throbber
JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS
`Vol. 38, No. 2, February 2015
`
`Flight Control of a Small-Diameter Spin-Stabilized Projectile
`Using Imager Feedback
`
`Frank Fresconi∗
`U.S. Army Research Laboratory, Aberdeen Proving Ground, Maryland 21005
`and
`Jonathan Rogers†
`Georgia Institute of Technology, Atlanta, Georgia 30332
`
`DOI: 10.2514/1.G000815
`
`Small-diameter gun-launched projectiles pose a challenging platform on which to implement closed-loop guidance
`and control. This paper presents a novel imager-based guidance and control algorithm for small-diameter spin-
`stabilized projectiles. The control law is specifically formulated to rely on feedback only from a strapdown detector
`and roll angle sensors. Following introduction of the projectile nonlinear dynamic model, an integrated guidance and
`control algorithm is presented in which control commands are computed directly from detector feedback using a gain-
`scheduled proportional control. Time-varying controller gains are derived through a surrogate modeling approach,
`and controller performance is further enhanced through use of an observer that filters unwanted angular motion
`components from detector feedback. Example closed-loop flight simulations demonstrate performance of the
`proposed control system, and Monte Carlo analysis shows a factor of 2 accuracy improvement for the closed-loop
`system over ballistic flight. Results indicate that delivery system improvements are achievable in small, gyroscopically
`stabilized projectiles containing low-cost guidance elements using the proposed integrated guidance and control
`approach.
`
`Cm0 ,
`
`Clp
`Cm,
`Cmα , Cmα3
`Cmq
`CN,
`CNα , CNα3
`Cnpα
`CX, CX0 , CX α2
`
`CN0 ,
`
`CMX , CMZ ,
`
`
`
`CMl , CM
`m
`CYpα
`D
`fL
`I
`Ke, K _e
`
`Kε, K _ε
`
`L, M, N
`
`M
`m
`q
`
`qM
`
`rc:g:→c:p:i
`
`=
`=
`
`=
`=
`
`Nomenclature
`roll damping moment coefficient
`total,
`trim,
`linear, and nonlinear pitching
`moment coefficient
`pitch damping moment coefficient
`total, trim, linear, and nonlinear normal force
`coefficient
`= Magnus moment coefficient
`=
`total, zero-yaw, and yaw-dependent axial force
`coefficient
`= wing actuator aerodynamic coefficients
`
`=
`
`=
`
`= Magnus force coefficient
`=
`reference diameter, m
`=
`lens focal length, m
`= moment of inertia matrix, kg · m2
`=
`controller proportional and derivative gain
`values
`surrogate model proportional and derivative
`gain values
`aerodynamic moments acting on projectile,
`Nm
`= Mach Number
`= mass, kg
`=
`dynamic pressure at projectile mass center,
`N∕m2
`dynamic pressure at wing actuator center of
`pressure, N∕m2
`vector from center of gravity to ith movable
`aerodynamic surface center of pressure, m
`
`=
`
`=
`
`Received 7 June 2014; revision received 3 September 2014; accepted for
`publication 4 September 2014; published online 2 January 2015. This material
`is declared a work of the U.S. Government and is not subject to copyright
`protection in the United States. Copies of this paper may be made for personal
`or internal use, on condition that the copier pay the $10.00 per-copy fee to the
`Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923;
`include the code 1533-3884/15 and $10.00 in correspondence with the CCC.
`*Guided Flight Dynamics Lead, Weapons and Materials Research
`Directorate.
`†Assistant Professor, School of Mechanical Engineering.
`
`181
`
`
`
`rIOT, rI
`
`OP
`
`PT, rV
`rB
`
`PT
`
`~ry, ~rz
`
`S
`TBE
`
`TBM
`
`TBMi
`
`TVE
`
`_u, _v, _w
`
`V, Vc:g:∕I
`
`X, Y, Z
`XG, YG, ZG
`x, y, z
`_x, _y, _z
`xB, yB, zB
`xc:p:, rc:p:
`xE, yE, zE

`α



`
`
`
`ϵy, ϵ
`
`z
`
`~ϵ
`y, ~ϵ
`
`z
`
`
`
`~ϵ y , ~ϵ
`
`z
`
`=
`
`=
`
`=
`
`=
`=
`
`=
`
`=
`
`=
`
`=
`
`=
`
`=
`=
`=
`=
`=
`=
`=
`=
`=
`=
`=
`=
`
`=
`
`=
`
`=
`
`vector from origin to target and from origin
`to
`projectile
`expressed
`in Earth-fixed
`frame, m
`vector from projectile to target, expressed in
`body frame and velocity vector frame, m
`guidance command reference signals in fixed-
`plane coordinates, m
`reference area, m2
`transformation matrix from body- to Earth-
`fixed coordinates
`transformation matrix from body-fixed to
`wing surface coordinates
`transformation matrix from body-fixed to
`movable aerodynamic surface coordinates
`transformation matrix from velocity vector
`coordinates to Earth-fixed coordinates
`time rate of change of body-fixed coordinate
`system translational velocity, m∕s2
`total velocity of projectile with respect to
`wind, velocity of center of gravity with respect
`to inertial frame in body-fixed coordinates,
`m∕s
`aerodynamic forces acting on projectile, N
`gravity forces acting on projectile, N
`inertial position, m
`inertial translational velocity, m∕s
`body-fixed coordinate system
`axial and radial center of pressure, m
`Earth-fixed coordinate system
`pitch angle of attack, rad
`total angle of attack, rad
`yaw angle of attack, rad
`normalized control magnitude
`state variable for surrogate model representing
`y or z component of rV
`PT
`body-fixed coordinates of lateral components
`of vector from projectile to target in detector
`plane, m
`fixed-plane coordinates of lateral components
`of vector from projectile to target in detector
`plane, m
`filtered ~ϵy, ~ϵz signals from Kalman filter
`
`Downloaded by Michael Daniels on August 3, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.G000815
`
`Yuneec Exhibit 1016 Page 1
`
`

`
`182
`
`FRESCONI AND ROGERS
`
`
`
`ηy, η
`
`z
`
`ϕA
`ξ, ω
`
`n, fμ, μ
`
`C, ϕ
`
`B,
`

`ϕ, θ, ψ

`CMD, ϕ
`Φ, ^ϕ

`B∕I
`
`=
`
`=
`=
`
`=
`=
`=
`
`=
`
`as
`
`zero-mean
`
`noise, modeled
`detector
`Gaussian white noise, m
`aerodynamic roll angle, rad
`damping, natural frequency, control scaling,
`and control input of surrogate model
`atmospheric density, kg∕m3
`roll, pitch, and yaw Euler angles, rad
`Earth-fixed commanded, Earth-fixed control,
`body-fixed control, relative, and stowed roll
`angles, rad
`body-fixed
`coordinate
`velocity, rad∕s
`
`system rotational
`
`I.
`
`Introduction
`
`D EVELOPMENT of viable guidance and control technologies
`
`for gun-launched projectiles is a unique engineering challenge.
`Over the past several decades, numerous solutions have been
`proposed, primarily for large artillery projectiles [1,2] or for slowly
`rolling airframes [3–5]. The general methodology employed by most
`of these studies has been to use GPS [1,6], inertial measurements [7],
`and seekers [8] to provide feedback to a flight control system. Course
`correction maneuvers are then initiated through aerodynamic mech-
`anisms [2,4,9] or jet control [10]. The numerous engineering
`challenges associated with fielding viable smart weapons systems
`have been repeatedly noted in the literature, including lack of payload
`volume for electronics, harsh accelerations at gun launch, highly
`nonlinear dynamics, and low-cost requirements.
`At the same time, guidance and control technologies developed to
`date largely do not scale to smaller man-portable munitions. Assault
`and carbine rifles, grenade launchers, and machine guns have been
`the backbone of the infantry for many years. Currently there is a
`desire to increase lethality of the infantry squad through precision
`guidance technologies designed for small-diameter projectiles.
`Successful application of miniaturized low-cost sensors and actu-
`ators to projectiles in this domain has the potential to revolutionize
`squad-level
`lethal capability. However, small-caliber projectiles
`impose severe constraints on control system development. Volume is
`a critical concern because the diameter of these projectiles ranges
`from only 5.56 to 40 mm. Additionally, these projectiles are launched
`from rifled guns that produce spin rates in the hundreds to thousands
`of rotations per second. The nonlinear dynamics caused by these high
`spin rates complicates sensor stimuli and stresses the actuator
`bandwidth. In addition, high-bandwidth and efficient control inputs
`are necessary because time of flight is often only a few seconds.
`Finally, because small-diameter projectiles are meant for squad-level
`weapons, it is essential that the integrated projectiles can be fielded at
`low cost. Low-cost requirements are manifested in terms of loose
`manufacturing tolerances, limitations in the type of feedback that can
`be obtained, and limitations in the degree of sensor calibration that
`can be performed. As a result, applicable guidance laws must be
`robust to large measurement uncertainty and require minimal state
`feedback.
`This paper outlines a novel guidance and control algorithm for
`small-diameter, gyroscopically stabilized projectiles equipped with a
`strapdown imager. The integrated guidance and control loop is
`designed to transform detector feedback directly into control surface
`commands. The proposed guidance law is formulated through
`control response modeling and development of a surrogate model for
`the input–output response. Once the surrogate model is developed, a
`linear control system is derived and stability bounds are determined.
`Because strapdown sensor feedback is severely perturbed by body
`angular motion during controlled flight, an extended Kalman filter is
`added to the feedback loop to remove body angular rate components
`from the feedback signals. This is feasible because the characteristic
`fast and slow-mode frequencies of the projectile are approximately
`known. It is assumed that the imager used in this study may be
`modeled as a pinhole camera [11]. Thus, details concerning whether
`the strapdown (or “stiff-neck”) detector signals are derived from a
`laser-designated spot or
`through image-based navigation are
`irrelevant for this study, because both seekers and standard cameras
`
`approximately satisfy the pinhole camera modeling assumptions
`[12]. The control system presented here is applied to an example
`40 mm grenade projectile equipped with a rotating wing maneuver
`mechanism. This grenade projectile exhibits flight characteristics
`typical of small-diameter spin-stabilized rounds, including high-
`amplitude angular motion, which highlight controller performance in
`a challenging flight environment.
`The paper proceeds as follows. First, the nonlinear projectile
`dynamic model is defined, and some representative flight behaviors
`are highlighted that complicate the guidance problem and preclude
`traditional solutions. An integrated guidance and control approach is
`presented, including surrogate modeling of the open-loop control
`response, followed by design of a linear controller that computes
`control surface deflections directly from detector outputs. Stability
`bounds on surrogate model control gains are analyzed. Example
`ballistic and controlled flight simulation results are presented, along
`with Monte Carlo simulations, which analyze performance with and
`without the angular rate estimator. Overall, resulting closed-loop
`performance of the proposed guidance system is shown to be favor-
`able, leading to significant reductions in impact point dispersion.
`
`II. Projectile Flight Dynamic Model
`This section describes the nonlinear flight dynamic model used in
`this study, including simulation of dynamics and aerodynamics,
`control actuation, and detector feedback. An example projectile con-
`cept is described, followed by a description of the mathematical
`model for each relevant component. A sample ballistic trajectory is
`provided to highlight key flight behaviors of the system.
`
`A. Projectile Concept
`The guidance and control formulation proposed in this study
`applies to the gamut of small-diameter spin-stabilized projectiles. For
`these small airframes, high control authority is difficult to obtain due
`to gun-launched packaging constraints and dynamic flight insta-
`bilities that often arise [13,14]. In this paper, a 40-mm-diam actively
`controlled grenade projectile is selected as an illustrative example.
`This system features a subsonic launch and a projectile spin rate of
`approximately 60 Hz. A rendering of the example projectile, in-
`cluding maneuver and navigation technologies, is shown in Fig. 1.
`An imaging device is packaged in a strapdown manner in the nose to
`reduce cost and ensure survivability during gun launch. Fuzing and
`warhead are contained in the central portion of the projectile body.
`The maneuver mechanism is placed at the aft end of the spin-
`stabilized projectile to maximize control authority [15,16]. This
`mechanism consists of a rotary motor (shown in red in Fig. 1) linked
`to a deflected wing (light blue in Fig. 1). As the projectile body spins
`over one revolution, the motor rotates the wing in the opposite
`direction at the same rate. This deploys the wing into the airstream for
`a portion of the roll cycle to create an aerodynamic asymmetry.
`
`Fig. 1 Example spin-stabilized 40 mm projectile with navigation and
`maneuver components.
`
`Downloaded by Michael Daniels on August 3, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.G000815
`
`Yuneec Exhibit 1016 Page 2
`
`

`
`FRESCONI AND ROGERS
`
`MR ˆ qS
`
`
`LR ˆ qSD
`Clp…M† pD
`
`Cmα…M†α ‡ Cmα3…M†α3 ‡ Cmα…M† qD
`
`‡ Cnpα…M; α†β pD
`
`NR ˆ qS
`−Cmα…M†β − Cmα3…M†β3 ‡ Cmq…M† rD
`
`‡ Cnpα…M; α†α pD
`
`
`
`2V
`
`2V
`
`2V
`
`183
`
`(4)
`
`2V
`
`2V
`
`Various means may be used to modulate the control effort because the
`wing has a fixed deflection angle. Control commands can “fishtail”
`the flight, a clutch can decouple the motor and wing, the wing can
`rotate at a frequency off the projectile spin rate, or multiple motor–
`wing assemblies can mix the relative phasing of the roll angles to
`deploy the wings into the airstream. For the purposes of this study,
`two motor–wing assemblies are considered to be packaged near the
`projectile base oriented 180 deg apart in roll angle. Because the focus
`of this study is on the imager-based guidance and control algorithm,
`additional details regarding this particular maneuver mechanism are
`omitted here for brevity but may be found in [17–20].
`
`B. Flight Dynamic Model
`The projectile body-fixed coordinate system is illustrated in Fig. 2,
`including the definition of aerodynamic angles of attack used to
`describe forces and moments. A standard body frame B is defined, as
`is a velocity vector frame V, where the V frame is derived by first
`rotating by β about the body z axis, and then by α about the resulting y
`axis. For this airframe, aerodynamic forces and moments are
`composed of components from the projectile rigid body (denoted by
`superscript R) and movable wing surfaces (denoted by super-
`script M).
`Total forces and moments on the projectile are given by
`
`X ˆ XR ‡ XM
`Y ˆ YR ‡ YM
`Z ˆ ZR ‡ ZM
`
`L ˆ LR ‡ LM
`M ˆ MR ‡ MM
`N ˆ NR ‡ NM
`
`(1)
`
`(2)
`
`Projectile body forces include contributions from axial force, normal
`force, and side (Magnus) force and can be described by the following
`aerodynamic expansion:
`
`
`…M† ‡ CXα2…M† α2Š
`XR ˆ −qS‰CX0
`YR ˆ −qS
`CNα…M†β ‡ CNα3…M†β3 − CYpα…M; α†α pD
`
`CNα…M†α ‡ CNα3…M†α3 − CYpα…M; α†β pD
`
`
`
`
`2V
`
`2V
`
`ZR ˆ −qS
`
`To model the wing forces and moments, first define the relative roll
`angle of wing actuator i with respect to the inertial frame as
`
`Φi ˆ ϕCi
`
`− ϕ
`
`(5)
`
`As the projectile rolls, the wing area exposed to the airstream varies
`such that it is maximum when the projectile roll angle coincides with
`the desired maneuver plane. Thus, define the amount of wing area
`exposed to the airstream through the stowed roll angle as
`
`^ϕ
`
`i ˆ Φ
`
`i
`
`− ϕ
`
`Bi
`
`(6)
`
`The total angle of attack of wing actuator i αMi and the yaw angle of
`attack βMi are given by the following expressions:
`
`q
`q
`
`2 ‡ wMi
`vMi
`2 ‡ vMi
`2 ‡ wMi
`uMi
`
`(7)
`
`1A
`
`2
`
`2
`
`0@
`
`Mi ˆ a sin
`
`α
`
`(8)
`
`1A
`
`q
`
`vMi
`2 ‡ vMi
`2 ‡ wMi
`uMi
`
`2
`
`0@
`
`βMi ˆ a sin
`
`where the local velocity at wing actuator i in local wing coordinates is
`obtained according to
`
`(3)
`
`‰ uMi
`
`vMi wMi ŠT ˆ TBMi…Vc:g:∕I ‡ ωB∕I × rc:g:→c:p:i†
`
`(9)
`
`Likewise, projectile body moments include contributions from roll
`damping, static pitching, pitch damping, and Magnus moments and
`can be described by the following expansion:
`
`Given the stowed roll angle and angle-of-attack relationships, the
`aerodynamic coefficients for each wing are calculated using aero-
`dynamic expansions dependent on the local Mach number and the
`stowed roll angle (which describes the wing area exposed to the
`freestream flow). These coefficients are given by
`
`…M; ^ϕ
`
`i†β
`
`Mi ‡ CM
`
`Xα2
`
`M
`
`M
`
`
`
`…M; ^ϕi†β2
`
`Mi
`
`Xα
`
`Mi
`
`Fig. 2 Body-fixed coordinate system.
`
`Finally, Eq. (11) provides the aerodynamic forces and moments of
`wing actuator i in projectile body coordinates:
`
`lα
`
`M
`
`mα
`
`Mi
`
`Mi ‡ CM
`
`mα3
`
`
`
`…M; ^ϕi†β3
`
`Mi
`
`M
`
`(10)
`
`M
`
`i† ‡ CM
`…M; ^ϕ
`X ˆ CM
`CMi
`‡ CM
`…M; ^ϕi†β3
`…M; ^ϕi† ‡ CM
`Z ˆ CM
`CMi
`l ˆ CM
`…M; ^ϕ
`i† ‡ CM
`CMi
`m ˆ CM
`…M; ^ϕ
`i† ‡ CM
`CMi
`
`Zα
`
`M
`
`…M; ^ϕi†β3
`
`Mi
`
`M
`
`Zα3
`
`…M; ^ϕi†βMi ‡ CM
`…M; ^ϕ
`i† α
`…M; ^ϕ
`i†β
`
`Xα3
`
`M
`
`X0
`
`Z0
`
`l0
`
`m0
`
`Downloaded by Michael Daniels on August 3, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.G000815
`
`Yuneec Exhibit 1016 Page 3
`
`

`
`184
`
`FRESCONI AND ROGERS
`
`Fig. 3 Earth- and body-fixed coordinate systems and euler angles.
`
`(11)
`
`3775
`
`3775
`
`CMi
`X
`CMi
`
`Z 0
`2664
`
`control system and ϕ
`C be the angle actually achieved by the wing
`actuator. To isolate controlled flight performance, we assume for the
`remainder of this study that the actuator responds at least 10 times
`faster than the projectile roll rate (∼60 Hz) so that actuator dynamics
`CMD ˆ ϕC. Experimental studies with
`may be neglected, and thus ϕ
`this actuator described in [20] have shown that an actuator bandwidth
`of approximately 600 Hz, or 10 times the projectile roll rate, is likely
`achievable using a rotary actuator design.
`
`C. Detector Modeling
`is
`To minimize cost, a simple strapdown feedback concept
`proposed. A strapdown detector provides the location of the target in
`the image plane and is modeled as a pinhole camera. The pinhole
`camera is used to project a three-dimensional scene onto a two-
`dimensional image plane, neglecting distortion effects from the lens
`or finite aperture size [11]. As mentioned previously, a pinhole
`camera model may be used to represent both seekers and standard
`camera devices [12]. Thus, details concerning whether detector data
`arrive from a laser-designated spot detector or through image-based
`navigation are irrelevant for this study and it is assumed that the target
`is identifiable in the detector plane. To formulate the pinhole model
`for the detector, start by defining the relative position between the
`projectile and target according to
`
`where Eq. (17) is written in body coordinates. Note that the presence
`of the roll angle in the transformation matrix in Eq. (17) yields a
`rapidly changing target location as seen by the strapdown detector for
`a spin-stabilized projectile.
`In practice, a passive detector can sense only the lateral relative
`positions, and the actual detector displacements are scaled based on
`the lens properties according to
`
`
`
`
`

`y
`ϵz
`

`
`fL
`
`
`
`PT…2†
`rB
`PT…3†
`rB
`
`
`
`
`
`
`

`

`y
`ηz
`
`(18)
`
`2664
`
`ˆ −qMi ST
`
`−1
`BMi
`
`3775
`
`3775
`
`2664
`
`XMi
`
`YMi
`
`ZMi
`
`LMi
`
`MMi
`
`NMi
`
`2664
`
`CMi
`
`l 0
`
`ˆ −qMi SDT
`
`−1
`BMi
`
`CMi
`m
`
`
`rB
`
`PT ˆ TTBE…rI
`
`OT
`
`OP†
`− rI
`
`(17)
`
`Given this force and moment model, the equations of motion for the
`projectile are formulated using a Newton–Euler approach. The
`inertial, flat-Earth coordinate system (denoted by frame E) and the
`body-fixed coordinate system B are related by Euler roll ϕ, pitch θ,
`and yaw ψ angles, as shown in Fig. 3. The translational and rotational
`kinematic equations are given by
`
`(12)
`
`35
`
`u v w
`
`24
`35
`
`cθcψ
`cθsψ
`−sθ
`
`sϕsθcψ − cϕsψ
`sϕsθsψ − cϕcψ
`sϕcθ
`
`cϕsθcψ ‡ sϕsψ
`cϕsθcψ ‡ sϕsψ
`cϕcθ
`
`24
`

`
`35
`
`_x
`_y
`_z
`
`24
`
`fL
`PT…1†
`− rB
`
`where the left-hand side of Eq. (18) represents the raw displacements
`of the target in the image plane and ηy, ηz are zero-mean Gaussian
`white noise representing detector quantization error, distortion
`effects, and other sources of detector noise [12]. A diagram of the
`projective transformation for ϵ
`z is shown in Fig. 4, and a similar
`diagram can be constructed for ϵ
`y. Note, however, that the raw
`detector measurements in Eq. (18) are not sufficient to compute a
`desired maneuver plane. Roll estimation is necessary to resolve the
`target error components in the fixed-plane (i.e., nonrolling) co-
`ordinate system. Numerous authors have studied roll angle esti-
`mation for smart weapons based on magnetometer [21–23] or
`infrared sensor measurements [24,25]. For this study, it is assumed
`that some source of roll feedback is available, and thus the target error
`components can be transformed into the fixed-plane frame according
`to
`
`Lf
`
`z
`

`
`Bx
`c.g. IV
`/
`
`PTr
`
`Target
`
`Bz
`
`Ex
`
`Ez
`Fig. 4 Definition of detector displacement ϵz. Image plane is shown in
`gray. (This diagram not to scale.)
`
`(13)
`
`35
`
`p q r
`
`24
`35
`
`sϕtθ
`1
`cϕ
`0
`0 sϕ∕cθ
`
`cϕtθ
`−sϕ
`cϕ∕cθ
`
`24
`

`
`35
`
`_ϕ
`_θ
`_ψ
`
`24
`
`where cα ≡ cos α, sα ≡ sin α, and tα ≡ tan α. Application of
`Newton–Euler kinetics to a projectile in free flight yields the
`following translational and rotational dynamic equations:
`
`24
`
`−
`
`35
`
`X ‡ XG
`Y ‡ YG
`Z ‡ ZG
`
`24
`
`ˆ 1
`m
`
`35
`
`35
`
`L M N
`
`24
`
`ˆ I
`
`−1
`
`35
`
`24
`
`_u
`_v
`_w
`
`_p
`_q
`_r
`
`24
`
`(14)
`
`(15)
`
`35
`
`u v w
`
`p q r
`
`35
`
`24
`35
`
`24
`
`35
`
`I
`
`−r
`q
`0
`0 −p
`r
`−q p
`0
`
`24
`
`− I
`
`−1
`
`−r
`q
`0
`0 −p
`r
`−q p
`0
`
`Note that aerodynamic forces and moments are computed as
`shown earlier, and the gravity force is included according to
`ŠT
`
`‰ XG YG ZG ŠT ˆ TT
`BE‰ 0 0g
`
`(16)
`
`The equations of motion given by Eqs. (12–15) are integrated
`forward in time using a fixed time step Runge–Kutta integration
`routine to obtain a single flight trajectory. Note that control action is
`input to this system through specification of the roll angle ϕC, which
`describes the maneuver plane in which the wing actuator is fully
`extended. Let ϕ
`CMD be the commanded roll angle output from the
`
`Downloaded by Michael Daniels on August 3, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.G000815
`
`Yuneec Exhibit 1016 Page 4
`
`

`
`FRESCONI AND ROGERS
`
`185
`
`Fig. 5 is the bias in the yaw angle of attack of about −1 deg, a result of
`the so-called yaw of repose. Yaw of repose is due to an interaction
`between the spin rate and overturning motion [27] and yields a cross-
`range drift of about 1 m over the 200 m range trajectory.
`For projectiles equipped with a maneuver mechanism, previous
`work has proven that flight instabilities occur for spin-stabilized
`projectiles maneuvering perpendicular to the gravity field when the
`control effectiveness is sufficiently high [13,14]. The potential for
`such instability exists in the current application. To demonstrate this,
`the aerodynamic force and moment coefficients for the wing were
`multiplied by a factor of 3 and a flight simulation was conducted with
`a constant maneuver command of ϕ
`to the right). Figure 6 demonstrates the resulting large-amplitude
`angular motion (counterclockwise when viewed from behind the
`projectile) which is a typical feature of this flight instability for spin-
`stabilized rounds. The implications of this instability are that,
`although it is difficult to maneuver small-diameter spin-stabilized
`projectiles, there is an upper bound on possible control authority,
`which is not particularly high, to maintain stability during maneuver.
`This leads to the conclusion that efficient control design and
`disturbance rejection are at least as important, if not more important,
`than generation of large control authority for accuracy improvement
`in these types of rounds.
`
`CMD ˆ 90 deg (full deflection
`
`III. Guidance and Flight Control Design
`Classical missile autopilots use an inner–outer loop guidance and
`control approach where proportional navigation (PN) guidance is
`based on line-of-sight and closing velocity estimates. In many
`implementations of these algorithms, accelerometers are used in an
`inner-loop control to track desired center of mass acceleration
`commands from PN guidance, and gyroscopes may be used to damp
`unwanted angular motion. Although PN-based inner–outer loop
`control may be suitable for some types of guided munition concepts,
`it is difficult to apply here for a variety of reasons. In the strapdown
`detector plane, the target rotates at the spin rate, and motion due to the
`trajectory arc and coning are superimposed. This action, along with a
`time of flight on the order of seconds, precludes high-accuracy
`estimation of line-of-sight rate from a low-cost detector. In addition,
`low-cost measurement devices such as accelerometers are severely
`stressed in the spin-stabilized environment. Manufacturing con-
`straints typically mean that these devices are subject to some error in
`misalignment and misposition, which can yield feedback that is
`largely unusable. To demonstrate this, Fig. 7 shows a component of
`lateral acceleration for a typical ballistic flight of the 40 mm grenade,
`both the truth value and as measured by a notional accelerometer. A
`random misalignment and misposition error were selected from a
`Gaussian distribution with standard deviations of 0.5 deg and 5 mm,
`respectively (small errors on the order of manufacturing tolerance).
`As shown in Fig. 7, high acceleration from the fast-spinning axis
`bleeds over into the lateral channels and largely overwhelms the true
`
`Fig. 5 Angular motion for ballistic flight of the 40 mm grenade concept.
`
`
`
`
`
`
`
`
`
`
`

`
`~ϵy
`~ϵ
`
`z
`
`cos ϕ − sin ϕ
`sin ϕ
`cos ϕ
`
`ϵy

`z
`
`(19)
`
`The target error components in the fixed-plane frame in Eq. (19)
`form the feedback signal used by the controller. Thus, a strapdown
`detector and roll angle sensor(s) are the only measurement devices
`assumed in this study. A detector update rate of 100 Hz, focal length
`
`of fL ˆ 0.1 m, and standard deviation of 0.001 m for ηy and ηz are
`
`used in Sec. IV of this paper for simulation studies.
`
`D. Example Flight Trajectories
`Several example simulations are provided here before control
`design to demonstrate key aspects of the flight behavior of small-
`diameter actively controlled projectiles. The example 40 mm system
`under investigation is launched with conditions necessary to fly to a
`ballistic range of 200 m, and no control inputs are given. Mass
`properties and aerodynamics for the example projectile obtained
`from solid modeling, semi-empirical aeroprediction, wind-tunnel
`tests, and computational fluid dynamics are provided in [18]. The
`angular motion over the entire ballistic flight (with no initial angular
`body rates) is shown in Fig. 5. A counterclockwise coning motion
`with amplitude of approximately 1 deg is visible due to Magnus
`moments on the projectile [26]. Magnus moments are an aero-
`dynamic effect resulting from the high projectile spin rate (which is
`necessary for gyroscopic stabilization) and often result in a limit cycle
`oscillation manifest as a coning motion. Aeromechanics theory
`dictates that this coning occurs at a natural frequency of about 2 Hz
`for this projectile [27]. Note that the Magnus moment is highly
`nonlinear with angle of attack, and thus maneuvering flight may
`significantly alter the response. Another important effect evident in
`
`Fig. 6 Instability of 40 mm grenade projectile with high control
`effectiveness.
`
`Fig. 7 Example true and measured lateral accelerometer signal for
`ballistic flight.
`
`Downloaded by Michael Daniels on August 3, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.G000815
`
`Yuneec Exhibit 1016 Page 5
`
`

`
`186
`
`FRESCONI AND ROGERS
`
`lateral acceleration. There is also significant high-frequency content
`in the signal at the spin rate (approximately 60 Hz) as shown in the
`oscillations for the zoomed-in portion of the measured acceleration.
`Clearly, use of this feedback in an inner-loop controller would not
`yield suitable performance. Similar arguments apply to gyroscope
`measurements in this environment.
`To address these issues with sensor measurements, this study
`proposes to use a detector and roll angle sensors as the sole feedback
`measurements in an integrated inner–outer loop control. This inte-
`grated control design is similar to that proposed for missile end-game
`guidance as described in [28–30]. The idea behind integrated
`guidance is that the controller is designed to compute control inputs
`directly from feedback measurements without any intermediate
`loops. This design allows for minimal feedback but requires some
`knowledge of the input–output dynamics of the system.
`Control design proceeds as follows. First, a second-order surrogate
`model is developed to describe the relationship between control
`inputs and lateral vehicle response. Once this model is identified, a
`proportional control gain is derived that provides suitable closed-
`loop response characteristics for the surrogate model. A model
`equivalency procedure is then performed between the surrogate
`model and the full nonlinear model, and a proportional controller is
`designed for the closed-loop nonlinear system. The gain for this
`controller is designed based on the surrogate model feedback gain
`and detector dynamics. This results in a time-varying control gain
`with a linear dependence on time of flight. Finally, an observer is
`introduced that filters body angular motion from detector feedback.
`
`A. Control System Overview
`The basic structure for the reduced-state control system proposed
`here is shown in Fig. 8. The nonlinear system dynamics formulated in
`the preceding flight models is contained in the function H. The
`z ŠT. This detector
`measurements are the raw detector data y ˆ ‰ ~ϵ
`~ϵ
`y
`data are filtered using the observer labeled L, which attempts to filter
`out components due to body angular rates and produces feedback
`signals y ˆ ‰ ~ϵ
`z ŠT. The error signal is formed by subtracting a
`~ϵ
`y
`reference from the feedback. The reference in the downrange
`direction is nonzero to account for trajectory arc. Likewise, the
`reference in the cross-range direction is nonzero to account for the
`difference between the center of the coning motion and the velocity
`vector caused by yaw of repose (as shown in Fig. 5). The control
`vector consists of the magnitude and direction of maneuver given by
`
`Fig. 8 Control system block diagram.
`
`Fig. 9 Representative downrange reference signal.
`
`
`
`q
`CMD ŠT
`u ˆ ‰ δ ϕ


`− ~ry†2 ‡ …~ϵ
`− ~rz†2
`~ϵ
`
`Ke
`
`y
`
`
`
`T
`
`(20)
`
`−1 −…~ϵ z − ~rz†
`
`−…~ϵ
`y − ~ry†
`tan
`
`z
`
`where δ represents the desired normalized control magnitude and
`varies between zero and one. The proportional gain value Ke is a
`critical parameter in control system performance because it provides
`a direct relationship between error magnitude in the detector plane
`and control magnitude. It was noted during numerous simulation
`trials with this projectile that a constant gain value yielded either
`closed-loop instability or, if the gain was lowered, insufficient control
`response. It was determined that the nonlinear dynamics of the
`problem, and the limitations on practical feedback, prohibits the
`classic control design method of finding a constant linear relationship
`between control input and feedback output. Thus, a surrogate model
`approach was used to determine a time-varying gain that yields
`reasonable performance.
`Note that, if a zero reference value ~rz were used in Eq. (20), the
`projectile would unnecessarily spend valuable control effort trying to
`maneuver toward the ground due to the trajectory arc. To overcome
`this, a downrange reference signal representing the curvature of the
`trajectory can be estimated preflight using the approximate launch
`angle derived from a weapon fire control system. Because the muzzle
`velocity, mass properties, and aerodynamics are relatively fixed
`quantities, a downrange reference can be formed for a variety of
`launch angles and stored in a lookup table. These data may be indexed
`as a function of time of flight to effectively zero the target location on
`the detector in the downrange direction. An example of a reference
`signal for a launch angle of about 15 deg and a focal length of
`
`fL ˆ 0.1 m is presented in Fig. 9.
`
`Likewise, the cross-range reference signal is nonzero due to yaw of
`repose, which causes the projectile to cone about a vector that is at
`some angular offset to the velocity vector. Let this angular offset be
`given by β, and assume it is approximately constant throughout the
`direct fire trajectory. For instance, in Fig. 5, β is observed to be
`approximately 1 deg for the example projectile. Using the pinhole
`camera model in Eq. (18), the cross-range reference signal is given
`
`by ~ry ˆ fL tan β. For the example projectile, if ⠈ 1 deg and
`fL ˆ 0.1 m, ~ry is approximately 0.0017 m.
`
`B. Surrogate Model Development
`To develop a surrogate model for control response, nonlinear
`simulations were conducted wherein open-loop maneuvers were
`commanded at different magnitudes and directions (i.e., to cause the
`projectile to travel left, right, up, or down). The resulting response
`data due to these maneuvers were collected. More specifically, the
`relative position (21) and velocity (22) of the projectile with respect to
`the target in velocity vector coordinates was tabulated according to
`
`
`rV
`
`PT ˆ TTVE…rI
`
`OT
`
`OP†
`− rI
`
`PT ˆ _rI
`_rV
`
`PT
`
`− rI
`
`OP

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket