Ada 425202
Ada 425202
Ada 425202
Disclaimers
The findings in this report are not to be construed as an official Department of the Army position unless
so designated by other authorized documents.
Citation of manufacturer’s or trade names does not constitute an official endorsement or approval of the
use thereof.
Destroy this report when it is no longer needed. Do not return it to the originator.
Army Research Laboratory
Aberdeen Proving Ground, MD 21005-5066
14. ABSTRACT
Throughout its development, slender body theory has been generalized to predict a large variety of aerodynamic coefficients for
a wide class of flight bodies. For most applications, slender body theory provides only a qualitative predictive capability. There
is, however, a set of slender body relationships which has been previously derived by Sacks that allows the individual pitch-
damping coefficients and the pitch-damping coefficient sums to be related to each other. Until recently, it has been difficult to
assess the accuracy of these relationships due to the lack of high-quality pitch-damping data or the lack of a higher order theory.
The current work applies a recently developed computational fluid dynamics capability for predicting all three pitch-damping
coefficients. From this analysis, the accuracy of these relationships has been assessed and their engineering significance
demonstrated. One important result is that the pitch-damping relations developed by Sacks allow the individual pitch-damping
coefficients to be determined from the pitch-damping coefficient sum with a high degree of accuracy.
ii
Contents
List of Figures iv
1. Introduction 1
1.1 Helical Motions ...............................................................................................................3
1.2 Coning Motion ................................................................................................................6
2. Computational Technique 7
3. Results 9
3.1 Ogive-Cylinder Results ...................................................................................................9
3.2 Flared Projectile Results................................................................................................19
4. Summary 23
5. References 25
Distribution List 30
iii
List of Figures
direction CFD predictions, predictions using Sacks’ relation (equation 28) and
equation 2, ANSR, Mach 2.5, xcg /L = 0.5598 , and L/D = 9. ..................................................19
Figure 16. Comparison of the pitch-damping moment coefficient sum predicted
from Bryson’s relation (equation 1) with CFD predictions, ANSR, Mach 2.5,
xcg /L = 0.5598 , and L/D = 9. ..................................................................................................20
Figure 17. Schematic of the flared projectile geometry................................................................20
Figure 18. Comparison of the three computed deviations, ∆1 , ∆ 2 , and ∆ 3 relative to
the damping moment coefficients, flared projectile, and Mach 2............................................21
iv
Figure 19. Comparison of the damping moment coefficient predicted from Sacks’ relations
with CFD predictions, flared projectile, and Mach 2...............................................................22
Figure 20. Comparison of the damping moment coefficient predicted from Sacks’ relations
with CFD predictions, flared projectile, and Mach 5...............................................................22
Figure 21. Comparison of the damping moment coefficient predicted from Sacks’ relations
with CFD predictions, 15° flared projectile, and Mach 5. .......................................................23
v
INTENTIONALLY LEFT BLANK.
vi
1. Introduction
The pitch-damping moment coefficients C mq (due to body transverse angular rate) and C mα
(due to angular rate associated with angle of attack) play an important role in the performance
and dynamic stability of flight bodies. The pitch-damping moment coefficient sum Cmq + C mα
is of most practical importance, although the individual damping coefficients are often required
in aerodynamic analyses. Throughout the past several decades, a variety of techniques and
theories have been developed for predicting the pitch-damping coefficients (1–10). These
techniques vary in their ease of use as well as their ability to accurately predict the pitch-
damping coefficients.
During the course of its development, slender body theory was generalized to predict a large
variety of aerodynamic coefficients including the pitch-damping coefficients (1–2). In general,
direct application of these methods provides only qualitative results for the aerodynamic
coefficients. However, elements of slender body theory have been incorporated into current
engineering methods. These methods (3–5) have evolved considerably, although their
implementation is fairly complex. Apart from implementation issues, modern engineering
methods, once embodied into a computer code, are relatively easy to use and provide fast and
reasonably accurate aerodynamic predictions for a large variety of flight geometries.
From slender body theory, some important relationships between the various aerodynamic
coefficients can be derived, although these relationships only hold rigorously within the context
of theories from which they were obtained. Bryson (2) derived the relatively well-known slender
body result that relates the pitch-damping moment coefficient sum to the normal force
coefficient, shown in equation 1:
2
⎛ L − x cg ⎞
C mq + C mα = −⎜⎜ ⎟ CN .
⎟ α
(1)
⎝ D ⎠
Sacks (1), using the Blasius method for calculating the forces and moments on slender bodies
from the cross-flow potential, found that many of the aerodynamic coefficients were related to
each other. Sacks obtained expressions that directly related the individual pitch-damping
coefficients to the normal force coefficient, such as the pitch-damping force coefficient shown in
equation 2. These expressions have a form similar to Bryson’s result shown in equation 1:
⎛ L − x cg ⎞
C Nq = ⎜⎜ ⎟C N .
⎟ α (2)
⎝ D ⎠
In practice, these relations that directly relate the damping coefficients to the normal force
coefficient do not perform particularly well, even when the slender body evaluation of the
1
normal force coefficient is replaced with a more accurate evaluation of the normal force
coefficient from sources such as experimental measurement or computational fluid dynamics
(CFD) (6). However, these relations can be combined with empirical corrections to yield more
reliable results (6, 7).
Sacks also found expressions that related the individual damping coefficients to each other,
including the following relationship between the pitch-damping force coefficients:
C Nq = C Nα − C mα . (3)
Sacks’ explicitly derived relation shown in equation 3 can be easily generalized using his theory
to the individual pitch-damping force and moment coefficients, and the pitch-damping force and
moment sums as shown in equations 4–9. For the purposes of this report, these relationships will
be referred to as Sacks’ relations:
C Nq = C Nα − C mα , (4)
C mq = C mα − C 2 mα , (5)
[C Nq ]
+ C Nα = 2C Nα − C mα , (6)
[C mq ]
+ C mα = 2C mα − C 2 mα , (7)
[C Nq + C Nα ] = 2C Nq + C mα , (8)
and
[C mq ]
+ C mα = 2C mq + C 2 mα . (9)
Sacks’ contribution is the recognition that these relationships exist, although the same relations
are implicitly contained in other theories such as that presented by Bryson (2).
For slender bodies, these relations are independent of configuration and are applicable to both
winged and wingless bodies. The importance of these relationships is that if just one of the three
the other two damping coefficients can be obtained using simple closed-form expressions. This,
of course, assumes that the pitching moment coefficient C mα (first moment of the normal force)
and the second moment of normal force C 2 mα can be obtained as well. Both of these
coefficients can be obtained if the normal force distribution is known as shown in equations 10
and 11:
2
x ( x cg − x ) d C Nα
C mα = ∫ dx , (10)
0
D dx
and
x ( x cg − x ) 2 d C Nα
C 2mα = ∫ dx . (11)
0 D2 dx
Predictive methods for the static normal force and static pitching moment are well established,
even for fast-design methods.
Because Sacks’ relations were derived using simple approximate theories, it remains to be shown
whether the validity of the relations shown in equations 4–9 exists only within the context of the
theories from which they were derived or whether they are universally valid for slender bodies,
or perhaps, more importantly, whether they are of general engineering significance. It is
important to note that the relations shown in equations 4–9 differ somewhat from the results in
equations 1 and 2 because they relate the pitch-damping coefficients to each other rather than
purely to the static normal force coefficient. This suggests that Sacks’ relations may more
properly represent the physics compared to their counterparts which directly relate the damping
coefficients to the normal force coefficient. Because of this, there is reason to investigate the
validity and accuracy of Sacks’ relations.
Recently, a computational approach for predicting all three of the pitch-damping coefficients has
been developed (8, 9). The approach solves the three-dimensional (3-D) thin-layer Navier-
Stokes equations for three different imposed motions that allow the three pitch-damping
coefficients to be predicted independently. The pitch-damping force and moment coefficient
sums are determined from the computation of a body undergoing an imposed coning motion.
The individual pitch-damping coefficients are obtained from computations of a body undergoing
two specific types of imposed helical motions. Each of these motions is described in the
following sections. One of the key components of this method is that steady flow techniques can
be employed to predict aerodynamic derivatives normally associated with time-dependent
motions.
3
Figure 1. Helical motion with nonzero α and zero q .
components associated with the coefficients C mα and C mα , respectively. This motion is
referred to as “q = 0 helical motion” because the angular rates associated with the damping
coefficient C mq are zero.
For the second motion, the longitudinal axis of the flight vehicle remains tangent to the helical
flight path at each point along the trajectory. Figure 2 shows a 3-D view of this motion. The
angle of attack of the incident airstream is zero because both the longitudinal axis of the body
and the free-stream velocity vector are tangent to the flight path. The resulting yawing rate is
also zero because the angle of attack is constant. The angular orientation of the flight body
changes continuously with respect to an earth-fixed reference frame, producing a nonzero
transverse angular rate.
As a result, moment components associated with the damping moment C mq are produced. This
motion is referred to as “ α = 0 helical motion” because the angular rates associated with the
damping coefficient C mα are zero.
For each of the helical motions, the transverse aerodynamic moment in the nonrolling frame will
be periodic in time, which also indicates that the flowfield will be periodic in time when viewed
from the nonrolling coordinate frame. The time dependency is removed by transforming to an
orthogonal right-handed coordinate system that has its x-axis aligned with the longitudinal axis
of the body and its z-axis along a line between the body’s CG and the axis of rotation of the
helix.
4
Figure 2. Helical motion with zero α and nonzero q .
For each of the helical motions previously described, the spin rate of the body has not been
defined. To eliminate any contributions to the aerodynamic forces and moments from the
Magnus forces and moments, the spin rate is fixed to zero (see references [8] and [9] for details).
The resulting in-plane moment (C m ) and side moment (C n ) coefficients in the transformed
coordinate system for both type of helical motions are shown in equations 12 and 13:
Zero − spin q = 0 helical motion
⎛ ΩD ⎞ R o Ω R Ω
C m + iC n = −C mα ⎜ ⎟ + iC mα o , (12)
⎝ V ⎠ V V
5
Zero − spin α = 0 helical motion
⎛ ΩD ⎞ R o Ω
C m + iC n = C mq ⎜ ⎟ . (13)
⎝ V ⎠ V
Here, Ω is the angular velocity of the body about the helix axis, R o is the perpendicular
distance between the helix axis and the body CG, and V is the total linear velocity of the CG.
Similar expressions for the individual damping force coefficients can be developed using the
same approach as applied for the moment coefficients.
Relative to a nonrolling coordinate frame, both the transverse angular rate of the body and the
angular rate associated with the angle of attack vary in a periodic manner thereby exciting the
aerodynamic forces and moments associated with both of the individual pitch-damping
coefficients. Here, a specific form of coning motion, described as zero-spin coning motion, is
employed. In zero-spin coning motion, the total angular velocity of the body along the
longitudinal axis (the spin rate) is zero. By imposing zero spin rate on the body, the
contributions from the Magnus forces and moments are eliminated.
The time dependency is removed by transforming the body-fixed nonrolling coordinate frame to
an orthogonal right-handed coordinate system that has its x-axis aligned with the longitudinal
axis of the body and its z-axis in the pitch-plane of the body. Within this transformed coordinate
frame, in-plane moment (C m ) and side moment (C n ) coefficients have the following form:
6
Zero-spin coning motion
⎛ ΩD ⎞
C m + iC n = iδ ⎜
⎝ V ⎠
[ ]
⎟ C mq + γ C mα + C mα δ . (14)
Here, the side-moment is proportional to the pitch-damping moment coefficient sum and varies
linearly with the coning rate Ω and sine of the total angle of attack δ . For small angles of
attack, the cosine of the total angle of attack, γ, can be assumed to be one.
The pitch-damping coefficient sum can also be determined by simply adding the individual
damping coefficients. In practice, there is very little difference in the two results. However, for
the current study, directly predicting the pitch-damping coefficient sum using coning motion
provides an alternative prediction of the pitch-damping sum and additional confirmation of the
predictions of the individual coefficients.
2. Computational Technique
In the previous section, several types of steady motion were presented that produce aerodynamic
forces and moments from which the various pitch-damping coefficients can be obtained. One
unique feature of these motions is that they are steady. The advantage of a steady motion over
an unsteady motion is that a potentially time-independent flowfield can be produced by a steady
motion, permitting analysis using steady flow CFD techniques. Such techniques can be
computationally less expensive than time-dependent CFD approaches. To fully exploit the
steady character of the flow, special body-fixed coordinate systems have been employed to
capture the steady flowfield. One feature of these coordinate frames is that they are rotating at a
constant rate with respect to an inertial frame. Because of this, the governing equations of fluid
motion must be modified to take into account the centrifugal and Coriolis force terms associated
with the noninertial rotating frame.
The steady thin-layer Navier-Stokes equations are shown in equation 15:
The inviscid flux vectors, Ê , F̂ , and Ĝ , the viscous term Ŝ , the inviscid and viscous source
terms due to the cylindrical coordinate formulation, Ĥ c and Ŝc , and the source term Ĥ,
containing the Coriolis and centrifugal force terms that result from the rotating coordinate frame,
are functions of the dependent variables represented by the vector q T = ( ρ , ρu, ρv, ρw , e) , where
ρ and e are the density and the total energy per unit volume, and u , v , and w are the velocity
components in axial, circumferential, and normal directions. The inviscid flux vectors and the
7
source term are shown in equation 16. Details of the thin-layer viscous term are available in the
literature (11):
⎡ ρU ⎤ ⎡ ρV ⎤
⎢ ρuU + ξ p⎥ ⎢ ρuV + η p ⎥
1⎢
x ⎥
1⎢ ⎥
x
Ê = ⎢ ρvU ⎥ F̂ = ⎢ ρvV + ηφ p / r ⎥
J⎢ ⎥ J⎢ ⎥
⎢ ρwU ⎥ ⎢ ρwV + η r p ⎥
⎢⎣ (e + p )U ⎥⎦ ⎢⎣ (e + p )V ⎥⎦
(16)
⎡ ρW ⎤ ⎡ 0 ⎤
⎢ ρuW + ζ p ⎥ ⎢ ρ fx ⎥
1⎢
x ⎥ 1 ⎢ ⎥
Ĝ = ⎢ ρvW + ζ φ p / r ⎥ Ĥ = ⎢ ρ fφ ⎥.
J⎢ ⎥ J⎢ ⎥
⎢ ρwW + ζ r p ⎥ ⎢ ρ fr ⎥
⎢⎣ (e + p )W ⎥⎦ ⎢ ρ u f x + ρ v fφ + ρ w f r ⎥
⎣ ⎦
The pressure, p, can be related to the dependent variables by applying the ideal gas law:
⎡ ρ 2 ⎤
p = (γ − 1) ⎢ e −
⎣ 2
(u + v2 + w 2 ⎥ .
⎦
) (17)
The turbulent viscosity, µt , which appears in the viscous matrices, was computed using the
Baldwin-Lomax turbulence model (12).
The Coriolis and centrifugal acceleration terms due to the rotating coordinate system, which are
contained in the source term, Hˆ , are shown in equation 18:
G G G G G G
(
f = 2Ω × u + Ω × Ω × R . ) (18)
The Coriolis acceleration is a function of the angular velocity of the coordinate frame with
G G
respect to the inertial frame, Ω , and the fluid velocity vector, u, which can be represented by the
velocity components, u , v , and w . The centripetal acceleration is a function of the angular
G G
velocity of the rotating frame, Ω , and the displacement vector, R, between the axis of rotation
G
and the local position in the flowfield. The acceleration vector, f, can be written in terms of its
components along the x , φ , and r axes, f x , f φ , and f r .
The steady thin-layer equations are solved using the parabolized Navier-Stokes (PNS) technique
of Schiff and Steger (11). This “space-marching” approach integrates the governing equations
from the nose of the flight body to the tail. Following the approach of Schiff and Steger, the
governing equations, which have been modified here to include the Coriolis and centrifugal force
terms, are solved using a conservative, approximately factored, implicit finite-difference
8
numerical algorithm as formulated by Beam and Warming (13). Details of the implementation
of the source term that contains the Coriolis and centrifugal force terms are given in references
(8) and (9).
The technique has been validated with available experimental data where possible, and excellent
agreement is found (8, 14). Grid resolution studies were also performed in the original studies to
ensure grid-independent solutions (8, 9).
3. Results
In the current context, this computational procedure allows the general applicability of Sacks’
pitch-damping relations to be examined. Arguably, up to now, it has not been possible to assess
the validity or accuracy of these relationships due to the uncertainty associated with
experimentally derived pitch-damping data and the lack of a higher order theory. In the current
research effort, the technique has been applied to two axisymmetric body geometries: an ogive-
cylinder configuration and a cone-cylinder body with a flared afterbody. The accuracy of Sacks’
relations is examined for each of these body geometries in supersonic flight.
Note that by simple algebraic manipulations, the following relations that are algebraically
equivalent to equation 19 are found:
[ ]
∆ = C Nq + C Nα − C mα − 2C Nq ≡ ∆ 2 , (20)
and
[ ]
∆ = 2C Nα − C mα − C Nq + C Nα ≡ ∆ 3 . (21)
9
Figure 4. Schematic of ANSR configuration.
Essentially, these deviations are a measure of the accuracy of the Sacks’ relations shown in
equations 4, 6, and 8. It could be argued that the differences ∆1 , ∆ 2 , and ∆ 3 are numerical
errors that are not representative of physical phenomena. However, it should be noted that the
deviation ∆ can be computed independently in three different ways. The deviation ∆1 is
computed using the individual damping coefficients C Nq and C Nα obtained independently
from two different types of helical motion, while ∆ 2 and ∆ 3 are computed from the pitch-
damping coefficient sum C Nq + C Nα and one of the two individual damping coefficients C Nq
and CNα , respectively. The importance of having three independent methods for determining ∆
is to demonstrate that the deviation is not due to an “error” in the prediction of any one of the
three damping coefficients. Similar expressions for the deviation of the moment coefficients ∆
can also be obtained and are shown in equations 22–24:
∆ = C mα − C 2 mα − C mq ≡ ∆1 , (22)
[ ]
∆ = C mq + C mα − C 2mα − 2C mq ≡ ∆ 2 , (23)
and
[ ]
∆ = 2C mα − C 2 mα − C mq + C mα ≡ ∆ 3 . (24)
10
Figure 5 shows the computed deviation for the force coefficient as a function of body length at
Mach 2.5. Here, the body has been lengthened to 20 calibers to more clearly illustrate the
variation of the deviation with body length. There is very good correlation between the three
different methods of computing ∆ , suggesting that the deviation is representative of a physical
effect rather than simply numerical error. The deviation is very small over the nose and begins
to grow in a linear fashion a couple of body diameters aft of the nose. The deviation (for the
force) can be shown to be independent of CG position using the CG translation relations for the
individual force coefficients. This was also confirmed by varying the CG position in the
computations as well. The computed deviation was also found to be somewhat dependent on
nose length and Mach number, although the results shown are representative of the trends
observed for the other flight conditions.
5 ∆1
∆2
∆3
4
Deviations
0
0 4 8 12 16 20
X/D
Figure 6 shows the computed deviations for the force, ∆1, ∆ 2 , and ∆ 3 , compared with the
individual pitch-damping coefficients and the pitch-damping sum. Relative to C Nα , and the
pitch-damping force sum, the deviation is quite small. The distribution of the deviation is also
small compared with the pitch-damping force coefficient C Nq , except near the end of the body
where C Nq itself is nearly zero. Equation 4 shows that according to Sacks’ relation, the
difference between C Nα and C Nq is the pitching moment coefficient Cmα . This is also
graphically shown in figure 6. Near the nose of the body, the difference between the two
damping coefficients, C Nα and C Nq , increases at nearly the same rate as the pitching moment
11
20
C Nq + C Nα
15
10
C mα C Nα
Coefficients
∆1 , ∆ 2 , ∆ 3
0
C Nq
-5
-10
0 3 6 9
X/D
coefficient. On the aft of the body, the pitching moment coefficient is nearly constant, and the
difference between C Nα and C Nq becomes relatively constant.
Figure 7 shows the computed deviation for the moment relative to the three damping coefficients
of interest. Again, there is very good correlation between the three methods for computing the
deviation. The computed deviation is small compared to individual damping moment
coefficients and to the pitch-damping moment sum. Sacks’ relation for the moment (equation 5)
shows that the difference between C mα and C mq is the pitching second-moment coefficient
C 2 mα . This also can be seen in figure 7. In a sense, figures 6 and 7 provide some measure of
the expected error in applying Sacks’ relations to obtain the various damping moment
coefficients given that normal force distribution (or C mα and C 2 mα ) and one of the pitch-
damping coefficients is known.
The most likely application of Sacks’ relations in practical situations is to compute the individual
damping coefficients from the pitch-damping coefficient sum because the pitch-damping
moment sum is much easier to measure. Through simple algebraic manipulations of equations
7 and 9, the following form of the Sacks’ relations can be obtained:
12
C mα = ( [C mq ] )
+ C mα + C 2 mα / 2 , (25)
and
C mq = ( [C mq ] )
+ C mα − C 2 mα / 2 . (26)
60
40
C 2 mα
20
C mα
Coefficients
∆1 , ∆ 2 , ∆3
-20
-40
C mq
C mq + C mα
-60
-80
0 3 6 9
X/D
Figure 8 shows a comparison of the pitch-damping moment coefficients Cmq and C mα
obtained by applying the Sacks’ relationships using the pitch-damping moment coefficient sum
⎡Cmq + Cm ⎤ and the second-moment coefficient of the normal force C 2 mα . Here, CFD has
⎣ α ⎦
been used to compute both ⎡Cmq + Cm ⎤ and C 2 mα . The results from the application of these
⎣ α ⎦
relations are compared with the CFD predictions of the individual damping coefficients. The
predictions were obtained at a flight velocity of Mach 2.5. The coefficients C mq and C mα are
overpredicted and underpredicted by ~5% and 12%, respectively, although the absolute error is
similar for both coefficients. The distribution of the damping coefficients over the body is also
very well predicted.
13
10
C mα
Pitch-Damping Coefficients
-10
-20
C mq
-30
-40
Direct CFD Prediction
Derived from Sacks' Relations
-50
-60
0 3 6 9
X/D
Figures 9–11 show similar comparisons for flight velocities of Mach 1.8, 3.5, and 4.5. The
relative comparisons are very similar to the predictions at Mach 2.5 shown in figure 8. The
largest Mach number effect appears to occur on the cylindrical portion of the body rather than on
the nose. The relative Mach number effect, though relatively small, is captured in the application
of the Sacks’ relations.
Similar analysis can be performed for the force coefficients. Figure 12 shows the comparison of
the individual pitch-damping moment coefficients C Nq and C Nα predicted using the form of
Sacks’ relations in equations 27 and 28 with direct CFD predictions. The distribution of the
force coefficients along the body is very well predicted using Sacks’ relations:
C Nα = ( [C Nq ]
+ C Nα + C mα / 2 , ) (27)
and
C Nq = ( [C Nq ]
+ C Nα − C mα / 2 . ) (28)
14
10
Pitch-Damping Coefficients
-10
C mα
-20
-30
C mq
-40
-60
0 3 6 9
X/D
10
C mα
Pitch-Damping Coefficients
-10
-20
C mq
-30
-60
0 3 6 9
X/D
Figure 10. Comparison of the damping moment coefficients predicted from Sacks’
relations with CFD predictions, ANSR, Mach 3.5, xcg /L = 0.5598 , and
L/D = 9.
15
10
C mα
Pitch-Damping Coefficients
-10
-20
C mq
-30
-40
Direct CFD Prediction
Derived from Sacks' Relations
-50
-60
0 3 6 9
X/D
Figure 11. Comparison of the damping moment coefficients predicted from Sacks’
relations with CFD predictions, ANSR, Mach 4.5, xcg /L = 0.5598 , and
L/D = 9.
20
15
Pitch-Damping Coefficients
C Nα
5
C Nq
-5
-10
0 3 6 9
X/D
Figure 12. Comparison of the damping force coefficient predicted from Sacks’ relations
with CFD predictions, ANSR, Mach 2.5, xcg /L = 0.5598 , and L/D = 9.
16
Figure 13 shows the results for the individual pitch-damping coefficients for various CG
positions for the body with the L/D ratio of 9 at Mach 2.5 obtained by applying equations 27 and
28. The results are compared with CFD results and with direct slender body theory results.
30
20
CFD Prediction
10 Slender Body Theory C mα
Pitch-Damping Coefficients
Sacks' Relations
0
-10
-20
-30
-40
C mq
-50
-60
-70
3 4 5 6 7
Xcg/D
Figure 13. Comparison of the predicted damping moment coefficients with CFD and
slender body theory results, ANSR, Mach 2.5, and L/D = 9.
There is very good correlation of the results obtained with Sacks’ relations and the CFD results,
and the results are a significant improvement over the slender body theory.
For the results shown in figures 8–12, comparisons have been made between the force and
moment distributions along the body. This was done because these distributions are easily
extracted from the computational results, and they allow detailed examination of the
performance of Sacks’ relations. It must be emphasized that it is not necessary to obtain the
pitch-damping force and moment distributions in order to apply Sacks’ relations because they are
applicable for the global force and moment coefficients as well. This is particularly important to
consider when the source of the pitch-damping coefficient data is from experiment where only
global coefficients are typically available. However, Sacks’ relations do require the normal force
distribution to determine C 2 mα because generally only the normal force and pitching moment
are available as global coefficients from many sources. In lieu of more sophisticated
computational methods for determining the normal force distribution, fast-design aeroprediction
codes (3) should provide acceptable accuracy for a variety of flight vehicle geometries.
17
To demonstrate the use of Sacks’ relations when only global forces and moments are available,
pitch-damping moment coefficient sum data from range firings of the ANSR have been used to
estimate the individual pitch-damping moment coefficients using Sacks’ relations. The normal
force distribution from the fast-design aeroprediction code AP02 (3) has been used to predict
C 2 mα . Figure 14 shows the predicted pitch-damping coefficients obtained from Sacks’ relations
using experimental measurements of the pitch-damping coefficient sum and C 2 mα obtained
from AP02. The results are compared with direct CFD predictions of the individual damping
coefficients as well as the results obtained from Sacks’ relation using CFD predictions of the
pitch-damping coefficient sum and C 2 mα shown previously in figure 10. The experimentally
derived values of the individual pitch-damping coefficients compare well with the predicted
results. The results also indicate that the biggest source of error is produced by the uncertainty in
the experimentally derived pitch-damping moment coefficient rather than the error in Sacks’
relations. (The pitch-damping data used here are from a highly regarded data set and are
representative of the expected accuracy for the pitch-damping coefficient sum from range
firings.)
10
Direct CFD
0 Direct CFD
Sacks w/Exp and AP02
Sacks w/Exp and AP02
-10
Damping Coefficient
Sacks w/CFD
Sacks w/CFD C mα
-20
-30
C mq
-40
-50
-60
3.5 4 4.5 5 5.5 6
Xcg/D
Figure 14. Comparison of the damping moment coefficients derived from experimental
data or CFD using Sacks’ relations with direct CFD prediction, ANSR, Mach
2.5, and L/D = 9.
Both Sacks (1) and Bryson (2) obtained relations between the pitch-damping coefficients and the
normal force coefficient shown previously in equations 1 and 2. Relatively speaking, equations
1 and 2 do not perform particularly well, especially when compared to Sacks’ relations
18
(equations 4–9). Figure 15 shows a comparison of the CFD prediction of the pitch-damping
coefficient C Nq with results obtained using equation 2 and Sacks’ relation (equation 28). As
before, CFD results were used to obtain the aerodynamic coefficients required in equations 2 and
28. The results obtained with equation 2 provide only a qualitative prediction of the distribution
of C Nq compared with the result obtained using Sacks’ relation (equation 28), which is in much
better agreement with the CFD predictions. Instead of relating the pitch-damping coefficients to
the normal force coefficient as in equation 2, Sacks’ relations relate the pitch-damping
coefficient to each other. These results seem to imply the pitch-damping coefficients are more
closely related to each other than they are to the normal force coefficient. Figure 16 shows a
comparison of the pitch-damping moment coefficient sum C mq + C mα from direct CFD
prediction with the results obtained using Bryson’s relation (equation 1), which like equation 2,
relates the damping coefficient to the normal force coefficient. The distribution of C mq + C mα
is relatively poorly predicted compared with the CFD result, although total moment for entire
configuration differs by only 25%.
15
10
Direct CFD Prediction
Derived from Sacks' Relations
Eq. 2
5
C Nq
0
-5
-10
0 3 6 9
X/D
19
10
-20
-30
-40
-60
-70
0 3 6 9
X/D
Figure 16. Comparison of the pitch-damping moment coefficient sum predicted from
Bryson’s relation (equation 1) with CFD predictions, ANSR, Mach 2.5,
xcg /L = 0.5598 , and L/D = 9.
this configuration (14). For the current report, additional predictions of the individual pitch-
damping coefficients were also performed for flight velocities between Mach 2 and Mach 5.
20
Figure 18 shows the predicted deviations of the moment compared with the individual pitch-
damping coefficients at Mach 2. Again, the deviation over the body is small compared to the
pitch-damping coefficients even on the afterbody where both the geometry and moment
coefficients are changing significantly. Similar results were found at the other flight velocities.
300
200
C 2 mα
100
Coefficients
0
∆1, ∆ 2 , ∆3
C mα
-100
C mq
-200
-300
0 4 8 12 16
X/D
Similar to the results for the secant ogive/cylinder configuration, the form of Sacks’ relations
embodied in equations 25 and 26 was applied to determine the individual pitch-damping
coefficients from the pitch-damping moment coefficient sum and the pitching second-moment
coefficient for the flared projectile geometry. Comparisons were made with direct CFD
predictions.
Figures 19 and 20 show comparisons of the individual pitch-damping moment coefficients at
Mach 2 and 5, respectively, for the flare projectile geometry. Very good agreement between the
results obtained by applying Sacks’ relations and direct CFD predictions are found. Additional
results for the same flare projectile with the 6° flare replaced by a 15° flare were also obtained
and are shown in figure 21. Despite a nearly doubling of the pitch-damping coefficient due to
the larger flare, the Sacks’ relations results are in very good agreement with the direct CFD
predictions.
21
50
C mα
Pitch-Damping Coefficients
-50
-100
C mq
-150
-200
Direct CFD Prediction
Derived from Sacks' Relations
-250
-300
0 4 8 12 16
X/D
Figure 19. Comparison of the damping moment coefficient predicted from Sacks’
relations with CFD predictions, flared projectile, and Mach 2.
50
C mα
Pitch-Damping Coefficients
-50
-100
C mq
-150
-200
Direct CFD Prediction
Derived from Sacks' Relations
-250
-300
0 4 8 12 16
X/D
Figure 20. Comparison of the damping moment coefficient predicted from Sacks’
relations with CFD predictions, flared projectile, and Mach 5.
22
100
C mα
0
-100
-300
C mq
-400
-700
0 4 8 12 16
X/D
Figure 21. Comparison of the damping moment coefficient predicted from Sacks’
relations with CFD predictions, 15° flared projectile, and Mach 5.
4. Summary
In conclusion, the results presented here indicate that Sacks’ pitch-damping relations are only
strictly valid under the context of the theory from which they were originally developed. They
do, however, provide a reasonably good means of estimating the pitch-damping coefficients
when one of the three pitch-damping coefficients can be determined. The most likely practical
use of these relations might be to provide estimates of the individual pitch-damping coefficients
using values of the pitch-damping coefficient sum determined from some other source, such as
experimental data and engineering estimation approaches or when the additional expense of the
separate CFD computations of the individual pitch-damping coefficients is not justified. In some
cases, it appears that the error in applying these relationships is smaller than the error associated
with generating the initial pitch-damping coefficient (such as with engineering estimation
approaches) from which the other two damping coefficients are derived using Sacks’ relations.
Applying Sacks’ relations to determine the individual pitch-damping coefficients from the pitch-
damping sum represents only one possible application of Sacks’ relations. These relations could
also benefit theoretical developments because theories for predicting the pitch-damping
coefficients need only focus on a single damping coefficient. The other damping coefficients
23
could then be obtained from Sacks’ relations. Such an approach has been already used as an
estimation procedure for the damping coefficients (6). In this work, the distribution of
C Nα along the body is predicted using slender body theory with empirically based corrections.
The damping coefficient C Nq can then be obtained from Sacks’ relations. Once the damping
force distributions are known, the damping moments can be easily obtained by integration of the
force loadings. Improvements in the estimates of C Nq from Sacks’ relations can also be
obtained by correlating the error ∆ as well – an approach used in reference (6) to further
improve the estimates of C Nq .
Finally, in the currently report, only axisymmetric configurations in supersonic flight have been
considered. The theory from which Sacks derived the relations considered here is applicable to
both wingless and winged vehicles. Further research is still required to assess the performance
of Sacks’ relations for winged vehicles and for other flight velocity regimes.
24
5. References
1. Sacks, A. H. Aerodynamic Forces, Moments, and Stability Derivatives for Slender Bodies of
General Cross Section; NACA Technical Note 3283; National Advisory Committee for
Aeronautics: Washington, DC, November 1954; p 27.
2. Bryson, A. E., Jr. Stability Derivatives for a Slender Missile With Application to a Wing-
Body-Vertical-Tail Configuration. Journal of the Aeronautical Sciences May 1953, 20 (5),
297–308.
3. Moore, F. G.; Hymer, T. C. The 2002 Version of the Aeroprediction Code: Part I –
Summary of New Theoretical Methodology; NSWCDD/TR-01/108; Naval Surface Warfare
Center: Dahlgren, VA, March 2002.
4. Vukelich, S. R.; Jenkins, J. E. Missile DATCOM: Aerodynamic Prediction on Conventional
Missiles Using Component Build-up Techniques; AIAA Paper 84-0388; American Institute
of Aeronautics and Astronautics: Reston, VA, January 1984.
5. Whyte, R. E. Spinner – A Computer Program for Predicting the Aerodynamic Coefficients
of Spin Stabilized Projectiles; Class 2 Report No. 69APB3; General Electric Company:
Burlington, VA, August 1969.
6. Danberg, J. E.; Weinacht, P. Approximate Computation of Pitch-Damping Coefficients;
AIAA Paper 2002-5048; American Institute of Aeronautics and Astronautics: Reston, VA,
August 2002.
7. Sigal, A. Correlation of the Damping in Pitch Stability Derivatives for Body-Tail
Configurations; AIAA Paper 94-3482; American Institute of Aeronautics and Astronautics:
Reston, VA, August 2002.
8. Weinacht, P.; Sturek, W. B.; Schiff, L. B. Navier-Stokes Predictions of Pitch-Damping for
Axisymmetric Projectiles. Journal of Spacecraft and Rockets 1997, 34 (6), 753–761.
9. Weinacht, P. Navier-Stokes Predictions of the Individual Components of the Pitch-Damping
Sum. Journal of Spacecraft and Rockets 1998, 35 (5), 598–605.
10. Park, S. H.; Kim, Y.; Kwon, J. H. Prediction of Damping Coefficients Using the Unsteady
Euler Equations. Journal of Spacecraft and Rockets 2003, 40 (3), 356–362.
11. Schiff, L. B.; Steger, J. L. Numerical Simulation of Steady Supersonic Viscous Flow. AIAA
Journal 1980, 18 (12), 1421–1430.
25
12. Baldwin, B. S.; Lomax, H. Thin Layer Approximation and Algebraic Model for Separated
Turbulent Flows; AIAA Paper 78-257; American Institute of Aeronautics and Astronautics:
Reston, VA, January 1978.
13. Beam, R.; Warming, R. F. An Implicit Factored Scheme for the Compressible Navier-
Stokes Equations. AIAA Journal 1978, 16 (4), 85–129.
14. Weinacht, P. Navier-Stokes Predictions of Pitch-Damping for a Family of Flared
Projectiles; AIAA Paper 91-3339; American Institute of Aeronautics and Astronautics:
Reston, VA, September 1991.
26
List of Symbols, Abbreviations, and Acronyms
a Speed of sound
M
Cm Pitching moment coefficient,
1
ρ ∞ V 2S ref D
2
∂ Cm
C mα Pitching moment coefficient slope with respect to angle of attack,
∂α
∂ C 2m
C 2 mα Pitching second moment coefficient slope with respect to angle of attack,
∂α
∂ Cm
C mα Pitch-damping moment coefficient slope,
⎛ α D ⎞
∂⎜ ⎟
⎝ V ⎠
∂ Cm
C mq Pitch-damping moment coefficient slope,
⎛qD⎞
∂⎜ ⎟
⎝ V ⎠
C mq + C mα Pitch-damping moment coefficient
F
CN Normal force coefficient,
1
ρ ∞ V 2S ref
2
∂ CN
C Nα Normal force coefficient slope with respect to angle of attack,
∂α
∂ CN
C Nα Pitch-damping force coefficient slope,
⎛ α D ⎞
∂⎜ ⎟
⎝ V ⎠
∂ CN
C Nq Pitch-damping force slope,
⎛qD⎞
∂⎜ ⎟
⎝ V ⎠
C Nq + C Nα Pitch-damping force coefficient
27
D Reference diameter
e Total energy per unit volume
F Force
L Body length
M Mach number
M Moment
p Pressure
r Radial coordinate
Re Reynolds number, a ∞ ρ ∞ D / µ ∞
Ro Helix radius
π D2
S ref Reference area, S ref =
4
V Freestream velocity
X e , Ye , Z e Earth-fixed coordinates
28
Greek Symbols
α Angle of attack
α Angular rate associated with angle of attack
β Yaw angle
ρ Density
φ Circumferential coordinate
Subscripts
∞ Quantity evaluated at freestream conditions
29
NO. OF NO. OF
COPIES ORGANIZATION COPIES ORGANIZATION
1 COMMANDING GENERAL
US ARMY MATERIEL CMD
AMCRDA TF
5001 EISENHOWER AVE
ALEXANDRIA VA 22333-0001
1 US MILITARY ACADEMY
MATH SCI CTR EXCELLENCE
MADN MATH
THAYER HALL
WEST POINT NY 10996-1786
1 DIRECTOR
US ARMY RESEARCH LAB
AMSRD ARL CS IS R
2800 POWDER MILL RD
ADELPHI MD 20783-1197
3 DIRECTOR
US ARMY RESEARCH LAB
AMSRD ARL CI OK TL
2800 POWDER MILL RD
ADELPHI MD 20783-1197
3 DIRECTOR
US ARMY RESEARCH LAB
AMSRD ARL CS IS T
2800 POWDER MILL RD
ADELPHI MD 20783-1197
30
NO. OF NO. OF
COPIES ORGANIZATION COPIES ORGANIZATION
1 COMMANDER ARDEC
AMSTR AR FSF X
W TOLEDO
BLDG 95 SOUTH
PICATINNY ARSENAL NJ 07806-5000
1 US ARMY AMRDEC
AMSAM RD SS AT
R W KRETZSCHMAR
BLDG 5400
REDSTONE ARSENAL AL 35898-5000
1 US ARMY AMRDEC
AMSAM RD SS AT
M E VAUGHN
BLDG 5400
REDSTONE ARSENAL AL 35898-5000
1 DYNETICS INC
M S MILLER
PO BOX 5500
HUNTSVILLE AL 35814-5500
31
NO. OF
COPIES ORGANIZATION
32