3.7719 eprint/17466.
[13] Lijun, X., Yunjun, Y., Zhou, L., and Weijiang, Z., “High-Performance [25] Atmosphere, U. S., National Oceanic and Atmospheric Administration,
Computing of Periodic Unsteady Flow Based on Time Spectral Method,” NASA, United States Air Force, Washington, D.C., 1976, Table I.
Procedia Engineering, Vol. 99, Feb. 2015, pp. 1526–1530. [26] Abdel-Gawad, M. S., Khalil, M. S., and Ahmed, M. Y. M., “Unsteady
https://doi.org/10.1016/j.proeng.2014.12.704 Numerical Analysis of Drag on Non-Winged Hypersonic Vehicle
[14] Chou, Y., and Laitone, E. V., “Phugoid Oscillations at Hypersonic During Re-Entry,” IOP Conference Series: Materials Science and
Speeds,” AIAA Journal, Vol. 3, No. 4, 1965, pp. 732–735. Engineering, Vol. 610, Nov. 2019, Paper 012096.
https://doi.org/10.2514/3.2956 https://doi.org/10.1088/1757-899X/610/1/012096
[15] Hui, W., and Tobak, M., “Unsteady Newton-Busemann Flow Theory. [27] FLUENT14.5.7 User’s Guide, ANSYS, Inc., Lebanon, NH, 2011.
I-Airfoils,” AIAA Journal, Vol. 19, No. 3, 1981, pp. 311–318.
https://doi.org/10.2514/3.7773.
[16] Ye, Y. D., Zhao, Z. L., Tian, H., and Zhang, X. F., “The Stability S. Silton
Analysis of Rolling Motion of Hypersonic Vehicles and Its Validations,” Associate Editor

You might also like