ADA582403

Download as pdf or txt
Download as pdf or txt
You are on page 1of 54

Projectile Roll Dynamics and Control With a Low-Cost

Skid-to-Turn Maneuver System

by Frank Fresconi, Ilmars Celmins, Mark Ilg, and James Maley

ARL-TR-6363 March 2013

Approved for public release; distribution is unlimited.


NOTICES

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 21005-5069

ARL-TR-6363 March 2013

Projectile Roll Dynamics and Control With a Low-Cost


Skid-to-Turn Maneuver System

Frank Fresconi, Ilmars Celmins, Mark Ilg, and James Maley


Weapons and Materials Research Directorate, ARL

Approved for public release; distribution is unlimited.


REPORT DOCUMENTATION PAGE Form Approved
OMB No. 0704-0188
Public reporting burden for this collection of information is estimated to average 1 hour per response, including the time for reviewing instructions, searching existing data sources, gathering
and maintaining the data needed, and completing and reviewing the collection information. Send comments regarding this burden estimate or any other aspect of this collection of information,
including suggestions for reducing the burden, to Department of Defense, Washington Headquarters Services, Directorate for Information Operations and Reports (0704-0188), 1215 Jefferson
Davis Highway, Suite 1204, Arlington, VA 22202-4302. Respondents should be aware that notwithstanding any other provision of law, no person shall be subject to any penalty for failing to
comply with a collection of information if it does not display a currently valid OMB control number.
PLEASE DO NOT RETURN YOUR FORM TO THE ABOVE ADDRESS.
1. REPORT DATE (DD-MM-YYYY) 2. REPORT TYPE 3. DATES COVERED (From - To)
March 2013 Final October 2011–September 2012
4. TITLE AND SUBTITLE 5a. CONTRACT NUMBER
Projectile Roll Dynamics and Control With a Low-Cost Skid-to-Turn Maneuver
System 5b. GRANT NUMBER

5c. PROGRAM ELEMENT NUMBER

6. AUTHOR(S) 5d. PROJECT NUMBER


Frank Fresconi, Ilmars Celmins, Mark Ilg, and James Maley AH80
5e. TASK NUMBER

5f. WORK UNIT NUMBER

7. PERFORMING ORGANIZATION NAME(S) AND ADDRESS(ES) 8. PERFORMING ORGANIZATION


REPORT NUMBER
U.S. Army Research Laboratory
ATTN: RDRL-WML-E ARL-TR-6363
Aberdeen Proving Ground, MD 21005-5066
9. SPONSORING/MONITORING AGENCY NAME(S) AND ADDRESS(ES) 10. SPONSOR/MONITOR’S ACRONYM(S)

11. SPONSOR/MONITOR'S REPORT


NUMBER(S)

12. DISTRIBUTION/AVAILABILITY STATEMENT


Approved for public release; distribution is unlimited.

13. SUPPLEMENTARY NOTES

14. ABSTRACT
A low-cost maneuver control system was successfully developed to provide enhanced maneuverability of gun-launched
munitions. The unique environment and aspects of the system are described. Electromechanical design of the maneuver
system was undertaken. Laboratory experiments were conducted to ensure survivability at gun launch and characterize the
response. Aerodynamics and flight mechanics of the projectile roll behavior and actuator dynamics were formulated. These
nonlinear dynamics were expressed in a linear state-space model for controllability, control design, and stability analysis. Roll
response was investigated through nonlinear simulation and wind tunnel experiments. Monte Carlo simulations illustrate
suitable performance of the maneuver control system. Experiments further validate the unique projectile maneuver technology
and approach, reveal nonideal characteristics of the system, and verify modeling and simulation.
15. SUBJECT TERMS
projectile, maneuver, flight control, roll
17. LIMITATION 18. NUMBER 19a. NAME OF RESPONSIBLE PERSON
16. SECURITY CLASSIFICATION OF: OF ABSTRACT OF PAGES
Frank Fresconi
a. REPORT b. ABSTRACT c. THIS PAGE 19b. TELEPHONE NUMBER (Include area code)
Unclassified Unclassified Unclassified UU 54 410-306-0794
Standard Form 298 (Rev. 8/98)
Prescribed by ANSI Std. Z39.18

ii
Contents

List of Figures v

List of Tables vii

1. Introduction 1

2. Low-Cost, Gun-Hard Skid-to-Turn Maneuver System 2


2.1 Maneuver System Design ................................................................................................2
2.2 Experimental Characterization ........................................................................................3
2.2.1 Shock ...................................................................................................................3
2.2.2 Response ..............................................................................................................4

3. Dynamics 5
3.1 Projectile Roll Dynamics.................................................................................................5
3.2 Actuator Dynamics ..........................................................................................................6

4. Control Technique 6
4.1 State Space Formulation and Controllability ..................................................................6
4.2 Linear Quadratic Regulator Algorithm ...........................................................................7
4.3 Stability Analysis ............................................................................................................8

5. Simulation Setup 10

6. Experimental Setup 11
6.1 Airframe, Instrumentation, and Wind Tunnel Facility ..................................................11
6.2 Procedure .......................................................................................................................13

7. Results 13
7.1 Nonlinear Simulation ....................................................................................................13
7.2 Experiment ....................................................................................................................20

8. Conclusions 34

iii
9. References 36

Nomenclature 39

Distribution List 41

iv
List of Figures

Figure 1. Solid model rendering of maneuver system (left) utilizing COTS servomechanism
(right). ........................................................................................................................................3
Figure 2. Maneuver system in shock facility. .................................................................................4
Figure 3. Representative servomechanism in assembly command and response without
aerodynamic loading. .................................................................................................................5
Figure 4. System eigenvalues. ........................................................................................................8
Figure 5. Root locus. .......................................................................................................................9
Figure 6. Bode diagram.................................................................................................................10
Figure 7. Model mounted in wind tunnel with high-speed camera and telemetry data
acquisition. ...............................................................................................................................12
Figure 8. Representative wind tunnel profile. ...............................................................................13
Figure 9. Nominal controlled response without fin cant at zero angle of attack and 10-Hz
initial roll rate – roll angle. ......................................................................................................14
Figure 10. Nominal controlled response without fin cant at zero angle of attack and 10-Hz
initial roll rate – roll rate. .........................................................................................................14
Figure 11. Nominal controlled response without fin cant at zero angle of attack and 10-Hz
initial roll rate – deflections. ....................................................................................................15
Figure 12. Monte Carlo controlled response without fin cant at zero angle of attack and
10-Hz initial roll rate – modulo roll angle histogram at 2 s. ....................................................16
Figure 13. Monte Carlo controlled response without fin cant at zero angle of attack and
10-Hz initial roll rate – roll rate. ..............................................................................................17
Figure 14. Monte Carlo controlled response without fin cant at zero angle of attack and
10 Hz initial roll rate – deflections. .........................................................................................17
Figure 15. Monte Carlo controlled response with 2° fin cant at zero angle of attack and
10-Hz initial roll rate – modulo roll angle histogram at 2 s. ....................................................18
Figure 16. Monte Carlo controlled response with 2° fin cant at zero angle of attack and
10-Hz initial roll rate – roll rate. ..............................................................................................19
Figure 17. Monte Carlo controlled response with 2° fin cant at zero angle of attack and
10-Hz initial roll rate – deflections. .........................................................................................19
Figure 18. Individual canard impulse response without fin cant at zero angle of attack – roll
rate............................................................................................................................................20
Figure 19. Individual canard impulse response without fin cant at zero angle of attack –
deflections. ...............................................................................................................................21
Figure 20. Group canard impulse response with 2° fin cant at zero angle of attack – roll rate. ...22

v
Figure 21. Group canard impulse response with 2° fin cant at zero angle of attack –
deflections. ...............................................................................................................................22
Figure 22. Monte Carlo simulation and experimental controlled response without fin cant at
zero angle of attack at 10 Hz initial roll rate – roll rate. ..........................................................23
Figure 23. Monte Carlo simulation and experimental controlled response without fin cant at
zero angle of attack at 10 Hz initial roll rate – deflections. .....................................................24
Figure 24. Controlled response without fin cant at zero angle of attack at 10-Hz initial roll
rate – roll angle. .......................................................................................................................25
Figure 25. Controlled response without fin cant at zero angle of attack at 10-Hz initial roll
rate – roll rate. ..........................................................................................................................25
Figure 26. Controlled response without fin cant at zero angle of attack at 10-Hz initial roll
rate – deflections. .....................................................................................................................26
Figure 27. Controlled response with 2° fin cant at zero angle of attack at 10-Hz initial roll
rate – roll angle. .......................................................................................................................27
Figure 28. Controlled response with 2° fin cant at zero angle of attack at 10-Hz initial roll
rate – roll rate. ..........................................................................................................................27
Figure 29. Controlled response with 2° fin cant at zero angle of attack at 10-Hz initial roll
rate – deflections. .....................................................................................................................28
Figure 30. Controlled response with 2° fin cant at zero angle of attack at 2-Hz initial roll rate
– roll angle. ..............................................................................................................................29
Figure 31. Controlled response with 2° fin cant at zero angle of attack at 2-Hz initial roll rate
– roll rate. .................................................................................................................................29
Figure 32. Controlled response with 2° fin cant at zero angle of attack at 2-Hz initial roll rate
– deflections. ............................................................................................................................30
Figure 33. Controlled response without fin cant at 5° angle of attack at 10-Hz initial roll rate
– roll angle. ..............................................................................................................................31
Figure 34. Controlled response without fin cant at 5° angle of attack at 10-Hz initial roll rate
– roll rate. .................................................................................................................................31
Figure 35. Controlled response without fin cant at 5° angle of attack at 10-Hz initial roll rate
– deflections. ............................................................................................................................32
Figure 36. Controlled response with 2° fin cant at 5° angle of attack at 10-Hz initial roll rate
– roll angle. ..............................................................................................................................33
Figure 37. Controlled response with 2° fin cant at 5° angle of attack at 10-Hz initial roll rate
– roll rate. .................................................................................................................................33
Figure 38. Controlled response with 2° fin cant at 5° angle of attack at 10-Hz initial roll rate
– deflections. ............................................................................................................................34

vi
List of Tables

Table 1. System properties. ...........................................................................................................11


Table 2. System uncertainties (1 standard deviation). ..................................................................11

vii
INTENTIONALLY LEFT BLANK.

viii
1. Introduction

The overarching motivation for this work is to provide capability for gun-launched munitions to
engage targets that require high maneuverability. Past efforts resulted in precision indirect fire
for stationary targets (1–5). This report represents a component of a program focused on general
development of this enhanced maneuverability for application to relevant systems, such as
artillery, mortars, and shoulder-launched munitions.

The approach to achieve a low-cost, gun-hard skid-to-turn maneuver control system outlined in
this report is to leverage commercial off-the-shelf (COTS) technology for the high-G
environment and develop control algorithms that rely on a high-fidelity characterization of the
aerodynamics and flight mechanics. A variety of state-of-the-art actuation technologies (6–8)
may be considered for this application. Servomechanisms used by hobbyists to fly remote-
controlled aircraft are low cost because of the high-volume manufacturing. One outstanding
question is whether these devices perform to the standard necessary for guided projectile
applications. Components must withstand the high structural loadings during gun launch (9, 10).
Hardening components, such as a COTS servomechanism for gun launch, consist of properly
supporting structures, minimizing mass, and reducing the number of moving parts.

Projectile aerodynamics and flight dynamics can be well understood by applying mature
computational and experimental methods developed specifically for projectiles. Computational
fluid dynamics techniques have been applied to complex projectile aerodynamic phenomena
with much success (11–14). Spark ranges, in existence since the 1940s, provide extremely
accurate aerodynamic coefficients (15–17) due to the free-flight nature and motion measurement
accuracy of the experiment. Experimental techniques have enjoyed recent advancement with the
onboard sensor technique (18).
A canted fin-stabilized, canard-controlled projectile flying in a skid-to-turn configuration was
selected based on past work (19), which complicates the aerodynamic characterization resulting
from flow interactions (18, 20, 21). Packaging constraints for gun tube launch require deploying
stabilizing and control surfaces. Spin induced by gun rifling or fin cant improves ballistic launch
accuracy but requires a consideration of roll-yaw resonance (22) for portions of the flight. A
rocket motor, used for most missiles, is not currently a part of this system, therefore only the gun
propellant may be used to supply the dynamic pressure necessary for target engagements that
require high maneuverability. The skid-to-turn configuration was chosen for the high bandwidth
since the time of flight for some applications of interest may be only 1 s. Quantifying the flight
dynamics of this complex system through computational and experimental techniques is essential
to developing and implementing simple, effective control solutions.

1
Classical and modern control techniques have been applied to the roll control problem for
missiles (23–26). Coupled roll-pitch-yaw control of bank- and skid-to-turn maneuver systems
have also been considered (21, 27–29). The approach herein is to formulate these techniques in
the simplest manner possible for the unique projectile environment by including effects like fin
cant and canard-fin interactions. Additionally, flexible body dynamics for missiles have been
shown in simulation to cause instability; however, these effects may be ignored for this
application since the flight body has been hardened to withstand gun launch. While missiles
often integrate high-performance gyroscopes given a transfer alignment at separation to
determine roll feedback, the harsh gun launch environment precludes such an arrangement.
Alternate solutions, based on low-cost magnetometer or thermopile measurements, have been
formulated (30–33).

The goal of this report is to develop and demonstrate roll control performance of a low-cost, gun-
hard skid-to-turn maneuver system. The contribution of this work lies in leveraging COTS
devices to reduce cost, ensuring structural integrity, and using high-fidelity flight
characterization methods in the control strategy for the gun launch environment. A wide variety
of technical disciplines have been successfully synthesized to develop this technology using
theory, nonlinear simulation, and experiment.

The low-cost, gun-hard maneuver control system design and experimental characterization is
presented in this report. Next, the dynamics of the projectile flight and actuation are given along
with the state space construction, controllability, control design, and stability analysis. Setup and
results of the nonlinear simulations and wind tunnel experiments are provided, followed by
summary findings and conclusions.

2. Low-Cost, Gun-Hard Skid-to-Turn Maneuver System

2.1 Maneuver System Design


A generic class of fin-stabilized, canard-controlled munitions in the 60- to 155-mm-diameter
range (mortars, artillery, shoulder-launched munitions, air-dropped munitions, etc.) may utilize
this maneuver control system technology. The mechanical design of the maneuver system
internally mounted near the nose and an image of the COTS servomechanism (Hitec
HS-5056MG) are shown in figure 1. Four servomechanisms are each connected through a
pinned arm to a control surface. Proper design of this linkage provides the required deflections
of the control surfaces. Mechanical limits of ±10° were placed on the deflections.

The servomechanism consists of potentiometer feedback, processor for a motor control


algorithm, driving electronics, and a brushed DC motor. This closed system is driven by an
outside pulse-width modulated (PWM) signal that prescribes the angle of the output shaft of the
motor and operates at a rate of 50 Hz.

2
Figure 1. Solid model rendering of maneuver system (left) utilizing COTS servomechanism (right).

A digital signal processor (DSP) (Texas Instruments TMS320F28335) sends PWM commands to
the servomechanism based on the embedded roll control algorithm. An analog-to-digital
converter samples measurements for the roll feedback and potentiometer from each
servomechanism. Diagnostics are also packaged on the DSP into a pulse-code modulated stream
and sent to an S-band transmitter and antenna for data acquisition during experiments. Lithium
ion batteries power the maneuver system.

As a cost reference point, components necessary for the present maneuver system are a couple
hundred dollars, which is approximately a twofold decrease in the order of magnitude from some
currently fielded precision munitions in relatively low volume (thousand units per year).

2.2 Experimental Characterization

2.2.1 Shock
Experiments were conducted to initially assess the survivability of the maneuver system during
gun launch. Assemblies of the servomechanisms, housings, and linkages were built and tested
for functionality by running the motors. These test articles were mounted as shown in figure 2 in
a shock facility. The experimental shock setup is a plate instrumented with a reference
accelerometer and the article under test, which is constrained to one-dimensional translation.
The instrumented plate is accelerated from rest by releasing an elastic cable and undergoes a
short period (usually order of <1 ms) of high acceleration at impact, which can be shaped by
dampers.

3
Figure 2. Maneuver system in shock facility.

The maneuver system was subjected to loadings up to 10,000 Gs. Post-shock inspection and
functionality did not reveal any failures in the assembly. The acceleration limit for survivability
was not found since the assemblies were tested successfully to 10,000 Gs (which is on the order
of most mortar and artillery launches) and a specific gun-projectile combination were not of
interest for this study.

2.2.2 Response
Simple experiments were performed on the bench in the laboratory on the maneuver system to
identify salient features for modeling the system. Step commands in 1° increments from –9° to
+9° were provided to each servomechanism, and the response was measured through the
potentiometer. Calibrations were performed on an optical bench with a laser reflection
measurement technique to relate potentiometer voltage to control surface deflection angle. An
example of the step response is provided in figure 3. Analysis of these data illustrates that the
significant actuator dynamics can be modeled as a first-order system with a delay. Major
phenomena that these experiments do not capture are the aerodynamic loading and coupling of
the projectile flight dynamics on the maneuver system.

4
Figure 3. Representative servomechanism in assembly command and response
without aerodynamic loading.

3. Dynamics

3.1 Projectile Roll Dynamics


Applying Newtonian kinetics yields the equation of motion for the projectile roll dynamics. The
rate of change of the angular momentum is equal to the sum of the moments about the roll axis.

. (1)

Aerodynamic moments include static roll torque due to canted fins, deflected canards, and roll
damping. Proper modeling and characterization of the aerodynamics are critical to developing a
low-cost skid-to-turn maneuver system. The static roll torque coefficient is often a complex
function of cant or bevel angle, Mach number, total angle of attack, aerodynamic roll angle, and
canard deflection to account for the flow interactions of fins with the body and canards. Roll
damping coefficient is a function of Mach number. Center-of-pressure (CP) of the canards can
be a function of Mach number and angle of attack. Canard normal force coefficient is usually
nonlinear with angle of attack and Mach number; the deflection angle is also sometimes used in
the aerodynamic model. An odd-order Taylor series expansion up to the fifth order in angle of
attack with look-up tables for the Mach variation models the canard normal force nonlinearities.

5
. (2)

Here, the local angle of attack at each blade is a summation of the deflection angle and the flow
velocity at each canard in the canard frame.

. (3)

Components of the flow velocity in the local canard frame are obtained by finding the velocity at
each canard in the projectile body frame.
. (4)

The transformation matrix from the projectile body frame to each canard frame is applied to
calculate the velocity at each canard in the canard frame.
. (5)

3.2 Actuator Dynamics


Inspection of the benchtop experimental response shows that the actuator dynamics can be
described using a first-order system with a delay. Effects such as oscillation, overshoot,
deadbanding, backlash, or hysteresis were not included. An open question that this research
addresses is whether actuation technology shortfalls may be overcome through a systems
approach. The model for the first-order system with time constant is given in the following
equation. The time delay was incorporated by simply delaying the deflection commands by a
prescribed amount of time.

. (6)

4. Control Technique

4.1 State Space Formulation and Controllability


The projectile roll and actuator dynamics can be packaged into a state space representation to
apply an array of control techniques. The states for this problem are roll, roll rate, and the four
canard deflections.

. (7)

The controls are the deflection commands for the four canards.
. (8)

6
By linearizing the aerodynamic model and re-expressing the equations of motion, we obtained
the system dynamics matrix.

. (9)

The time constant of each canard may be unique. The controls matrix has the following form.

. (10)

Additionally, the static roll torque adds a forcing function.

. (11)

The system is controllable, as the rank of the controllability matrix


(Co  [ B AB A2 B A3B A4 B A5 B]) is equal to the number of states.

4.2 Linear Quadratic Regulator Algorithm


The linear quadratic regulator expresses the control as . The weightings for control
effort and control error enable proper tuning of the controller for a given application. The gain is
given by the following expression.

. (12)

The matrix is calculated by solving the matrix Riccatti equation.

7
. (13)

The error signal used in the control law is formed by differencing the feedback from the desired
states. The airframe flies in an “X” configuration for optimal maneuverability. For reasons of
symmetry, the desired roll angle can be any of four locations, which are 90° apart in roll. A
trapezoidal error signal was formed using these symmetry locations, and the roll feedback to
drive the roll angle to the nearest symmetry location. Roll rate feedback was regulated to zero.
While canard deflection feedback is available, after some analysis it was decided not to use this
information for closed loop control for reasons of increased cost of calibration, processing, and
reliability necessary for a sufficiently clean feedback signal. Additionally, feed forward gain
may be used when fins are canted since the effectiveness of both the fins and canards and fin
cant angle is known. Saturation occurred when the commands were greater than the mechanical
limits of ±10°.

4.3 Stability Analysis


The linear state space representation of the nonlinear projectile roll and actuator dynamics was
used to conduct a stability analysis during control design. Recall that nonlinearities in the
aerodynamics, actuator delay, and deflection limiter were ignored for this analysis.

Eigenvalues for roll and roll rate are presented in figure 4. The uncontrolled system has a
marginally unstable eigenvalue. Comparing the uncontrolled and controlled cases shows that
damping is increased and oscillations are induced for the controlled system.

Figure 4. System eigenvalues.

8
The root locus analysis illustrates the behavior of the roll and roll rate as the control gain is
changed. Figure 5 shows that the roll response increases in frequency as the gain increases. As
gain increases to about 3, the rolling motion becomes unstable before again becoming stable at
even higher values of gain (above 50). Roll rate stability increases as gain increases with a
minor decrease in frequency. After a gain of about 0.05, the dominant roll rate mode shifts, and
roll rate behavior is unstable for high gain.

Figure 5. Root locus.

The Bode plot in figure 6 shows the controlled response decline beyond 2 Hz for roll and at
about 8 Hz for roll rate. Roll features a gain margin of about 4.6 dB and phase margin of about
42°. The Bode plot illustrates that the controlled roll rate characteristics are excellent; –180° in
phase is never crossed and the phase margin is about 40°.

9
Figure 6. Bode diagram.

5. Simulation Setup

The nonlinear dynamics were implemented in a simulation environment to assess full-spectrum


system performance. A fourth-order Runge-Kutta scheme was used to integrate the equations of
motion. System properties are provided in table 1. The Mach number of interest was 0.49. The
axial moment of inertia was obtained by solid modeling and measurement. Semi-empirical
aeroprediction in the PRODAS (34) Finner code and parameter identification using Bayesian
filter error methods from wind tunnel data were applied to estimate the aerodynamic coefficients.
Actuator time constant and delays came from benchtop laboratory and wind tunnel data. System
properties in table 1 were also used for the stability analysis.

Monte Carlo simulations were performed. System properties were varied in each Monte Carlo
replication according to the uncertainties given in table 2. Bias and random errors were also
added to the canard deflections, roll feedback, and roll rate feedback.

10
Table 1. System properties.

Property Value Unit


0.49 —
0.08 m
2
0.0044 kg-m
–6.76 —
0.82 cal.
0.674 —
0.0263 —
0.015 s
0.030 s
0.0225 —
0.0021 —

Table 2. System uncertainties (1 standard deviation).

Property Value Unit


0.1 %
10 %
30 %
15 %
30 %
30 %
20 %
20 %
1 deg
1 deg
1 deg
0.1 Hz
0.1 Hz

6. Experimental Setup

6.1 Airframe, Instrumentation, and Wind Tunnel Facility


Experiments were conducted on the low-cost, gun-hard maneuver control system in a wind
tunnel. Figure 7 shows an airframe that featured a hemispherical nose, increasing body diameter
with length toward the base (to 80 mm maximum), and six fins. Détentes in the fins allowed
cant angles of 0° and 2° to examine the effect of static roll torque. The majority of the maneuver
system was internally mounted in the ogive of the airframe. The DSP, batteries, and S-band

11
Figure 7. Model mounted in wind tunnel with high-speed camera and telemetry data acquisition.

transmitter and antenna were contained in the midbody of the airframe. The airframe was
mounted to a sting in the wind tunnel and free to rotate about the longitudinal axis through a
bearing.

In this experimental setup, bearing friction, which is not present in free flight, adds another
moment to the roll dynamics. The experimental model was spun in the lab to estimate the
moment due to bearing friction. A simple linear model of the form L  sgn   F was used to
obtain an experimentally derived value of –0.0604 Nm for the bearing friction moment.

An optical encoder with 8192 counts per revolution mounted in the base of the airframe
measured the roll angle of the body with respect to the sting (wind tunnel). This measurement
was differentiated on the DSP to estimate roll rate.

A telemetry ground station acquired the data transmitted from the wind tunnel model. High-
speed photography recorded the motion of the airframe in the wind tunnel. The telemetry ground
station and helical antenna and high-speed camera are evident in the background of figure 7.

The wind tunnel was located at the Edgewood Area of Aberdeen Proving Ground, MD. This
tunnel features a 20- × 20-in cross section and can accommodate Mach numbers from about 0.5
to 1.2. A Mach of 0.48–0.49 was used for experiments. Figure 8 shows a typical velocity profile
as the tunnel ramps up to steady state Mach number.

12
Figure 8. Representative wind tunnel profile.

6.2 Procedure
The roll control commenced during experiments based on the airframe reaching a roll rate
threshold. The roll rate threshold was achieved either by deflecting canards or canting fins and
starting the wind tunnel. For a given experiment, the roll rate threshold value and initial canard
deflection were programmed into the DSP through a laptop with the model sitting on a bench.
The model was then mounted in the wind tunnel, and the batteries were connected. As the wind
tunnel blow began, the model spun to the roll rate threshold and then performed controlled
deflections; telemetry and high-speed camera data were recorded. After the wind tunnel blow
ended, the model was removed from the tunnel, power was turned off, data was reviewed, and
the cycle repeated for the next experiment.

7. Results

7.1 Nonlinear Simulation


Simulations were performed according to the setup described earlier. Some nominal responses
for an initial roll rate of 10 Hz, zero angle of attack, and no fin cant are shown in figures 9–11.
The projectile body completes one revolution prior to locking into the “X” configuration at
= 405° (alternately, 360N + [45 135 225 315] degree roll could have been the symmetry
point) with a steady-state roll error of <1°. Roll rate decreases to zero by about 0.3 s. Some
overshoot and oscillation is apparent in the roll and roll rate, but using the flight and actuation
model in the control algorithm provides sufficient performance for a nominal simulation.

13
Figure 9. Nominal controlled response without fin cant at zero angle of attack and
10-Hz initial roll rate – roll angle.

Figure 10. Nominal controlled response without fin cant at zero angle of attack and
10-Hz initial roll rate – roll rate.

14
Figure 11. Nominal controlled response without fin cant at zero angle of attack and
10-Hz initial roll rate – deflections.

The deflection history for all four canards for the nominal simulation is presented in figure 11.
Canards deflect in the negative direction to decrease the initially positive roll rate. The
difference between the commands and modeled response is due to the delay and lag. Sample and
hold at 50 Hz is also present in the data.

Monte Carlo simulation results for 500 samples at zero angle of attack are provided in the next
six figures. For these simulations, initial roll was randomized between 0° and 360° according to
a uniform random number generator, and initial roll rate was randomly drawn from a normal
distribution with a standard deviation of 10 Hz. All other parameters were randomized with a
normal distribution, with mean and standard deviation provided in tables 1 and 2. Each canard
featured unique values for lag, delay, and bias.

Histograms of the roll angle manipulated by a modulo at 2 s are given in figure 12. Roll
symmetry points of 45°, 135°, 225°, and 315° are dashed red vertical lines. Many simulations
are grouped around the 45° and 225° symmetry points with a few distributed about the 135° and
315° roll locations. The reason behind these groupings is likely the interaction of the initial roll
rate and the roll control effectiveness, as seen for the nominal simulation responses. Essentially,
the roll rate is controlled at a time where 45° or 225° is the closest roll symmetry location.

15
Figure 12. Monte Carlo controlled response without fin cant at zero angle of attack
and 10-Hz initial roll rate – modulo roll angle histogram at 2 s.

Individual Monte Carlo samples for the roll rate and deflections are shown in figures 13 and 14.
The dashed red lines represent plus and minus one standard deviation about the mean of all
simulations. The initial roll rate is reduced to <1 Hz by about 0.3 s. The deflections show a
region close to zero time with small deflection due to the delay and lag, followed by a moderate
deflection (standard deviation of about 5°) for a few hundred milliseconds, then a near-constant
value as the desired control is achieved. Bias errors in the roll rate and deflections are illustrated
in these results.

16
Figure 13. Monte Carlo controlled response without fin cant at zero angle of attack
and 10-Hz initial roll rate – roll rate.

Figure 14. Monte Carlo controlled response without fin cant at zero angle of attack
and 10 Hz initial roll rate – deflections.

17
Monte Carlo analysis was conducted for zero angle of attack with fin cant. These results, given
in figures 15–17, are similar to the no fin cant case. The roll histograms are grouped at 45° and
225°. The roll rate falls below 1 Hz statistically within about 0.3 s. Deflections peak within the
first few hundred milliseconds and feature an offset on average of about –1° to counteract the fin
roll torque. A feedforward gain resulting in a deflection command of –0.9° was used in these
simulations. Results did not change appreciably without the feedforward gain.

Figure 15. Monte Carlo controlled response with 2° fin cant at zero angle of attack
and 10-Hz initial roll rate – modulo roll angle histogram at 2 s.

18
Figure 16. Monte Carlo controlled response with 2° fin cant at zero angle of attack
and 10-Hz initial roll rate – roll rate.

Figure 17. Monte Carlo controlled response with 2° fin cant at zero angle of attack
and 10-Hz initial roll rate – deflections.

19
7.2 Experiment
Wind tunnel experiments were performed according to the setup and procedure outlined
previously. Scatter in the experimental data is due to bit errors in the telemetry. Initial
experiments were designed to assess the open-loop performance of the maneuver system in the
wind tunnel. Maneuver system properties (e.g., bias, linearity), aerodynamic estimates, flow
conditioning, and accuracy of model fixturing in the tunnel were investigated during these initial
experiments.

One series of experiments to address these phenomena swept each canard through a sequence of
angles every 800 ms. This method is illustrated by examining figures 18 and 19 in tandem. As
the tunnel started, canard 1 was deflected to 4°. Once roll rate exceeded 2 Hz, canard 1 changed
value to 0°, –4°, 4°, –8°, 8° with all other canards held to zero. After canard 1 finished sweeping
through these commands, canard 2 began an equivalent sweep. This process was repeated until
all canards were cycled. The wind stopped just as canard 4 began deflections so these data were
not captured. Bias in the roll rate is apparent in these results; –4° and –8° deflections yield about
a –2 and –4 Hz roll rate, respectively, while 4° and 8° deflections produce 2.5 and almost 5 Hz
roll rate, respectively. With just the wind tunnel data, it is indeterminate whether this is due to
some mix of maneuver system bias or nonlinearity or some cross-flow due to flow conditioning
or model fixturing. Unsteady flow effects are also present. Regardless, these measurements
provide experimental evidence of nonideal response that the maneuver control system must
sufficiently address.

Figure 18. Individual canard impulse response without fin cant at zero angle of attack
– roll rate.

20
Figure 19. Individual canard impulse response without fin cant at zero angle of attack
– deflections.

Another set of experiments was designed to determine maneuver system response in the tunnel at
0° angle of attack with 2° fin cant and collective deflection. These data are provided in figures
20 and 21. Fins were used to increase the roll rate as the wind commenced. Once the 1-Hz roll
rate was reached, all canards deflected to 0°, –0.5°, –1°, –1.5°, and –2° at 800-ms intervals.
Deflecting all canards to –0.5° decreased the roll rate. As the deflection angle changes to –1°,
the canard and fin roll torques are nearly equal and opposite. Friction in the roll bearing
becomes more important as the roll rate approaches zero. The precise characterization of bearing
friction is likely highly nonlinear near zero rotation speed and was not undertaken. Bearing
friction is not acting during free-flight, and this phenomenon adds difficulty to the problem of
controlling roll in the wind tunnel, as at very low rotational speeds, overcoming the bearing
friction becomes more important. Increasing the deflection to –1.5° yields a negative roll rate,
and –2° deflection increases the negative roll rate. As another cycle of deflections began,
however, the wind stopped at about 6–7 s. This experiment provides a value for canard
deflection necessary to overcome the bearing friction at zero roll rate and also the deflection that
may be used in a feedforward gain construction during control with fin cant.

21
Figure 20. Group canard impulse response with 2° fin cant at zero angle of attack –
roll rate.

Figure 21. Group canard impulse response with 2° fin cant at zero angle of attack –
deflections.

22
Control was performed in the wind tunnel. These data provided experimental support for the
performance of the maneuver control system and an opportunity to validate the modeling and
simulation. Figures 22 and 23 provide the experimental data for zero angle of attack and no fin
cant and the corresponding Monte Carlo analysis. Here, the canards were initially deflected to
4°, and control began once the wind started and the roll rate exceeded 10 Hz. In the experiment,
the roll was controlled within about 0.3 s, similar to previous simulations. Roll rate and
deflection response lie within the 1 standard deviation bounds of the Monte Carlo simulations for
the majority of the time. Differences in the roll behavior (and subsequent deflection commands
using roll and roll rate feedback) that occur as the roll rate passes through zero are likely due to
nonlinear bearing friction or unsteady, nonlinear aerodynamics. Excellent agreement between
experiment and simulation will only improve further as aerodynamic and maneuver system
characterizations mature.

Figure 22. Monte Carlo simulation and experimental controlled response without fin
cant at zero angle of attack at 10 Hz initial roll rate – roll rate.

23
Figure 23. Monte Carlo simulation and experimental controlled response without fin
cant at zero angle of attack at 10 Hz initial roll rate – deflections.

Experiments were repeated at the same conditions (zero angle of attack, no fin cant, 10-Hz initial
roll rate). These results are presented in figures 24–26. The roll angle increased prior to locking
into the nearest symmetry point (1305°) with about a 37° error. Roll rate decreased with some
oscillation near zero. The concomitant deflections show that the demand after the roll rate is
zero is almost 1° because of the roll angle error; however, the static bearing friction is too high to
roll the model in the wind tunnel. This response agrees with the collective deflection open-loop
experiments, where a deflection of 1.5° was required to break the bearing friction from rest. This
phenomenon is not present in free-flight.

24
Figure 24. Controlled response without fin cant at zero angle of attack at 10-Hz initial
roll rate – roll angle.

Figure 25. Controlled response without fin cant at zero angle of attack at 10-Hz initial
roll rate – roll rate.

25
Figure 26. Controlled response without fin cant at zero angle of attack at 10-Hz initial
roll rate – deflections.

Fin cant of 2° was emplaced on the model, and controlled experiments were executed. For the
results in figures 27–29, a feedforward gain yielding a deflection of –0.8° was used to combat the
static roll torque of the fins. Experimental roll was within 1° of the desired point (1125°), and
roll rate was within 1 Hz of zero within about 200 ms from start of control. The –0.8° deflection
offset for longer times was to counteract the fin cant. The long (>2 s) drop-out in data was due
to the transmitting telemetry antenna facing away from the receiving antenna once roll control
started.

26
Figure 27. Controlled response with 2° fin cant at zero angle of attack at 10-Hz initial
roll rate – roll angle.

Figure 28. Controlled response with 2° fin cant at zero angle of attack at 10-Hz initial
roll rate – roll rate.

27
Figure 29. Controlled response with 2° fin cant at zero angle of attack at 10-Hz initial
roll rate – deflections.

The effect of initial roll rates was studied by keeping the angle of attack at 0°, fin cant to 2°, and
setting the threshold for control to start at 2 Hz. The roll, roll rate, and deflections for this case
are shown in figures 30–32. As expected, the response was improved for lower initial roll rate.
Again, the feedforward gain was used, and the roll was <2° from the symmetry point (225°) by
the end of the blow. Some slow transient in roll response was evident; there was about 15° of
error when roll rate was regulated to zero. Deflection histories illustrate that less control demand
is required to sufficiently control the model when initial conditions are closer to the desired
states.

28
Figure 30. Controlled response with 2° fin cant at zero angle of attack at 2-Hz initial
roll rate – roll angle.

Figure 31. Controlled response with 2° fin cant at zero angle of attack at 2-Hz initial
roll rate – roll rate.

29
Figure 32. Controlled response with 2° fin cant at zero angle of attack at 2-Hz initial
roll rate – deflections.

It is well known that the aero-mechanics of projectiles, especially while spinning, can be highly
nonlinear with angle of attack. Wind tunnel experiments were conducted to assess the
performance of this maneuver control system at angle of attack. The data in figures 33–35 were
taken with the model at a 5° angle of attack and no fin cant. Roll is regulated to within 31° of the
desired angle (1125°). Roll rate response and deflection histories are similar to previous
experimental results. The roll rate response beyond the threshold for control to start and
subsequent deflection commands and roll rate overshoot were slightly more dramatic since initial
deflections were 8° to account for aerodynamic roll angle-dependent lifting surface effectiveness.
Deflections were almost a degree after the roll rate was controlled to zero because of remaining
roll error. Again, the bearing friction near roll rate zero is likely the culprit for the steady-state
roll behavior.

30
Figure 33. Controlled response without fin cant at 5° angle of attack at 10-Hz initial
roll rate – roll angle.

Figure 34. Controlled response without fin cant at 5° angle of attack at 10-Hz initial
roll rate – roll rate.

31
Figure 35. Controlled response without fin cant at 5° angle of attack at 10-Hz initial
roll rate – deflections.

The final fundamental phenomenon under investigation was the effect of angle of attack with fin
cant. These data are given in figures 36–38. The feedforward gain was not used in this
experiment. Roll was regulated to about 31° of the desired angle (1305°). Roll and roll rate
oscillated a few hundred milliseconds after control started, potentially because of unsteady flow
interactions of canards and fins. Steady-state deflection commands exist because of the steady-
state error in roll.

32
Figure 36. Controlled response with 2° fin cant at 5° angle of attack at 10-Hz initial
roll rate – roll angle.

Figure 37. Controlled response with 2° fin cant at 5° angle of attack at 10-Hz initial
roll rate – roll rate.

33
Figure 38. Controlled response with 2° fin cant at 5° angle of attack at 10-Hz initial
roll rate – deflections.

8. Conclusions

The motivation for this effort is a new capability of affordable interception of targets requiring
high maneuverability from gun-launched systems, which represents an enormous leap-ahead in
lethal technology. The novel low-cost, gun-hard maneuver control system successfully
developed and outlined in this report supports this goal. These technologies are applicable to a
generic class of munitions from 60 to 155 mm in diameter. This system flies in a skid-to-turn
arrangement, may feature time of flight as short as 1 s, does not have a rocket motor, and is gun
hard. The approach to achieve this novel system was to use a systems approach to relax
subsystem interface requirements that enabled the application of COTS devices. The approach
also relied on a high-fidelity aero-mechanics characterization in the control scheme.

The mechatronics of the maneuver system was provided. The suitability of this design for
survival at gun launch was assessed through laboratory shock experiments. Modeling of the
actuation system was underpinned through benchtop experiments.

Dynamics of the projectile roll and actuator were derived from first principles. The
aerodynamics are a major contributor to the system response; therefore, detailed modeling was
undertaken. The dynamics were linearized and a formal controllability analysis illustrated that

34
the system was formulated properly. The linear quadratic regulator was applied for this unique
system. Stability analysis evaluated closed-loop behavior and helped select optimal control
gains.

Nonlinear simulations and wind tunnel experiments were described and conducted. The effects
of initial roll rate, fin cant, and angle-of-attack phenomena were investigated. These results
demonstrated that this maneuver control system achieves the required performance over the
necessary conditions. Monte Carlo simulations showed satisfactory control within about 0.3 s
over a wide range of variation in system parameters, such as mass properties, aerodynamics,
actuation (delay, lag, and bias), and feedback (roll, roll rate). Wind tunnel experiments verified
the modeling and simulation and further validated the low-cost, gun-hard maneuver control
system. Open-loop experiments illustrated the non-ideal performance of the COTS-based
maneuver system; however, the closed-loop results indicated that using appropriate dynamic
modeling in the control scheme with a systems approach led to successful control. After all
experimental results were reviewed, it appears that steady-state roll errors in the wind tunnel are
driven by bearing friction, as cases with fin cant and zero angle of attack (i.e., when a steady,
constant force other than the canards is available) feature roll error within 1° to 2°. While
bearing friction in the wind tunnel rig led to reduced roll response, this phenomenon is not
present in the relevant environment and Monte Carlo simulations of free-flight suggest sufficient
performance.

Future efforts focus on improved aerodynamic modeling based on computational and


experimental techniques. These data are essential to synthesizing the simplest, most effective
controller. Pitch and yaw control are necessary for lateral maneuvers to the target. Integration
with the navigation technology must be undertaken to achieve target engagements requiring high
maneuverability from a gun.

35
9. References

1. Morrison, P. H.; Amberntson, D. S. Guidance and Control of a Cannon-Launched Guided


Projectile. Journal of Spacecraft and Rockets 1977, 14 (6), 328–334.

2. Grubb, N. D.; Belcher, M. W. Excalibur: New Precision Engagement Asset in the Warfight.
Fires October–December 2008, 14–15.

3. Moorhead, J. S. Precision Guidance Kits (PGKs): Improving the Accuracy of Conventional


Cannon Rounds. Field Artillery January–February 2007, 31–33.

4. Fresconi, F.; Brown, T.; Celmins, I.; DeSpirito, J.; Ilg, M.; Maley, J.; Magnotti, P.; Scanlan,
A.; Stout, C.; Vazquez, E. Very Affordable Precision Projectile System and Flight
Experiments. Proceedings of the 27th Army Science Conference, Orlando, FL, 2010.

5. Fresconi, F. E. Guidance and Control of a Projectile With Reduced Sensor and Actuator
Requirements. Journal of Guidance, Control, and Dynamics 2011, 34 (6), 1757–1766.

6. Habibi, S.; Roach, J.; Luecke, G. Inner-Loop Control for Electromechanical Flight Surface
Actuation Systems. Journal of Dynamic Systems, Measurement, and Control 2008, 130,
051002-1–051002-13.

7. Mani, S.; Sahjendra, S. N.; Parimi, S. K.; Yim, W. Adaptive Servoregulation of a Projectile
Fin Using Piezoelectric Actuator. Journal of Dynamic Systems, Measurement, and Control
2007, 129, 100–104.

8. Rabinovitch, O.; Vinson, J.; On the Design of Piezoelectric Smart Fins for Flight Vehicles.
Smart Materials and Structures 2003, 12, 686–695.

9. Mermagen, W. H. High-g Resistant Electronic Fuse for Projectile Payloads. Journal of


Spacecraft and Rockets 1971, 8 (8), 900–903.

10. Brown, T. G.; Davis, B.; Hepner, D.; Faust, J.; Myers, C.; Muller, P.; Harkins, T.; Hollis, M.;
Miller, C.; Placzankis, B. Strap-Down Microelectromechanical (MEMS) Sensors for High-
G Munition Applications. IEEE Transactions on Magnetics 2001, 37 (1), 336–342.

11. DeSpirito, J.; Silton, S.; Weinacht, P. Navier-Stokes Predictions of Dynamic Stability
Derivatives: Evaluation of Steady-State Methods. Journal of Spacecraft and Rockets 2009,
46 (6), 1142–1154.

12. Costello, M.; Sahu, J. Using Computational Fluid Dynamic/Rigid Body Dynamic Results to
Generate Aerodynamic Models for Projectile Flight Simulation. Journal of Aerospace
Engineering 2008, 22 (G7), 1067–1079.

36
13. Sahu, J. Time-Accurate Numerical Prediction of Free-Flight Aerodynamics of a Finned
Projectile. Journal of Spacecraft and Rockets 2008, 45 (5), 946–954.

14. Weinacht, P. Projectile Performance, Stability, and Free-Flight Motion Prediction Using
Computational Fluid Dynamics. Journal of Spacecraft and Rockets 2004, 41 (2), 257–263.

15. Murphy, C. H. Free Flight Motion of Symmetric Missiles; report 1216; U.S. Army Ballistics
Research Laboratory: Aberdeen Proving Ground, MD, July 1963.

16. McCoy, R. L. Modern Exterior Ballistics; Schiffer Publishing Ltd.: Atlen, PA, 1999.

17. Whyte, R. H.; Winchenbach, G. L.; Hathaway, W. H. Subsonic Free-Flight Data for a
Complex Asymmetric Missile. Journal of Guidance and Control 1981, 4 (1), 59–65.

18. Fresconi, F. E. Experimental Flight Characterization of Asymmetric and Maneuvering


Projectiles From Elevated Gun Firings. Journal of Spacecraft and Rockets, accepted for
publication, 2012.

19. Fresconi, F.; Celmins, I.; Fairfax, L. Optimal Parameters for Maneuverability of Affordable
Precision Projectiles. AIAA Aerospace Sciences Meeting, Orlando, FL, January 2012.

20. Moore, F. G.; Moore, L. Y. Approximate Method to Calculate Nonlinear Rolling Moment
due to Differential Fin Deflection. Journal of Spacecraft and Rockets 2012, 49 (2),
250–260.

21. Wise, K. A.; Broy, D. J. Agile Missile Dynamics and Control. Journal of Guidance
Control, and Dynamics 1998, 21 (3), 441–449.

22. Pepitone, T. R.; Jacobson, I. D. Resonant Behavoir of a Symmetric Missile Having Roll
Orientation-Dependent Aerodynamics. Journal of Guidance and Control 1978, 1 (5),
335–339.

23. Nesline, F. W.; Wells, B. H.; Zarchan, P. Combined Optimal/Classic Approach to Robust
Missile Autopilot Design. Journal of Guidance, Control, and Dynamics 1981, 4 (3),
316–322.

24. Nesline, F. W.; Zarchan, P. Why Modern Controllers Can Go Unstable in Practice. Journal
of Guidance, Control, and Dynamics 1984, 7 (4), 495–500.

25. Ohta, H.; Kakinuma, M.; Nikiforuk, P. N. Use of Negative Weights in Linear Quadratic
Regulator Synthesis. Journal of Guidance 1991, 14 (4), 791–796.

26. Talole, S. E.; Godbole, A. A.; Kohle, J. P.; Phadke, S. B. Robust Roll Autopilot Design for
Tactical Missiles. Journal of Guidance, Control, and Dynamics 2011, 34 (1), 107–117.

37
27. Kang, S.; Kim, H. J.; Lee, J.; Jun, B.; Tahk, M. Roll-Pitch-Yaw Integrated Robust Autopilot
Design for a High Angle-of-Attack Missile. Journal of Guidance, Control, and Dynamics
2009, 32 (5), 1622–1628.

28. Arrow, A.; Williams, D. E. Comparison of Classical and Modern Missile Autopilot
Techniques. Journal of Guidance 1989, 12 (2), 220–227.

29. Williams, D. E.; Friedland, B.; Madiwale, A. N. Modern Control Theory for Design of
Autopilots for Bank-to-Turn Missiles. Journal of Guidance 1987, 10 (4), 378–386.

30. Maley, J. Roll Orientation from Commercial-Off-The-Shelf (COTS) Sensors in the Presence
of Inductive Actuators. Institute of Navigation Joint Navigation Conference, Colorado
Springs, CO, 2010.

31. Rogers, J.; Costello, M.; Hepner, D. Roll Orientation Estimator for Smart Projectiles Using
Thermopile Sensors. Journal of Guidance, Control, and Dynamics 2011, 34 (3), 688–697.

32. Rogers, J.; Costello, M.; Harkins, T.; Hamaoui, M. A Low-Cost Orientation Estimator for
Smart Projectiles Using Magnetometers and Thermopiles. Navigation 2012, 59 (1), 9–24.

33. Rogers, J.; Costello, M.; Harkins, T.; Hamaoui, M. Effective Use of Magnetometer
Feedback for Smart Projectile Applications. Navigation 2011, 58 (3), 203–219.

34. Arrow Tech Assoc. PRODAS User Manual; South Burlington, VT, 1997.

38
Nomenclature

= roll, roll rate, roll acceleration


= axial moment-of-inertia
= mass
= diameter
= reference area
= velocity
= Mach number
pitch angle-of-attack, yaw angle-of-attack, total angle-of-attack,
= aerodyanic roll angle
= atmospheric density
= dynamic pressure

= roll damping moment coefficient


= radial center-of-pressure
= axial center-of-pressure
= canard normal force coefficient, 1st, 3rd, and 5th order terms
= static roll moment coefficient
= number
= deflection
= velocity of projectile center-of-gravity
= angular velocity of projectile

vector from projectile center-of-gravity to center-of-pressure of ith lifting


= surface

= transformation matrix from body frame to ith lifting surface frame

= time constant
= state and controls vector
= system dynamics, controls, forcing function matrices
= gain matix
= control error and control effort matrices
= Riccati equation matrix
= time
= roll moment, friction moment

39
subscripts = canard
= fin
= ith lifting surface
= command
= delay
= bias
= random
= friction

40
NO. OF NO. OF
COPIES ORGANIZATION COPIES ORGANIZATION

1 DEFENSE TECHNICAL 2 US ARMY ARDEC


(PDF INFORMATION CTR RDAR MEM C
only) DTIC OCA D DEMELLA
8725 JOHN J KINGMAN RD A LICHTENBERG-SCANLAN
STE 0944 BLDG 94
FORT BELVOIR VA 22060-6218 PICATINNY ARSENAL NJ 07806-5000

1 DIRECTOR 2 US ARMY TACOM ARDEC


US ARMY RESEARCH LAB RDAR MEM M
IMAL HRA C MOEHRINGER
2800 POWDER MILL RD J TRAVAILLE
ADELPHI MD 20783-1197 BLDG 94
PICATINNY ARSENAL NJ 07806-5000
1 DIRECTOR
US ARMY RESEARCH LAB 1 US ARMY TACOM ARDEC
RDRL CIO LL AMSRD AAR AEPS
2800 POWDER MILL RD J ROMANO
ADELPHI MD 20783-1197 BLDG 407
PICATINNY ARSENAL NJ 07806-5000
10 US ARMY ARDEC
RDAR MEF E 4 US ARMY TACOM ARDEC
D CARLUCCI RDAR MEM A
M HOLLIS E VAZQUEZ
C STOUT G MALEJKO
A SANCHEZ W KOENIG
R HOOKE S CHUNG
J MURNANE BLDG 94S
BLDG 94 PICATINNY ARSENAL NJ 07806-5000
PICATINNY ARSENAL NJ 07806-5000
2 RDECOM ARDEC
8 US ARMY TACOM ARDEC RDAR MEM A
RDAR MEF S J GRAU
D PANHORST W TOLEDO
G MINER BLDG 94
N GRAY PICATINNY ARSENAL NJ 07806-5000
R FULLERTON
B DEFRANCO 3 US ARMY TACOM ARDEC
M MARSH RDAR MEF I
P FERLAZZO R GRANITZKI
D PASCUA J CHOI
BLDG 94 L VO
PICATINNY ARSENAL NJ 07806-5000 BLDG 95
PICATINNY ARSENAL NJ 07806
3 US ARMY TACOM ARDEC
RDAR MEM C 1 US ARMY ARDEC
R GORMAN RDAR MEM C
D CIMORELLI M LUCIANO
K SANTANGELO BLDG 65S
BLDG 94 PICATINNY ARSENAL NJ 07806
PICATINNY ARSENAL NJ 07806-5000

41
NO. OF NO. OF
COPIES ORGANIZATION COPIES ORGANIZATION

1 US ARMY TACOM ARDEC 3 US ARMY ARDEC


RDAR MEE W SFAE AMO CAS MS
J LONGCORE P BURKE
BLDG 382 G SCHWARTZ
PICATINNY ARSENAL NJ 07806 J HILT
BLDG 162 S
1 US ARMY ARDEC PICATINNY ARSENAL NJ 07806-5000
RDAR MEF
M HOHIL 1 US ARMY ARDEC
BLDG 407 SMO MAS LC
PICATINNY ARSENAL NJ 07806-5000 D RIGOGLIOSO
BLDG 354
4 RDECOM ARDEC PICATINNY ARSENAL NJ 07806
AMSRD AMR SG SD
J BAUMAN 1 PM MAS
H SAGE SFAE AMO MAS SETI
S DUNBAR J FOULTZ
B NOURSE BLDG 354
BLDG 5400 PICATINNY ARSENAL NJ 07806
REDSTONE ARSENAL AL 35898
1 FIRES DEPUTY MGR
1 US ARMY ARMAMENT RD&E CTR EXP MANEUVER WARFARE
RDAR MEM C OFC OF NAVAL RSRCH
D NGUYEN ONR 30
BLDG 94 875 N RANDOLPH ST
PICATINNY ARSENAL NJ 07806-5000 RM 1155B
ARLINGTON VA 22203
1 RDECOM AMRDEC
RDMR WDG N 1 US ARMY TACOM ARDEC
B GRANTHAM RDAR MEF E
BLDG 5400 T RECCHIA
REDSTONE ARSENAL AL 35898 BLDG 94
PICATINNY ARSENAL NJ 07806-5000
2 PM CAS
SFAE AMO CAS 2 NVL SURFC WARFARE CTR
R KIEBLER DAHLGREN DIVISION
P MANZ N COOK
BLDG 171 L STEELMAN
PICATINNY ARSENAL NJ 07806 G33
6210 TISDALE RD STE 223
2 PM CAS DAHLGREN VA 22448-5114
SFAE AMO CAS EX
J MINUS 2 ATK ADVANCED WEAPONS DIV
M BURKE R DOHRN MN07 MN07 MN14
BLDG 171 J WEYRAUCH
PICATINNY ARSENAL NJ 07806 4700 NATHAN LANE N
PLYMOUTH, MN 554422
1 US ARMY TACOM ARDEC
SFAE AMO MAS 1 SAIC
C GRASSANO D HALL
BLDG 354 1150 FIRST AVE STE 400
PICATINNY ARSENAL NJ 07806 KING OF PRUSSIA PA 19406

42
NO. OF NO. OF
COPIES ORGANIZATION COPIES ORGANIZATION

1 ALLIANT TECHSYSTEMS INC 1 GEORGIA INST OF TECHLGY


ALLEGANY BALLISTICS LAB SCHOOL OF AEROSPACE ENG
S OWENS M COSTELLO
MS WV01 08 BLDG 300 RM 180 ATLANTA GA 30332
210 STATE RTE 956
ROCKET CTR WV 26726-3548 1 TEXAS A&M
SCHOOL OF AEROSPACE ENG
1 GEN DYNAMICS ST MARKS J ROGERS
H RAINES COLLEGE STATION TX 77843
PO BOX 222
SAINT MARKS FL 32355-0222 1 ROSE-HULMAN INST OF TECHLGY
SCHOOL OF MECHANICAL ENG
3 GOODRICH SENS AND INERTIAL SYS B BURCHETT
T KELLY TERRE HAUTE IN 47803
P FRANZ
S ROUEN ABERDEEN PROVING GROUND
100 PANTON RD
VERGENNES VT 05491 5 COMMANDER
US ARMY TACOM ARDEC
1 US ARMY ARDEC AMSRD AR AEF D
RDAR MEM C J MATTS
P MAGNOTTI A SOWA
BLDG 94 J FONNER
PICATINNY ARSENAL NJ 07806 M ANDRIOLO
B NARIZZANO
5 BAE ARM SYS DIV B 305
T MELODY APG MD 21005
J DYVIK
P JANKE 45 DIR USARL
B GOODELL RDRL WM
O QUORTRUP P PLOSTINS
4800 E RIVER RD RDRL WML
MINNEAPOLIS MN 55421-1498 P PEREGINO
M ZOLTOSKI
1 US ARMY YUMA PROVING GROUND RDRL WML A
TEDT YPY MW W OBERLE
M BARRON R PEARSON
301 C STREET L STROHM
YUMA AZ 85365-9498 RDRL WML D
M NUSCA
1 TRAX INTRNTL CORP J SCHMIDT
R GIVEN RDRL WML E
US ARMY YUMA PROVING GROUND V BHAGWANDIN
BLDG 2333 I CELMINS
YUMA AZ 85365 G COOPER
J DESPIRITO
1 ARROW TECH ASSOC L FAIRFAX
W HATHAWAY F FRESCONI (5 CPS)
1233 SHELBURNE RD J GARNER
STE D-8 B GUIDOS
SOUTH BULINGTON VT 05403 K HEAVEY
G OBERLIN

43
NO. OF
COPIES ORGANIZATION

J SAHU
S SILTON
P WEINACHT
RDRL WML F
B ALLIK
T BROWN
B DAVIS
T HARKINS
M ILG
G KATULKA
D LYON
J MALEY
C MILLER
P MULLER
D PETRICK
B TOPPER
RDRL WML G
J BENDER
W DRYSDALE
M MINNICINO
RDRL WML H
M FERMEN-COKER
J NEWILL
R SUMMERS
RDRL WMP F
R BITTING
N GNIAZDOWSKI

44

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy