You are on page 1of 22

AIAA Atmospheric Flight Mechanics Conference AIAA 2010-8248

2 - 5 August 2010, Toronto, Ontario Canada

TIME-ACCURATE COMPUTATIONS FOR RAPID


GENERATION OF MISSILE AERODYNAMICS

Jubaraj Sahu * and Karen R. Heavey‡


U.S. Army Research Laboratory, Aberdeen Proving Ground, MD 21005-5066

This paper describes a computational research study undertaken to develop new


alternate computational fluid dynamics methodology for determination of aerodynamic
coefficients accurately and efficiently. A new time-accurate sweep procedure has been used
to compute the aerodynamics of complex projectile configurations. Given a steady-state
solution, this technique can be used to generate multiple time-accurate numerical solutions
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

over a select range of angles of attack or side-slip angles for a given set of flight parameters.
The sweep procedure has been used to compute solutions for an elliptic missile at a
supersonic speed and also a low transonic speed, for angle of attack and side-slip sweeps.
Separate steady-state solutions for various angles of attack and side-slip angles were also
computed for comparison. Finite-volume unstructured Navier-Stokes computational
techniques are used, including a two-equation Reynolds-Averaged Navier-Stokes (RANS)
turbulence model. Visualization of the numerical solutions provides the qualitative features
of the flow field with the projectile at various angles of incidence. Aerodynamic coefficients
were extracted from the computed solutions, with good agreement between the steady-state
and sweep solutions. The results of this study show that the time-accurate sweep procedure
can provide aerodynamic coefficients efficiently and accurately.

I. Introduction

A ccurate understanding of aerodynamics and fundamental flow physics is critical to the development of new
advanced low-cost precision munitions. The flow fields associated with these munitions can be very complex
and involve shock-boundary-layer interactions, unsteady wakes, body and wing vortices interactions.1-6 Simple
analytical techniques are not able to determine these complex three-dimensional flow interactions and associated
non-linear flow physics. Although time-consuming, computational fluid dynamics (CFD), with the advancement in
high performance computing (HPC), offers an alternate capability to compute these non-linear interacting flow
fields and the potential to provide detailed understanding of the associated nonlinear aerodynamics processes. The
advancement in CFD has had a major impact on projectile design and development 7-10 and has continued to emerge
as a critical technology for the aerodynamic design and assessment of flight vehicles. It is essential that CFD be
brought into the earlier stages of design and development to be more effective. Also, the designer needs the CFD
results in terms of the flow field and aerodynamic forces and moments to be obtained as fast and as accurately as
possible.
Two types of CFD methods, steady-state and time-accurate, are used to obtain the desired aerodynamics from
subsonic to supersonic speeds. In general, the aerodynamic coefficients at supersonic speeds can be generated
efficiently using steady-state CFD methods. Time-accurate techniques, however, are required in some cases
especially at transonic and subsonic speeds11 and for prediction of base flows and dynamic derivatives. Time-
accurate or unsteady CFD modeling techniques have proven more challenging but are increasingly being used for
numerical prediction of projectile and missile aerodynamic.11-16 Computational algorithm and computing advances
have also led to coupling of CFD codes to projectile body dynamics codes for time-accurate simulation of projectile
free flight motion. This technique allows determination of both static and dynamics derivatives using the same
numerical simulation. In general, such advanced CFD computations can be time-intensive since they are based on

*
Aerospace Engineer, AIAA Fellow

Mathematician, Non-member.

This work is declared work of the U.S. Government and is not subject to copyright protection in the United States.

1
American Institute of Aeronautics and Astronautics

This material is declared a work of the U.S. Government and is not subject to copyright protection in the United States.
solving the time-dependent Reynolds averaged Navier-Stokes (RANS) equations with very small time-steps. The
traditional steady-state solutions are also obtained using the same governing Navier-Stokes equations. Starting with
initial free stream conditions, solutions are marched in time until they are converged. However, a large number of
steady-state solutions is required to generate a complete aerodynamic database. This may include a range of angles
of attack and a range of side slip angles for a given Mach number. Generating the complete aerodynamic database
this way is not trivial and can be very time-intensive as well. Current efforts are therefore directed at developing
and applying alternate procedures for rapid generation of aerodynamics for both simple and complex configurations
at all speeds from subsonic to supersonic. In the present work, research was focused on the development of alternate
methodologies for rapid determination of static aerodynamic coefficients using time-accurate Navier-Stokes
techniques.
The research work presented in this paper is particularly focused on a new time-accurate sweep procedure that
allows determination of aerodynamic coefficients for an entire range of angles of attack quickly. This technique is
demonstrated on a complex non-axisymmetric missile configuration with an elliptic cross-section and a set of
multiple complex fins. Numerical computations for this complex configuration have been performed at a supersonic
speed, Mach 2.5 and a low transonic speed, Mach 0.9 for several angles of attack from -5° to 30° and side-slip
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

angles from -10° to 10° using the time-accurate sweep technique. These computed results are compared with
steady-state results obtained from a previous study17 for the same configuration at M = 2.5. In addition, we have
performed steady-state computations at M = 0.9 for comparison with the time-accurate sweep results. Computed
results are also compared with experimental data obtained at NASA Langley Research Center for the same
configuration and at the same velocities.18
A description of the computational techniques including the new time-accurate sweep procedure is presented in
the next section. Computational results are then presented and validated with data. Computational timing and
resources required show how rapidly the aerodynamic coefficient results are generated using the time-accurate
sweep procedure.

II. SOLUTION TECHNIQUES

Traditionally, aerodynamic information for projectiles has been obtained by computing the steady-state
coefficients at multiple individual angles of attack. This method requires repetitive editing of files and processing of
computer runs. The sweep procedure provides a method to consolidate the process.

a. Steady-State Computational Technique

A commercially available code, CFD++19,20, is used for the time-accurate CFD simulations. The basic numerical
framework in the flow solver contains unified-grid, unified-physics, and unified-computing features. Details of the
basic numerical framework can be found in References 19 and 20. Here, only a brief synopsis of this framework
and methodology is given.
The 3-D, time-dependent Reynolds-averaged Navier-Stokes (RANS) equations are solved using the following
finite volume method:


WdV   F  G dA   HdV
t V
(1)
V

where W is the vector of conservative variables, F and G are the inviscid and viscous flux vectors, respectively, H is
the vector of source terms, V is the cell volume, and A is the surface area of the cell face.
The numerical framework of CFD++ is based on the following general elements: (1) unsteady compressible and
incompressible Navier-Stokes equations with turbulence modeling [unified-physics], (2) unification of Cartesian,
structured curvilinear, and unstructured grids, including hybrids [unified-grid], (3) unification of treatment of
various cell shapes including hexahedral, tetrahedral and triangular prism cells (3-D), quadrilateral and triangular
cells (2-D) and linear elements (1-D) [unified-grid], (4) treatment of multiblock patched aligned (nodally
connected), patched-nonaligned and overset grids [unified-grid], (5) total variation diminishing discretization based
on a new multi-dimensional interpolation framework, (6) Riemann solvers to provide proper signal propagation
physics, including versions for preconditioned forms of the governing equations, (7) consistent and accurate
discretization of viscous terms using the same multi-dimensional polynomial framework, (8) pointwise turbulence
models that do not require knowledge of distance to walls, (9) versatile boundary condition implementation

2
American Institute of Aeronautics and Astronautics
includes a rich variety of integrated boundary condition types for the various sets of equations, (10) implementation
on massively parallel computers based on the distributed-memory message-passing model using native message-
passing libraries or MPI, PVM, etc. [unified-computing].
For supersonic and transonic flows of interest here, several techniques are used to achieve faster convergence. It
combines the basic ideas of implicit local time-stepping and relaxation. Use of an implicit scheme circumvents the
stringent stability limits suffered by their explicit counterparts, and successive relaxation allows update of cells as
information becomes available and thus aids convergence. Second-order discretization was used for the flow
variables and the turbulent viscosity equation. The turbulence closure is based on topology-parameter-free
formulations. Two-equation turbulence models21,22 were used for the computation of turbulent flows. These models
are ideally suited to unstructured book-keeping and massively parallel processing due to their independence from
constraints related to the placement of boundaries and/or zonal interfaces.
The traditional steady-state solutions are obtained using the same governing Navier-Stokes equations. Starting
with an initial free stream conditions, solutions are marched in time until they converge. This process usually
requires anywhere from 500 to a few thousand time-steps depending on the flow conditions.
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

b. Time-Accurate Sweep Procedure

A new time-accurate sweep procedure has been developed for rapid generation of static aerodynamic
coefficients across a range of angles of attack for a fixed side slip angle, for example. In this procedure, the first
step involves generating a steady-state solution at the lowest angle of attack in the desired sweep. This steady-state
solution serves as the initial condition for the angle of attack sweep which is done in the time-accurate mode. Dual-
time-stepping as described below is used to in the time-accurate sweep procedure. Starting with the steady state
solution, the first time-step is run in the unsteady (time-accurate) mode at the starting angle of attack.
Subsequently, at every time step after the first one, the computational grid is rotated by a small increment (on the
order of 0.5°) across the range of angles of attack and computed solutions are obtained in the unsteady mode
basically at the selected increment in angle of attack for the entire range of angle of attack. Figure 1 shows the
location of the projectile at various stages of the sweep. In this example, the rotation point is at the end of the
projectile; however, it can also be conveniently chosen to be the center of gravity of the projectile as well. Here, the
idea is to generate a number of steady-state solutions rapidly, using the solution generated for each angle as an initial
guess for the next time step (angle). The rotation angle is chosen small enough to capture any non-linear
aerodynamic effect that may be present.

Alpha Sweep

Figure 1. Location of the projectile at various stages of the sweep procedure.

Generally, dual time-stepping method is used to achieve the desired time-accuracy between the time-steps for
time-accurate simulations of missile aerodynamics. The term “dual-time-step” implies the use of two time steps.
The first one is an “outer” or global (and physical) time step that corresponds to the time discretization of the
physical time variation term. This time step can be chosen and is typically set to a smaller value to provide the

3
American Institute of Aeronautics and Astronautics
desired time-accuracy. This time step is applied to every cell (not separately varying). In the present time-accurate
sweep procedure, we make use of the time-accurate dual time-stepping method to determine a “quasi-steady”
solution at each time-step or increment of rotation angle very quickly. Therefore, the present sweep procedure
requires the use of a very large time step of the order of 1.0 sec which is ten to hundred thousand larger than a time-
step required for actual time-accurate solutions. The larger time step drives the rate of change of flow variables with
respect to angle of attack, for example, to zero as fast as possible. The solutions obtained are therefore, not time-
accurate, but are efficiently obtained using a time-accurate procedure.
Usually, an artificial or “inner” or “local” time variation term is added to the basic physical equations. This time
step and corresponding “inner-iteration” strategy is chosen to help satisfy the physical transient equations to the
desired degree. If the inner iterations converge, then the outer physical transient equations (their discretization) are
satisfied exactly; otherwise approximately. The time step for the inner iterations is allowed to vary spatially. Also,
relaxation with multigrid (algebraic) acceleration is employed to reduce the residues of the physical transient
equations. It is found that an order of magnitude reduction in the residues is usually sufficient to produce a good
transient iteration. This may require anywhere from a 5 and 50 inner iterations depending on the magnitude of the
outer time step, the nature of the problem, the nature of the boundary conditions and the consistency of the mesh
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

with respect to the physics at hand. Typically, 25 inner iterations are found to be adequate in the present simulations
using the time-accurate sweep procedure. This inner iteration is a critical parameter in the use of the present time-
accurate sweep procedure to obtain the final solution at each time-step or rotation angle increment.

III. Model Geometry and Computational Mesh

Previous work17 in this effort focused on two projectile models and several computational mesh types. The
computations performed using the sweep procedure are limited to one geometric model and one mesh type. The
information regarding this model and mesh is presented again at this time for clarity and readability of this paper.
The geometric model for the elliptic missile
is a sharp-nosed monoplane winged missile with
tail fins. The overall body length (L) is 711.2
mm with a 3:1 elliptical cross section, having
maximum body width and height of 176.0 mm
and 58.7 mm, respectively. There are four tail
fins on the aft end of the missile, aligned with
the base of the missile. This model, as shown in
Figure 2, has been used extensively in wind
tunnel tests at the NASA Langley Research
Center.18,23,24 and therefore, an extensive amount
of experimental data is also available for this
missile configuration for comparison with
computed numerical solutions. The center of
gravity is located at 0.6L from the nose of the
missile configuration.
The viscous grid previously generated for
this configuration was selected for the sweep
computation. It is an unstructured mesh created Figure 2. Schematic diagram of the complex body-wing-tail
using GRIDGEN25, a commercial software elliptic missile.
package for grid generation. Figure 3 shows the
expanded view of the full grid near the body of the missile. This grid consists of approximately 1.6 million cells for
the body-wing-tail configuration. This viscous mesh is a hybrid unstructured grid that consists of triangular prisms
and tetrahedral cells and thus, includes prism layers near the body surfaces to capture the boundary layer. Figure 4
shows an expanded view of the grid near the base for the missile configuration. It clearly shows the hybrid
unstructured mesh, including the prism layers near the body surface. The boundary spacing near the body wall was
selected in order to achieve a y+ value of less than 1.0 for both Mach number runs. In general, most of the grid
points are clustered in the boundary-layer, the nose and the near wake regions

4
American Institute of Aeronautics and Astronautics
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

Figure 3. Unstructured mesh for the body-wing-tail elliptic missile

Figure 4. Expanded view of the mesh near the base showing the prism layers.

IV. RESULTS

As stated earlier, numerical computations were performed to assess methodology used to predict the flow field
and aerodynamic coefficients for a complex missile configurations accurately and efficiently. Calculations were
performed at Mach Number 2.5 and matched the wind tunnel conditions provided: Re = 5.56 x 106 per meter, total
temperature, T0 = 339 K, and total pressure, P0 = 81.36 kPa. These conditions corresponded to a static temperature
and pressure of 150.67 K and 4761.8 Pa, respectively for the CFD calculations. Numerical computations were also
performed for this same configuration at Mach number M = 0.9; In this case, calculations were performed for
matching wind tunnel conditions of Re = 7.89 x 106 per meter, total temperature, T0 = 316 K, and total pressure, P0 =
59.95 kPa. All computations were performed for roll orientation of 0°. The angle of attack, α, varied from -5° to
30°, and the side slip (yaw) angle,  varied from -10° to 10°. Full three-dimensional computations were performed
and no symmetry was used. The results for both steady-state and sweep procedures are described below.
Computed results obtained for the Mach 2.5 case are presented first. Both qualitative and quantitative results
are presented. Figure 5 shows the longitudinal Mach contours for the body-wing-tail configuration in the pitch
plane of symmetry. The longitudinal Mach contours here are shown for three different angles of attack, α = -4.5o,
20o, and 31o and  = 0°. Note that the time-accurate sweep was performed from α = -4.5o to α =31o at increments of
0.5° for  = 0° and computed results are available for the entire range of angles of attack. As seen in Figure 5, the
flow field is slightly asymmetric at small angle of attack, α = -4.5o and as expected is highly asymmetric as angle of
attack is increased to higher values. One can clearly see a much strong shock wave emanating from the nose of the
missile in the wind side at larger angles of attack, α = 20o and 31o. The flow field in the near wake also shows the
strong asymmetry as well. Computations also were performed for various side slip angles for given angles of attack

5
American Institute of Aeronautics and Astronautics
to determine the lateral characteristics. Figure 6 shows some representative circumferential pressure contours at
four longitudinal stations X/L = 0.3, 0.6, 0.89 and 1.0. These results were obtained at an angle of attack, α = 20o and
side slip angle, = 8.5ofor the body-wing-tail configuration. One can clearly see the cross-flow separation in this
case at all four selected axial locations. The flow field is highly asymmetric between the wind and the lee sides due
to the large angle of attack with high pressures observed on the wind side and low pressures on the lee side (shown
in blue). One can also see the asymmetry in the computed flow field both on the lee and the wind sides between the
left and the right halves due to the side slip angle effect. The flow field seems to be even more complex due to the
presence of the aft fins. These fin-body interaction flow fields can be seen at X/L = 0.89 and this effect continues at
X/L = 1.0 which correspond to a station just after the end of the fin.
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

Figure 5. Longitudinal Mach contours for the pitch symmetry plane, M = 2.5,  = 0° and
 = -4.5° (bottom),  = 20° (middle), and  = 31° (top).
A representative set of surface pressure coefficients have been obtained from the computed solutions and
compared with the experimental data.24 Figure 7 shows the comparison of the computed circumferential surface
pressure coefficients with the data. Comparison is made at two axial locations, X/L = 0.6 and 0.95 and at two angles
of attack, α = 5o and 20o. As seen in this figure, computed results for both angle of attack cases match very well
with the data. One can observe lower pressures in the lee side and higher pressures on the wind side of the missile.

6
American Institute of Aeronautics and Astronautics
In addition, with increase in angle of attack from 5o to 20o as expected, the wind side pressures become higher and
the lee side pressures become lower.
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

(a) x/L = 0.30 (b) x/L = 0.60

(c) x/L = 0.89 (d) x/L = 1.0

Figure 6. Circumferential pressure contours at different x/L positions, M = 2.5,  = 20°,  = 8.5°.

To demonstrate the capability of the new time-accurate sweep procedure, aerodynamic force and moment
coefficients have been obtained from the computed CFD results and are shown next. Computed results have been
obtained using the time-accurate sweep procedure with the sweep performed from α = -4.5o to α =31o at increments
of 0.5° for a fixed  = 0°. As stated earlier, a critical parameter in the sweep procedure is the number of inner time
iterations used. The number of inner iterations was varied from 5 to 50 and solutions were obtained using the sweep
procedure. Figure 8 shows the convergence history of the solution residual for the inner iterations as a function of
the number of inner time iterations used. With 5 inner iterations, the solution residual drops by less than an order of
magnitude. With increasing number of inner iterations, the drop in the solution residual is increased. If an order of
magnitude drop in the residual is desired, the number of inner iterations required should be of the order of 25. When
the number of inner time iterations is increased to 50, the residual drops by almost two orders of magnitude. Figures
9,10, and 11 show the effect of the number of iterations on the computed solutions. In addition, steady-state
computations obtained earlier17 at various angles of attack are used for comparison with the results obtained from
the sweep calculations. Figures 9, 10, and 11 show the static aerodynamic coefficients (axial force, normal force,
and the pitching moment) as a function of angle of attack. As seen in Figure 9, one can see a very small difference
between the steady state results and the sweep results obtained with 5 inner time iterations. As the number of inner
time iterations is increased to 10, 25, and 50, computed total axial force is practically unchanged and matches well
with the steady state results. The predicted normal force with the sweep procedure is in good agreement with the
steady state results (Figure 10) and seems to be rather insensitive to the number inner time iterations. Comparison
of pitching moment coefficients between the sweep results and the steady state solutions is shown in Figure 11.
Clearly, the sweep results obtained with 5 inner iterations is inadequate. The comparison improves substantially
when 10 inner iterations or more are used in the sweep procedure. Hardly any difference is seen between the sweep
results obtained at 25 and 50 inner iterations. From these results and comparisons, it was decided that 25 inner time
iterations are sufficient and were subsequently used in all calculations performed using the time-accurate sweep
procedure.

7
American Institute of Aeronautics and Astronautics
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

X/L = 0.6

X/L = 0.95

Figure 7: Surface pressure coefficients at different angles of attack, M = 2.5,  = 0°.

8
American Institute of Aeronautics and Astronautics
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

9
(b) Iterations = 10
(a) Iterations = 5

American Institute of Aeronautics and Astronautics


Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

(c) Iterations = 25

(d) Iterations = 50

Figure 8. Time-history of solution residual as a function of inner iterations, M = 2.5,  = 20°,  = 8.5°.

10
American Institute of Aeronautics and Astronautics
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

Figure 9. Total axial force coefficient as a function of angle of attack, M = 2.5,  = 0°.

Figure 10: Normal force coefficient as a function of angle of attack, M = 2.5,  = 0°.

11
American Institute of Aeronautics and Astronautics
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

Figure 11: Pitching moment coefficient as a function of angle of attack, M = 2.5,  = 0°.

Computed results obtained using the time-accurate sweep procedure and 25 inner time iterations with the sweep
performed from α = -4.5o to α =31o are compared with available experimental data. With the sweep, computed
results are available for the entire range of angles of attack. In addition, steady-state computations obtained earlier17
at various angles of attack are also used for comparison with the data. Figures 12, 13, and 14 show the static
aerodynamic coefficients (axial force, normal force, and the pitching moment) as a function of angle of attack.
These results correspond to zero roll and zero side slip angle for the missile configuration. Both sets of computed
results obtained for the body-wing-tail configuration are shown in these figures and are compared with the
experimental data.23 Figure 12 shows the axial force coefficient as a function of angle of attack. The axial force
coefficient shown here does not include the base drag and represents the forebody only. The axial force coefficient
seems to increase quite a bit with the increase in angles of attack from 10 to 30 degrees. Both computed and the
experimental data show the same trends. Some discrepancy can be noted between the computed results and the
experimental data at angles of attack less than 10° and at higher angles of attack near α =25°. These results are still
considered quite good, given the modest size of the computational meshes used in the CFD calculations. Figure 13
shows the comparison of the computed normal force coefficient with the experimental data for the body-wing-tail
configuration. Computed normal force coefficients obtained from both sets of steady-state and sweep calculations
for this configuration are in excellent agreement with the data at all angles of attack. As seen in this figure, the
normal force coefficients in both cases increase with increasing angles of attack as expected. Comparison of the
computed pitching moment coefficients with the experimental data is shown in Figure 14. The moment is
referenced to the center of gravity of the missile located at 0.6L from the nose of the missile. Again, this figure
shows the computed results obtained using both sets of steady-state and sweep calculations. Again, as pointed out
earlier, both sets of computed results match very well with each other. The pitching moment coefficient is almost
constant for angles of attack -5° to 20° and it decreases slightly with further increase in the angle of attack.
Computed pitching moment coefficients obtained from both sets of computations agree well the measured data from
α = -5o to 12o and are slightly under-predicted at higher angles of attack.

12
American Institute of Aeronautics and Astronautics
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

Figure 12. Forebody axial force coefficient as a function of angle of attack, M = 2.5,  = 0°.

Figure 13: Normal force coefficient as a function of angle of attack, M = 2.5,  = 0°.

13
American Institute of Aeronautics and Astronautics
3.00

2.50

2.00

1.50

1.00

CM 0.50

0.00

-0.50
EXP - BWT
-1.00
CFD-BWT-VISC
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

-1.50 CFD BWT


SWEEP
-2.00
-5.0 0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0

Angle of Attack (deg)

Figure 14: Pitching moment coefficient as a function of angle of attack, M = 2.5,  = 0°.

Numerical computations were carried out for various side slip angles also. The side slip (yaw) angle,  varied
from -10° to 10° for a given angle of attack. Computed results for varying side slip angles were obtained at three
different angles of attack, α =0°, 10°, and 20°. These computed results are presented next. Figures 15 and 16 show
the results for the side force and the yawing moment coefficients, respectively at α =0°. Computed results were
obtained using both sets of steady-state and sweep calculations for the body-wing-tail configuration and are
compared against each other. In addition, computed results are also compared with available experimental data. As
seen in these figures, computed results obtained using the sweep procedure match very well with the results obtained
using standard steady-state method.17 As shown in Figure 15, the computed side force coefficients for this missile
configuration correctly predict the trend similar to that observed in the experimental data. The side force coefficient
decreases with increasing side slip angle. Computed side force coefficients are generally well predicted for the
entire range of side slip angles considered here. Another parameter that is of particular interest in the lateral
characteristics is the yawing moment coefficient. Computed yawing moment coefficients are shown as a function of
yaw angles and are compared to the experimental data in Figure 16. As shown in this figure, computed yawing
moment coefficients obtained from the sweep calculations are in very good agreement with the steady-state results.
Here, both the data and the computed results show the yawing moment coefficient to increase with increasing yaw
angles. The agreement between the computed results and the experimental data is good only for all positive side slip
angles. One can clearly observe somewhat larger discrepancy between the computed yawing moment coefficients
and the measured data at negative side slip angles. Comparison of computed results with the data is shown in
Figures 17 and 18 for α = 10° case and in Figures 19 and 20 for α = 20° case for the missile configuration. As seen
in these figures, computed side force coefficients and the yawing moment coefficients obtained from both sets of
computations match fairly well at α = 10° and show some difference at higher angle of attack, α = 20°. At this high
angle of attack, the flow field for this complex missile configuration is also very complex (see Figure 6) and is very
difficult to compute using either steady or sweep procedure. The maximum discrepancy between the steady state
and sweep computational results is approximately 7% at α = 20° and occurs at the highest side slip angle considered
( = 10°). Given the complexity of the flow field, computed results obtained can still be considered good.
Comparison of the computed results with the experimental data is similar to the α = 0° case. Some discrepancy
between the computed results and the data can be noted near both the higher negative and positive side slip angles.
All CFD results (Figure 15 through 20) obtained by both methods are anti-symmetric with respect to zero side slip
angle. Experimental results clearly are not and perhaps contain some errors.

14
American Institute of Aeronautics and Astronautics
0.80

0.60

0.40

0.20

CY 0.00

-0.20

EXP - BWT
-0.40
CFD - BWT-STEADY
-0.60
CFD - BWT-SWEEP
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

-0.80
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10

Side Angle Beta (deg)

Figure 15: Side force coefficient as a function of side slip (yaw) angle, M = 2.5,  = 0°.

0.50

0.40

0.30

0.20

0.10

Cn 0.00

-0.10

-0.20
EXP - BWT
-0.30
CFD - BWT-STEADY
-0.40
CFD - BWT-SWEEP
-0.50
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10
Side Angle Beta (deg)

Figure 16: Yawing moment coefficient as a function of side slip (yaw) angle, M = 2.5,
 = 0°.

15
American Institute of Aeronautics and Astronautics
0.80

0.60

0.40

0.20

CY 0.00

-0.20

-0.40 EXP - BWT


Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

CFD - BWT-STEADY
-0.60
CFD - BWT-SWEEP
-0.80
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10
Side Angle Beta (deg)

Figure 17: Side force coefficient as a function of side slip (yaw) angle, M = 2.5,  = 10°.

0.60

0.40

0.20

Cn 0.00

-0.20

EXP - BWT
-0.40
CFD - BWT-STEADY

CFD - BWT-SWEEP
-0.60
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10
Side Angle Beta (deg)

Figure 18: Yawing moment coefficient as a function of side slip (yaw) angle, M = 2.5,
 = 10°.

16
American Institute of Aeronautics and Astronautics
0.80

0.60

0.40

0.20

CY 0.00

-0.20

EXP - BWT
-0.40
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

CFD - BWT-STEADY
-0.60
CFD-BWT-SWEEP
-0.80
-5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10

Side Angle Beta (deg)

Figure 19: Side force coefficient as a function of side slip (yaw) angle, M = 2.5,  = 20°.

0.40

0.30

0.20

0.10 EXP - BWT

CFD - BWT-STEADY
Cn 0.00
CFD-BWT-SWEEP
-0.10

-0.20

-0.30

-0.40
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10

Side Angle Beta (deg)

Figure 20: Yawing moment coefficient as a function of side slip (yaw) angle, M = 2.5,
 = 20°.

Computed results were also obtained at a transonic speed, M = 0.9 for various angles of attack. These computed
results are compared with limited available data for this case. Figure 21 shows the longitudinal Mach contours for
the body-wing-tail configuration in the pitch plane of symmetry. The longitudinal Mach contours here are shown
for three different angles of attack, α = 0o, 10o, and 20o and  = 0°. Note that the time-accurate sweep was

17
American Institute of Aeronautics and Astronautics
performed from α = -5o to α =25o at increments of 0.5° for  = 0° and computed results are available for the entire
range of angles of attack. As seen in Figure 21, the flow field is symmetric at zero degree angle of attack. It is
asymmetric at angle of attack, α = 10o and as expected is highly asymmetric as angle of attack is increased to higher
values (α =20o). One can clearly see a much higher velocity field on the lee side and lower velocity field in the wind
side of the missile, especially at α = 20o. The flow field in the near wake also shows the strong asymmetry as well at
this high angle of attack.
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

Figure 21. Longitudinal Mach contours for the pitch symmetry plane, M = 0.9,  = 0° and
 = 0° (bottom),  = 10° (middle), and  = 25° (top).

Aerodynamic force and moment coefficients have been obtained from the computed CFD results. Again,
computed results have been obtained using the time-accurate sweep procedure with the sweep performed from α = -
5o to α =25o at increments of 0.5° for a fixed  = 0°. In addition, steady-state computations were also performed at
this Mach number at various angles of attack for comparison with the results obtained from the sweep calculations.
Figures 22, 23, and 24 show the static aerodynamic coefficients (axial force, normal force, and the pitching moment)
as a function of angle of attack. These results correspond to zero roll and zero side slip angle for the missile
configuration. Both sets of computed results obtained for the body-wing-tail missile configuration are shown in

18
American Institute of Aeronautics and Astronautics
these figures and are compared with the experimental data.24 The axial force coefficient shown in Figure 22 does
not include the base drag and represents the forebody only. Both sets of computed results (steady-state and sweep)
are seen to match very well with each other. Also, the axial force coefficient seems to decrease slowly up to an
angle of attack of about 4o and then decrease quite a bit with the increase in angles of attack up to 25o. Both
computed and the experimental data show the same trends and the computed results seem to match fairly well with
experimental data. Figure 23 shows the comparison of the computed normal force coefficient with the experimental
data. Computed normal force coefficients obtained from both sets of steady-state and sweep calculations for this
configuration are in excellent agreement with the data as well as with each other for all angles of attack. As seen in
this figure, the normal force coefficient increases with increasing angles of attack as expected. Comparison of the
computed pitching moment coefficients with the experimental data is shown in Figure 24. The moment is
referenced to the center of gravity of the missile located at 0.6L from the nose of the missile. Again, this figure
shows the computed results obtained using both sets of steady-state and sweep calculations. Both sets of computed
results match very well with each other. However, computed pitching moment coefficients obtained from both sets
of computations agree well the measured data from positive angles of attack and do not seem to match well at
negative angles of attack. Note that the experimental pitching moment is not zero at zero degree angle of attack.
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

The discrepancy at negative angles of attack could possibly be due to experimental errors and needs further
investigation.

0.30
EXP - BWT
0.25 CFD BWT SWEEP

CFD BWT STEADY


0.20

0.15

CXf 0.10

0.05

0.00

-0.05

-0.10
-5.0 0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0

Angle of Attack (deg)

Figure 22. Axial force coefficient as a function of angle of attack, M = 0.9,  = 0°.

19
American Institute of Aeronautics and Astronautics
12.00

10.00

8.00

6.00

CN EXP - BWT
4.00
CFD - BWT-SWEEP
CFD - BWT-STEADY
2.00

0.00
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

-2.00
-5.0 0.0 5.0 10.0 15.0 20.0 25.0 30.0

Angle of Attack (deg)

Figure 23: Normal force coefficient as a function of angle of attack, M = 0.9,  = 0°.

Figure 24: Pitching moment coefficient as a function of angle of attack, M = 0.9,  = 0°.

20
American Institute of Aeronautics and Astronautics
These numerical results were obtained utilizing the high performance computing resources at the U.S. Army
Research Laboratory (ARL) Department of Defense Supercomputer Resource Center (DSRC) at Aberdeen Proving
Ground, Maryland. All solutions were processed on MJM, a Linux Networx Advanced Technology Cluster,
providing 4400 cores and 8.8 TB of memory.26 Table 1 provides a comparison of resources used for a single
steady-state computation and a computation utilizing the time-sweep procedure (not including the steady-state
startup time). The sweep procedure provides seventy-two solutions in about the same amount of time needed for
three steady-state runs, for a 1.6 million cell mesh at Mach Number 2.5.

Table 1. Computational time and resource requirements, (1.6 million cells)

Run Type Steady-State Sweep (25 inner iterations)

Solutions per Run 1 (one angle of attack) 72 (in 0.5 deg increments)
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

Total Processing Hours 48 hrs 90 hrs

CPU Time 6 hrs on 8 CPUs 5.5 hrs on 16 CPUS

V. Concluding Remarks
A new time-accurate sweep procedure has been used to obtain aerodynamic data for a range of angles of attack
and side slip angles. Numerical computations using viscous Navier-Stokes methods were performed to predict the
flow field and aerodynamic coefficients on non-axisymmetric missile configurations for wind tunnel test conditions.
Full three-dimensional computations were performed and no symmetry was used. Computational results were
obtained for a complex body-wing-tail missile configuration with elliptic cross-section at a supersonic speed and
also a low transonic speed using a two-equation k- turbulence model. Steady-state solutions were also computed
for various angle of attack and side-slip angle parameters. Visually, contour plots of the numerical results show the
qualitative features of flow field for the missile at various angles of incidence. Aerodynamic coefficients have been
extracted from the computed solutions; results from both solution methods, the sweep and standard steady-state
simulations, have been found to generally match well with each other and with the available experimental data for
this configuration. The timing data shown in Table 1 for the Mach 2.5 steady-state and sweep procedures provides a
dramatic comparison. Clearly, the new sweep procedure provides results that compare well with the steady-state
solutions, but require significantly less resources, producing accurate aerodynamic data for complex missile
configurations in a timely and efficient manner.

References
1. Silton, S., “Navier-Stokes Computations for a Spinning Projectile from Subsonic to Supersonic Speeds,” Journal of
Spacecraft and Rockets, Vol 42, No 2, pp 223-231, 2005.
2. DeSpirito, J., M. Vaughn, D. Washington, “Numerical Investigation of Canard-Controlled Missile with Planar Grid Fins,”
Journal of Spacecraft and Rockets, Vol 40, No 3, pp 363-370, 2003.
3. Sahu, J., "Numerical Computations of Transonic Critical Aerodynamic Behavior," AIAA Journal, Vol. 28, No. 5, May
1990, pp. 807-816.
4. Sahu, J., Heavey, K. and Edge, H., “Numerical Computations of Supersonic Flow over Elliptic Projectiles”, ARL-TR-
2589, U.S. Army Research Laboratory, Aberdeen Proving Ground, MD, December 2001.
5. Weinacht, W. Sturek, L. Schiff, “Projectile Performance, Stability, and Free-Flight Motion Prediction Using
Computational Fluid Dynamics,“ Journal of Spacecraft and Rockets, Vol 41, No 2, pp 257-263, 2004.
6. Heavey, K. and Sahu, J. “Application of Computational Fluid Dynamics to a Monoplane Fixed-Wing Missile with Elliptic
Cross Sections.” ARL-TR-3549, U.S. Army Research Laboratory, Aberdeen Proving Ground, MD, July 2005.

21
American Institute of Aeronautics and Astronautics
7. Sahu, J., Despirito, J., Edge, H., Silton, S., and Heavey. K., “Recent Applications of Structured and Unstructured Grid
Techniques to Complex Projectile and Guided Missile Configurations”, Proceedings of the Eighth International Grid Generation
and Computational Field Simulations Conference, Honolulu, HI June 2002.
8. Sahu, J., “Unsteady Numerical Simulations of Subsonic Flow over a Projectile with Jet Interaction.” AIAA Paper 2003-
1352, Reno, NV, 6-9 January 2003.
9. Weinacht, P., “Coupled CFD/GN&C Modeling for a Smart Material Canard Actuator,” AIAA-2004-4712, AIAA
Atmospheric Flight Mechanics Conference, Providence, Rhode Island, 2004.
10. Heavey, K. and Sahu, J., “Computational Fluid Dynamics Results for Extended Area Protection System (EAPS)
Projectile Designs”, ARL-TR-4983, U.S. Army Research Laboratory, Aberdeen Proving Ground, MD, September 2009.
11. DeSpirito, J. and Heavey, K., ”CFD Computation of Magnus Moment and Roll-Damping Moment of a Spinning
Projectile,” AIAA-2004-4713, AIAA Atmospheric Flight Mechanics Conference, Providence, Rhode Island, 2004.
12. Sahu, J., “Unsteady CFD Modeling of Aerodynamic Flow Control over a Spinning Body with Synthetic Jet.” AIAA
Paper 2004-0747, Reno, NV, 5-8 January 2004.
13. Sahu, J., “Numerical Simulations of Supersonic Flow over an Elliptic-Section Projectile with Jet Interaction”, AIAA
Paper No. 2002-3260, AIAA Applied Aerodynamics Conference, St. Louis, Missouri, June 2002.
14. Sahu and C. J. Nietubicz, "Application of Chimera Technique to Projectiles in Relative Motion", Journal of Spacecraft
Downloaded by MONASH UNIVERSITY on November 28, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2010-8248

& Rockets, Sept-Oct 1995.


15. Sahu, J. and Heavey, K.R, “Unsteady CFD Modeling of Micro-Adaptive Flow Control for an Axisymmetric Body”,
International Journal of Computational Fluid Dynamics, Vol. 5, April-May 2006.
16. Sahu, J., “Time-Accurate Numerical Prediction of Free-Flight Aerodynamics of a Finned Projectile,” AIAA-2005-5817,
AIAA Atmospheric Flight Mechanics Conference, San Francisco, California, 2005.
17. Sahu, J. and K. Heavey, “Computations of Supersonic Flow over a Complex Elliptical Missile Configuration”, AIAA
Paper No. 2009-5714, Atmospheric Flight Mechanics Conference, Chicago, IL, August 2009.
18. Allen, J.M., G. Hernandez, and M. Lamb, “Body-Surface Pressure Data on Two Monoplane-Wing Missile
Configurations with Elliptical Cross Sections at Mach 2.50”, NASA Technical Memorandum 85645, Langley Research Center,
Hampton, VA, 1983.
19
Peroomian, O., S. Chakravarthy, and U. Goldberg, “A ‘Grid-Transparent’ Methodology for CFD.” AIAA Paper 97-07245,
1997.
20
Peroomian, O., S. Chakravarthy, S. Palaniswamy, and U. Goldberg, “Convergence Acceleration for Unified-Grid
Formulation Using Preconditioned Implicit Relaxation.” AIAA Paper 98-0116, 1998.
21
Goldberg, U. C., O. Peroomian, and S. Chakravarthy, “A Wall-Distance-Free K-E Model With Enhanced Near-Wall
Treatment.” ASME Journal of Fluids Engineering, Vol. 120, pp. 457-462, 1998.
22
Batten, P., U. Goldberg and S. Chakravarthy, "Sub-grid Turbulence Modeling for Unsteady Flow with Acoustic
Resonance", AIAA Paper 00-0473, 38th AIAA Aerospace Sciences Meeting, Reno, NV, January 2000.
23
Graves, E.B, “Aerodynamic Characteristics of a Monoplanar Missile Concept With Bodies of Circular and Elliptical Cross
Sections”, NASA Technical Memorandum 74079, Langley Research Center, Hampton, VA, 1977.
24
Graves, E.B. and R.H. Fournier, “Effect of Nose Bluntness and Afterbody Shape on Aerodynamic Characteristics of a
Monoplanar Missile Concept With Bodies of Circular and Elliptical Cross Sections at a Mach Number of 2.50”, NASA Technical
Memorandum 80055, Langley Research Center, Hampton, VA, 1979.
25
Pointwise, Inc., Gridgen Version 15, Bedford, TX, 2009.
26
ARL DSRC website: www.arl.hpc.mil/Systems/html, 2010.

22
American Institute of Aeronautics and Astronautics

You might also like