Sei sulla pagina 1di 13

AIAA 2009-6318

AIAA Guidance, Navigation, and Control Conference


10 - 13 August 2009, Chicago, Illinois

Linear Dynamics and PID Flight Control


of a Powered Paraglider
Yoshimasa Ochi 1 , Hiroyuki Kondo 2 , and Masahito Watanabe 3
National Defense Academy, Yokosuka, Kanagawa 239-8686, Japan
TPF

FPT

TPF

FPT

TPF

FPT

This paper describes analytical derivation of a linear dynamic model of a powered


paraglider (PPG) from a nonlinear dynamic model, which was derived by the authors. The
linear model indicates that the longitudinal dynamics are decoupled from the lateraldirectional one. The linear model obtained on the canopy coordinates is transformed into the
one on the payload coordinates by linear state transformation, since measurement units are
mounted on the payload. The transformation matrix is obtained by linearizing the relations
of the velocity and the angular velocity between the canopy and the payload, and also by
linearizing the relation among the coordinate transformation matrices. The authors are
developing a new design method of a proportional-integral-derivative (PID) control system,
which is based on plant model reduction with -gap metric and the integral-type optimal
servomechanism. This method is applied to flight controller design for three single-inputsingle-output systems of the PPG. A design example and computer simulation results show
linear dynamic properties of the PPG and illustrate good control performance and desirable
stability margins of the PID controllers.

I. Introduction

powered paraglider (PPG), shown in Fig. 1, can be a useful unmanned aerial vehicle (UAV) for land
observation, surveillance, space vehicle retrieval, etc. Although it is subject to wind, its light and foldable wing
or canopy makes it portable equipment for both civil and military uses. Actually PPG-type UAVs have been
developed and sold on a commercial basis. Such a UAV needs an autonomous/automatic flight control system, and
in order to design a control system we need a dynamic model of a PPG, unless we take an empirical design approach,
which is often employed for a black-box plant. However, a PPG is not a totally unknown plant. In fact, many papers
have been reported on modeling of PPG or paraglider (PG) nonlinear/linear dynamics.1-10 The authors also proposed
a nonlinear dynamic model with eight degrees of freedom, where all the internal forces between the canopy and the
suspended payload are analytically eliminated to obtain a model in the form of state equation.11 Its validity was
demonstrated through numerical simulation and comparison of the results with flight experiment data of a manned
paraglider.
In this paper, first we derive a linear model from the nonlinear model by analytical first-order approximation of
Taylor series expansion as well as numerical partial differential. Since the original nonlinear model is described with
the state variables of the canopy, the linear model also explicitly expresses motion of the canopy. However,
measurement sensors such as accelerometers and gyros are mounted on the payload; hence, it would be more
convenient in dynamics analysis and/or control system design, if the model is expressed by the state variables of the
payload. This motivated us to derive a linear transformation matrix between the canopy states and the payload states.
With the state transformation, we can obtain a payload-state linear model from a canopy one, and vice versa. Then,
we consider designing a flight control system for the PPG using the linear model. Although a number of studies on
PPG or PG flight control have been reported, many of them employ complicated control methods such as model
predictive control12 or inversely consider just open-loop control.13, 14 In our study, we employed PID (proportionalintegral-derivative) control, which is most commonly used in industry. The authors are proposing a new design
method of a PID controller15 based on integral-type optimal servomechanism (IOS), which is a derivative of the
linear quadratic regulator or LQR. In the design method, first we reduce a given linear plant model to a second-order
P

Professor, Department of Aerospace Engineering, 1-10-20 Hashirimizu, Senior Member AIAA.


Graduate student, Department of Aerospace Engineering, 1-10-20 Hashirimizu.
3
Graduate student, Department of Aerospace Engineering, 1-10-20 Hashirimizu.
1
American Institute of Aeronautics and Astronautics
TP

PT

2
TP

PT

TP

PT

Copyright 2009 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

system so that -gap16 between the original plant and the reduced plant becomes as small as possible. The plant
parameters are determined by parameter-space search, since the number of parameters is just two or three. Of course,
the three-term controller cannot always stabilize any plant; however, when stabilizable, it usually provides a set of
control gains with good control performance and substantial stability margins, which would be attributed to the
properties of the LQR. Moreover, the control gains can be adjusted through selection of weighting matrices for
trade-off between control performance and robust stability. In application to a PPG, three PID controllers are
designed to control the altitude by thrust, the forward speed by collective brake deflection of the canopy, and the
heading angle by differential brake deflection. Time responses to appropriate reference outputs are computed as well
as stability margins for each control loop, where other loops are closed.
This paper is organized as follows. In the next section, we describe an outline of the linear model and
transformation between the canopy-state model and the payload-state model. Then we present the design method of
a PID controller based on the IOS and the -gap metric. In Sections IV and V, numerical analysis and simulation
results are presented, followed by summary and conclusions in Section VI.
P

Canopy

Air-Intake
Suspension
lines
Propelling unit
Control line
Payload
Figure 1. Configuration of PPG

Figure 2. Coordinate systems

II. Linear Dynamic Model and State Transformation


A. Coordinate Systems and Degrees of Freedom of Motion
Figure 1 shows coordinate systems of the PPG. The inertial coordinate system I is defined as (XI, YI, ZI). The
XIYI-plane is horizontal and the positive direction of the ZI- axis is taken vertically downward. The location of the
origin and the positive direction of the XI-axis are appropriately chosen. Definition of the canopy coordinate system
c = (Xc, Yc, Zc) and the payload coordinate system p = (Xp, Yp, Zp) is also shown in Fig. 2. The origin Oc of the
canopy-fixed coordinate system is chosen at the center of mass (CM) of the canopy and the Zc-axis is chosen in the
direction from Oc to Om. The Xc-axis is perpendicular to the Zc-axis in the symmetry plane of the canopy and the
positive direction of the Xc-axis is taken forward. The Yc-axis is defined so that the coordinates form a right-hand
coordinate system. The payload is assumed to be symmetric, and the origin is taken at the CM. The Xp-axis is taken
forward along the thrust direction, the Zp-axis is downward and perpendicular to the Xp-axis in the symmetry plane,
and then the Yp-axis is defined to form a right-hand system. It is assumed that the XcZc-plane corresponds to the
XpZp-plane in a trimmed straight flight.
The payload is connected with the suspension lines of the canopy at two points, OmR and OmL. Om is the middle
point between OmR and OmL. It is assumed that the motion of the canopy has six degrees of freedom (DOF) and that
the suspension lines are deformed only about the Zc-axis. The payload then has two DOF, i.e., the relative yawing
about the Zc-axis and the relative pitching about the line OmROmL. Thus, the PPG is modeled as a system with eight
DOF.
B

2
American Institute of Aeronautics and Astronautics

B. Nonlinear and Linear State Equations


Let the velocity of the canopy at Oc represented in c be Vc = [uc vc wc]T, and the angular velocity of the canopy
also in c be c = [pc qc rc]T. Let the relative yaw angle and rate between the canopy and the payload be pc and
rpc=dpc/dt, respectively, and similarly let the relative pitch angle and rate be pc and qpc=dpc/dt, respectively. In
addition, let the pitch and roll angles of the canopy be c and c, respectively. We then define the state vector xc = [uc
vc wc pc qc rc qpc rpc pc pc c c]T = [VcT cT pcT pc pc c c]T, where pc = [qpc rpc]T. The actual control inputs are
thrust Fpthx, the right and left brake angles, R and L, which are deflection angles of the right and left rear-portions of
the canopy. The right (left) brake angle is defined as dR/cc (dL/cc,), where dR (dL) is the pull-length of the right (left)
control lines and cc is the mean aerodynamic-chord length of the canopy. For convenience and for derivation of
linear state equation, we define collective deflection angle e = R + L, and differential one r = R L. We then
define the control input vector as u = [Fpthx e r]T. With the state and control vectors, the 8-DOF motion of the PPG
is described by the nonlinear state equation,11
(1)
x& c = f ( xc ) + g ( xc )u
B

PB

PB

PB

where f(xc) 121 and g(xc) 123 are nonlinear functions of xc.
Let the equilibrium state and input vectors be x* and u*, i.e.,
f(xc*) + g(xc*)u* = 0.
(2)
Under the assumption of symmetry of the PPG with respect to the XZ-plane, the non-trivial equations in Eq. (2) are
four equations corresponding to forward and downward force trim and pitching moment trim of the canopy and the
payload. Defining small perturbations of the state and control variables around the equilibrium point (xc*, u*) as xc
= xc xc* and u = u u*, and applying the first-order approximation of Taylor series expansion, we obtain the
linearized state equation,
(3)
x& c = Ac x c + Bc u ,
B

PB

PB

PB

PB

where Ac1212 is a constant matrix given by partial differentiation of f(xc) + g(xc)u with respect to xc at the trim
point (xc*, u*), and Bc123 is given by g(xc*). We have analytically derived the linear state equation, although we
have no space to show details here. Since the velocity of the PPG is small and its angle of attack is relatively large in
trim flight, we approximate small perturbation of the angle of attack by
(cos c* ) 2
(4)
(wc tan c* uc ) ,
c =
u c*
where c = tan1(wc/uc). This approximation provides a better correspondence of the system matrix Ac between the
analytical linearization and the numerical one.
Appropriately rearranging the elements of the state vector makes the matrices Ac and Bc block-diagonal, so that
Eq. (3) is separated into two state equations, i.e., for the longitudinal motion we have
(5)
x& clong = Aclong xclong + Bclong ulong
B

PB

PB

and for the lateral-directional motion


x& clat = Aclat x clat + Bclat u lat ,

(6)

where xclong = [uc wc qc c qpc pc] , ulong = [Fbth e] , xclat = [vc pc rc c rpc pc c] and
ulat = r.
T

C. State Transformation
Since the canopy generates most of the aerodynamic forces and moments, using the canopy states makes the
derivation of the dynamic equation easier and straightforward. However, sensors such as accelerometers, gyros,
camera, etc. are mounted on the payload; hence, expressing the state equation in terms of the payload states, which
are defined below, would be more convenient for analysis and synthesis of a flight control system. Although we
could derive a nonlinear payload-state equation, we choose to derive a linear payload-state equation, considering
application of a linear control method.
The following relations hold between the canopy states and the payload states. First, the velocity of the payload
expressed in c is given by
(7)
Vpc = Vc + Kpc1c + Kpc2pc.
In Eq. (7), when we define lc (lp) to be the distance between Om and Oc (Op), Kpc1 and Kpc2 are defined, respectively,
as
B

3
American Institute of Aeronautics and Astronautics


l p cos pc + lc
0

= (l p cos pc + lc )
0
l p sin pc sin pc l p sin pc cos pc

K pc1

l p sin pc sin pc

l p sin pc cos pc

(8)

and
cos pc cos pc

= l p cos pc sin pc
sin pc

sin pc sin c

K pc 2
sin pc cos pc .

T
Defining the velocity of the payload expressed in p as Vp = [up vp wp] , we can also write Vpc as
Vpc = TpcVp
where Tcp is a coordinate transformation matrix from p to c and defined as
cos pc cos pc sin pc sin pc cos pc

Tcp = cos pc sin pc cos pc sin pc sin pc .

sin pc
0
cos pc

Eliminating Vpc from Eqs. (7) and (10), we have the first relation:
Vc = TcpVp Kpc1c Kpc2pc.
Second, the relation between the angular velocities is given by
c = Tcpp Kpc3pc,
where p = [pp qp rp]T is the angular velocity vector of the payload in p, and Kpc3 is defined by
sin pc 0
K pc 3 = cos pc 0 .
0
1
B

(10)

(11)

(9)

(13)

(12)

(14)

Third, from the definition of Euler angles and relative attitude angles between the canopy and the payload, we have
the identical equation
TcI(c, c, c) =Tcp(pc, pc)TpI(p, p, p),
(15)
where c is the yaw angle of the canopy, and p, p, and p are the roll, pitch, and yaw angles of the payload,
respectively. TpI (TcI) is the coordinate transformation matrix from the inertial coordinates to p (c).
Thus, we have obtained three relations, which are given by Eqs. (12), (13), and (15). Applying the first-order
approximation of Taylor series expansion to the equations yields a linear relation between the canopy state vector
and the payload one as follows.
First, from Eq. (12) we have
(16)
Vc + Kpc1*c + Kpc2*pc = Tpc*Vp +Kpc4*[pc pc]T,
where the superscript * means that the functions and the variables are evaluated at the trim states, and Kpc4* is
defined as
*
sin( *p pc

)
0
(17)
*
*
* ,
K pc 4 = V p
0
cos( *p pc
)
*
cos( *p pc

)
0

*
1
*
*
where p = tan (wp /up ). Next, Eq. (13) is linearized as
(18)
c = Tpc*p Kpc3*pc.
Finally, from Eq. (15) the following linear equations are obtained:
(19)
p = c+pc

c
*
*
0
p
1 cos c sin c
(20)
.


pc
*
*
*
*

sin
cos
cos

cos
p

pc
pc
p
p

c
B

PB

PB

PB

PB

PB

PB

PB

PB

PB

PB

Adding c to the state vector xc, defining the payload state vector as xp = [VpT pT pcT pc pc p p p]T, and
combining Eqs. (16), (18), (19), and (20), we obtain the relation
xp = Txc,
(21)
where xp =xp xp* and T 1313 is a constant transformation matrix.
With Eq. (21), Eq. (3) is transformed into the equation expressed by the payload state as
B

PB

PB

4
American Institute of Aeronautics and Astronautics

PB

PB

x& p = Ap x p + B p u ,

(22)

where Ap = TAcT and Bp = TBc. As Eqs. (5) and (6) resulted from Eq. (3), Eq. (22) is rewritten into the
longitudinal state equation
(23)
x& plong = Aplong x plong + B plong ulong
B

and the lateral-directional state equation


x& plat = Aplat x plat + B plat ulat ,

(24)

where xplong = [up wp qp p qpc pc] and xplat = [vp pp rp p rpc pc p] . This means that the
state transformation matrix itself becomes block-diagonal by rearranging the state variables.
T

III. PID Control based on Integral-Type Optimal Servomechanism


A. Plant Representation
Consider a second-order delay system described by a transfer function GC(s), i.e.
y(s) = GC(s)u(s),
(25)
where y(s) is an output and u(s) is an input, and GC(s) is defined as
n0
.
(26)
GC ( s ) = 2
s + d1 s + d 0
The plant is obviously controllable and observable. To give a state-space representation of the system, state vector is
defined as x= [x1 x2]T = [ y y& ]T . The system is then expressed as
(27)
x& = Ax + Bu
y = Cx,
(28)
where
1
0
0
, B = , C = [1 0] .
(29)
A=

d 0 d1
n0
B

PB

B. Integral-type Optimal Servo Controller


The first step is to design an integral-type optimal servo controller for the plant given by Eqs. (27) and (28). In
this study, we employ a design method of an IOS controller by Smith and Davison.17 With the definition of control
error e = r y, where r is a constant reference output, we defined an augmented system as
P

d x& A
=
dt e C

0 x& B .
+
u&
0 e 0

(30)

For a quadratic cost function defined as

J = ( x& T Q x x& + Qe 2 + Ru& 2 ) dt ,

(31)

the optimal control law that minimizes J can be derived as


u& = K x x& + K e e
by using the LQR theory. Integrating Eq. (32) yields
u = K x x + K e ed .

(32)
(33)

Defining Kx = [KP KD] and KI = Ke, we can rewrite the control law as
u = K P y + K D y& + K I ed .
B

(34)

This is exactly the I-PD (proportional and derivative preceded integral) control law, which means that the PID gains
are determined at a time as optimal state feedback gains. Once the I-PD control law is obtained, it can be arbitrarily
converted to the PID, PI-D or generic 2-DOF form.18 Note that any of the controller forms has the same closed-loop
property such as stability. In practical use, the pure derivative of y is replaced with an approximate one s/(TDs+1),
where TD is an appropriate constant.
P

C. Plant Reduction based on -Gap Metric


As stated above, we can design an optimal PID controller, if the plant dynamics are given by Eq. (25). However,
the order of a plant GP(s) is generally higher than the second. For a higher-order plant, we need to reduce its transfer
function to Eq. (26) and to achieve this we employ the model reduction technique based on the -gap metric.
B

5
American Institute of Aeronautics and Astronautics

Specifically, we search plant parameters of GC(s) that minimize the -gap between GP(s) and GC(s). It is known that
if the -gap is small, closed-loop properties for the two plants are close to each other.19, 20 Reference 21 also uses this
method in conjunction with LMI to find a reduced-order plant for controller design. In our study, parameter-space
search is adopted, since the number of parameters is small.
Before giving the definition of -gap, let us define the following function
B

G P ( s ) GC ( s )

(G P ( s), GC ( s)) =

(35)

1 + G P ( s)G P ( s) 1 + GC ( s)GC ( s)

and the conditions

(36)
1+GC(j)GP(j) 0,
(37)
wno (1+GC(j)GP(j))+(GP(s)) (GC(s)) 0(GC(s)) = 0,
where (GP(s)) denotes the number of poles of GP(s) in the open right-half complex plane and 0(GC(s)) the number
of imaginary axis poles of GC(s). wno denotes the winding number valuated on the standard Nyquist contour
indented around any imaginary axis poles of GP(s) and GC(s). With the definitions, the -gap is defined as
(GP , GC )
if Eqs. (36) and (37) are satisfied.
(38)
(GP , GC ) =
otherwise.
1

From our design experience, we have found that d0 of GC(s) can be set to zero for GP(s) having integral element
1/s. This also reduces the number of parameters to be searched. Note that depending on GP(s), there exists no GC(s)
for which the -gap is small. In that case, we cannot design a PID controller that possesses good closed-loop
properties or even stability. However, this is natural, since the three-term controller cannot stabilize all plants.
B

Thus, the design procedure is summarized as follows.


1) Approximately cancel closely located poles and zeros of GP(s), if necessary.
2) Determine the parameters of GC(s) based on the -gap metric via parameter-space search.
3) Design an integral-type optimal servo using GC(s).
Features of the design method are as follows.
1) PID gains are obtained at a time.
2) Trade-off between control performance and control effort can be performed through weight selection.
3) Trade-off among P, D, and I actions is also possible through weight selection of the diagonal elements of Qx and
Q, respectively.
4) The controller has desirable properties as an LQR such as good time response and desirable stability margins,
although some degradation may occur due to plant model reduction.
B

IV. Linear Dynamics Analysis


We use data of a manned paraglider22 for numerical analysis and simulation. The span and the chord length of
the canopy are 9.95 m and 3.05 m, respectively, and the weights of the canopy and the payload are 6.4 kg and 93.0
kg, respectively. The trim condition without control inputs is steady-state gliding, where the airspeed is 9.54 m/s, the
angle of attack is 15.1 deg, the pitch angle of the canopy is 3.45 deg, and the relative pitch angle is 4.83 deg. The
system and control matrices of the linear dynamic model were computed by analytical linearization as well as by
numerical partial differentiation. The two methods gave almost the same results. The matrices of the state equations
and state transformation matrices are given in Appendix.
To illustrate the difference between the canopy-state system and the payload-state system, we computed time
responses for the initial relative pitch angle pc(0) = 0.1 rad, where other initial states were zero. Figures 3 shows
time responses of the pitch angle of the canopy, c. Although the same initial states are given, the time responses of
c are very different. The reason for this is that the initial pitch angle c(0) for the payload-state system is 0.1 rad,
whereas it is exactly zero for the canopy-state system, where b(0) = 0.1 rad. Computing time responses of the
lateral-directional motion for the initial relative yaw angle pc(0) = 0.1 rad, we observe the similar difference
between the canopy-state and payload-state systems. Thus, we should consider physical meaning of the states, when
we deal with the systems.
Table 1 summarizes the eigenvalues of the linear models of the PPG and motion modes corresponding to the
complex conjugate pairs. Figure 4 shows time responses of the pitch rate of the canopy computed using the payloadstate model for pc(0) = 0.1 rad. We can see from the figure that the relative pitch rate has a faster, good-damping
mode, and that the canopy pitch rate has slower, more oscillatory mode. Hence, the former mode corresponds to the
P

6
American Institute of Aeronautics and Astronautics

relative pitching motion and the latter to the pitching motion of the canopy. Figure 5 shows time responses of the
yaw rate of the canopy computed using the payload-state model for pc(0) = 0.1 rad. We can also see from the
figure that the relative yaw rate has a faster, oscillatory mode, and that the canopy yaw rate has a slower, less
oscillatory mode. The former mode corresponds to the relative yawing motion and the latter to the Dutch-roll motion
of the canopy.

Canopy pitch angle, deg

2
0
-2
canopy-state system
payload-state system

-4
-6

10

20

time, s

Figure 3. Time responses of the canopy pitch angle for pc(0) = 0.1
B

Pitch rate, deg/s

Table 1 Eigenvalues of the linear systems (longitudinal and lateral-directional)


motion mode
eigenvalues
natural frequency, rad/s damping ratio
first-order modes
1.96, 0.557

relative pitching
3.89
0.863
3.361.97i
canopy pitching
0.983
0.290
0.285 0.941i
first-order modes
121, 1.35, 0

relative yawing
5.03
0.0297
0.1505.03i
Dutch-roll
0.755
0.297
0.2240.721i

qc
qpc

2
0
-2
-4
0

10

20

time, s

Figure 4. Time responses of the canopy pitch rate for pc= 0.1 rad

Yaw rate, deg/s

30
20
10
0
-10
-20
-30

rc
rpc

10

20

time, s

Figure 5. Time responses of the canopy yaw rate for pc= 0.1 rad
B

7
American Institute of Aeronautics and Astronautics

V. Flight Control System Design and Simulation


A. Flight Control System Design
The design method of a PID controller has been applied to the linear model of the PPG. Since the PPG has three
control inputs, we design three PID controllers for respective SISO systems as follows.
The transfer function from thrust, Fpthx(s), to the altitude variation, h(s), is given by
7.1214 10 5 ( s + 180.28)( s + 2.0410)( s + 0.4408)( s 2 + 6.2294 s + 14.678)
(39)
G P1 ( s ) =
s ( s + 1.9578)( s + 0.5567)( s 2 + 6.7208s + 15.170)( s 2 + 0.56976 s + 0.96647)
We have then obtained a second-order plant with a -gap of 0.0177 as
0.01061 .
(40)
GC 1 ( s ) = 2
s +s
Choosing the weighting matrices as Qx = 022, Q =100, and R =1 and the time constant as TD =0.025 s, the resultant
PID control gains are KP = 52.33, KI = 10.00, and KD = 42.67.
The transfer function from collective brake deflection angle, e(s), to the forward speed, up(s), is given by
0.19986( s + 69.954)( s + 12.736)( s + 1.5482)( s 2 + 2.5840 s + 4.0135)
(41)
G P 2 (s) =
( s + 1.9578)( s + 0.5567)( s 2 + 6.7208s + 15.170)( s 2 + 0.56976 s + 0.96647)
We have then obtained a second-order plant with a -gap of 0.298 as
69.24 .
(42)
GC 2 ( s ) = 2
s 6s 20
Choosing the weighting matrices as Qx = 022, Q =1, and R =1 and the time constant as TD =0.025 s, we obtained
PID control gains of KP = 0.9671, KI = 1.000, and KD = 0.2749.
The transfer function from differential brake deflection angle, r(s), to the heading angle, p(s) +p(s), is
given by
35.434( s + 0.30072)( s 2 + 2.1269 s + 37.670)( s 2 + 1.8946 s + 8.5392)
(43)
G P 3 ( s) =
s ( s + 121.0)( s + 1.347 )( s 2 + 0.2990 s + 25.35)( s 2 + 0.4488 s + 0.5698)
Before designing a PID controller, we first design a roll damper, whose control law is r = 0.5pp. Figure 6 shows
time responses of the canopy roll rate, pp, to the step input r = 0.1 rad with and without the roll damper. It is
obvious that the damping of the roll rate is improved.
B

Payload roll rate, deg/s

10

without roll damper


with roll damper

5
0
-5
0

10

20

time, s

Figure 6. Time responses of the roll rate to step input, r = 0.1 rad
B

The transfer function of the closed-loop system then becomes


35.434( s + 0.30072)( s 2 + 2.1269 s + 37.670)( s 2 + 1.8946 s + 8.5392)
(44)
G P 3dmp ( s ) =
s ( s + 119.09)( s + 2.6908)( s 2 + 0.79272 s + 28.537 )( s 2 + 0.47395 s + 0.26189)
For GP3dmp(s), we have then obtained a second-order plant with a -gap of 0.310 as
1.288 .
(45)
GC 3 ( s ) = 2
s + 0.9 s
We designed a PID controller for the reduced-order model of Eq. (45), choosing the weighting matrices as Qx = 10I2,
Q =10, and R =1 and the time constant as TD =0.025 s. The obtained gains are KP = 6.193, KI = 3.162, and KD =
3.785.
B

8
American Institute of Aeronautics and Astronautics

Stability margins are summarized in Table 2. The margins are computed by opening each closed-loop, while
other loops are closed. The controllers have large gain and phase margins. The margin of gain increase is for the
speed controller.
controller
altitude
speed
heading

Table 2 Stability margins


gain margin
phase margin
crossover frequency
28.63 dB
89.09 deg
0.5150 rad/s
42.44 deg
9.172 rad/s
18.35 dB
85.05 deg
2.597 rad/s

In this design example, we use the PID control law


K

u ( s ) = K P K D s + I e ( s ) .
s

where e = yref y and yref is given by


B

(46)

y ref ( s) =

2
nref

2
s 2 + 2 ref nref s + nref

r.

(47)

We have chosen ref = 1 and nref = 0.5.


A problem with flight control of a PPG is input saturation. Since the control lines can only be pulled and cannot
be pushed, the right and left brakes, R and L, can take a positive angle only. Hence, the control inputs tend to
become saturated. Particularly, when a controller has an integral action, the control system may suffer integralwindup due to input saturation, which can significantly degrade control performance. To alleviate the degradation,
we have employed an anti-reset-windup scheme23 and modified the PID control law as
k
KDs2 KPs + KI
(48)
u ( s) =
e( s ) +
u ( s )
s+k
s+k
where u(s) is a controller output and u ( s) is a plant input; namely, if u(s) violates input range limits, u ( s) takes the
maximum or minimum value, and otherwise, u(s) = u ( s) . k is a constant design parameter, and we have chosen
k=10.
B

B. Simulation
We conducted computer simulation, applying the controllers to the nonlinear model of the PPG. The reference
outputs are 0 m for the altitude variation, 8 m/s for the forward speed, and 90 deg for the heading angle. The initial
condition is the steady-state gliding, where the forward speed is 9.22 m/s and the heading angle is 0 deg. As seen
from the initial and the target outputs, this simulation considers the case where the PPG changes its flight from the
steady-state gliding to level flight with thrust, while turning to the right by 90 deg. Figures 7 through 11 show time
histories of the outputs. The altitude undergoes undershoot about 2.4 m, since the PPG is going down at the initial
time; however, it recovers to zero by increasing thrust to achieve level flight. Also, as is expected, the forward speed
is controlled to 8 m/s, and the heading angle is changed by 90 deg. Time histories of the control inputs are shown in
Figs. 12 and 13. Note that the brake deflection angles are saturated, since the neutral deflection angle is zero and the
inputs cannot take negative angle. In this design example, we have added the anti-reset windup compensator and it
really works. In fact, without it control performance considerably degrades as Fig. 14 shows. The thrust converges to
about 200 N, which is required to maintain level flight.

9
American Institute of Aeronautics and Astronautics

Speed, m/s

10
up
vp
wp

5
0
-5

10

20

time, s

30

40

50

Angular rate, deg/s

Figure 7. Time histories of the payload speeds

30
20
10
0
-10
-20

pp
qp
rp

10

20

time, s

30

40

50

Relative angular rate, deg/s

Figure 8. Time histories of the payload angular rates

40
relative pitch rate, qpc
relative yaw rate, rpc

20
0
-20
-40

10

20

time, s

30

40

50

Figure 9. Time histories of the relative angular rates

Euler angle, deg

100
80

roll angle, p
pitch angle, p
yaw angle, p
reference for p

60
40
20
0
0

10

20

time, s

30

40

Figure 10. Time histories of the Euler angles

10
American Institute of Aeronautics and Astronautics

50

Altitude variation, m

0
altitude variation
reference output
-5

10

20

time, s

30

40

50

Brake deflection angle, deg

Figure 11. Time histories of the altitude variation

10
left deflection angle, L
right deflection angle, R

5
0
-5

10

20

time, s

30

40

50

Figure 12. Time histories of the brake deflection angles

Thrust, N

300
200
100
0

10

20

time, s

30

40

50

Figure 13. Time histories of the thrust

Euler angle, deg

100
roll angle, p
pitch angle, p
yaw angle, p
reference for p

50

0
0

10

20

time, s

30

40

50

Figure 14. Time histories of the Euler angles without anti-reset windup compensation

11
American Institute of Aeronautics and Astronautics

VI. Conclusions
We have obtained a linear dynamic model of a PPG, which is analytically derived or numerically computed from
the nonlinear dynamic model. We have also derived the state transformation matrix between the canopy states and
the payload states. With the matrix we can obtain a linear model expressed by the payload states from the linear
model by the canopy states, and the payload-state linear model facilitates dynamical analysis and controller design,
since the measurement sensors and cameras are mounted on the payload. The design example of PID controllers and
the simulation results illustrate that the design method based on the integral-type optimal servomechanism and the gap metric actually works as an efficient design tool of a PID controller with good control performance and
adequate stability margin. A problem with PPG flight control is saturation of the brake deflection angles. We may
need to add an anti-reset windup compensator, in addition to carefully giving reference outputs.

Appendix
The coefficient matrices in the linear state equations, Eqs. (5) and (6), are as follows:
2.4840
2.8799 0.2382 0.61035
1.2966 10 3
0.53697 0.071012

0.23140 1.5516
4
1.1585
0.86154 0.12228 0.38697
8.2210 10

1.4230 10 3
0.02926
0.19233 0.98936 1.0546 0.38030 0.66983 ,
Aclong =
Bclong =
0
0
1
0
0
0
0

0.29554 1.4250
2.8436
1.0502
6.7271 15.844
1.2710 10 3

0
0
0
0
1
0
0

32.647

18.646
5.2030

4.8906

0.19269
1.1958
3.2798
5.8945 10 4 0.091265 0
2.2418
0.024676

0.0059881
4
9.7908 10
0.078414 0
1.1901
0.019645 0.062424 0.77299

20.311
529.09
0.64332
0 ,
152.59
122.48 2.5409 10 4 7.9074 10 3

Aclat =
0
0
1
0.060345
0
0
0
0 Bclat =

20.339
529.04
152.59
123.07
0.48119
0.31324
26.106 0

0
0
0
0
0
1
0
0

0
0
0
1.0018
0
0
0
0

Rearranging the state variables appropriately, we can also make the state transformation matrix diagonal; namely,
by the same transformation from Ac to block-diag(Along, Alat), we can transform T in Eq. (21) to a block-diagonal
matrix. Thus, we obtain the following state transformation matrices for the longitudinal states and the lateraldirectional states, respectively,
0.084212
6.8278
0 0.33200 1.6954
0.99645
0.084212 0.99645 0.54898 0
0
9.3928

0
0
1
0
1
0 and
Tlong =

0
0
0
1
0
1

0
0
0
0
1
0

0
0
0
0
0
1

0
0.027958 9.2167
1 6.8498 0.027958
0
0.99645
0.084212
0
0.084212
0

0 0.084212
0.99645
0
0.99645
0

Tlat = 0
0
0
0.99847
0
0.060252
0
0
0
0
1
0

0
0
0
0
1
0
0
0
0

0
.
084236
0
0
.
99674

which constitute a block-diagonal matrix, block-diag(Tlong, Tlat).


B

12
American Institute of Aeronautics and Astronautics

0
0
0 ,

0
0

0
1

References
1

A. Azuma, Flight Dynamics of Paraglider, Lecture Note, University of Tokyo, 1994.


2
Akasaka, T., The study of the flight characteristics of paraglider, Ph.D. Dissertation, Aeronautics and Astronautics Dept.,
Tokai Univ., Kanagawa, 1999 (in Japanese).
3
Muller, S., Wagner, O., and Sachs, G., A High-Fidelity Nonlinear Multibody Simulation Model for Parafoil Systems,
AIAA Aerodynamic Decelerator Systems Conference, CP2003-2120, 2003, pp.149-158.
4
Mooij, E., Wijnands, Q.G.J., and Schat, B., 9 dof Parafoil/Payload Simulator Development and Validation, AIAA
Modeling and Simulation Technologies Conference and Exhibit, Austin, TX, AIAA Paper 2003-5459, Aug. 2003.
5
Slergers, N. and Costello, M., Comparison of Measured and Simulated Motion of a Controllable Parafoil and Payload,
AIAA Atmospheric Flight Mechanics Conference and Exhibit, Austin, TX, AIAA Paper 2003-5611, Aug. 2003.
6
Hur, G.-B. and Valasek, J., System Identification of Powered Parafoil-Vehicle from Flight Data, AIAA Atmospheric
Flight Mechanics Conference and Exhibit, Austin, TX, AIAA Paper 2003-5539, Aug. 2003.
7
Wise, K., Dynamics of a UAV with Parafoil under Powered Flight, AIAA Guidance, Navigation, and Control Conference
and Exhibit, Keystone, CO, AIAA Paper 2006-6795, Aug. 2006.
8
Christiaan Redelinghuys, A Flight Simulation Algorithm for a Parafoil Suspending an Air Vehicle, Journal of Guidance,
Control, and Dynamics, Vol.30, No.3, 2007, pp.791-803.
9
Ricci, W. S., Wendt, T., Brocato, B., and Benney, R., Leaflet Delivery System, AIAA Aerodynamic Decelerator Systems
Conference, AIAA-2003-2139, 2003, pp.276-280.
10
Yakimenko, O. A, On the Development of a Scalable 8-DoF Model for a Generic Parafoil-Payload Delivery
System, AIAA Aerodynamic Decelerator Systems Conference, AIAA-2005-1665, 2005, pp.642-654.
11
Watanabe, T. and Ochi, Y., Modeling and Simulation of Nonlinear Dynamics of a Powered Paraglider, AIAA
Guidance, Navigation, and Control Conference and Exhibit, Honolulu, HI, AIAA Paper 2008-7418, Aug. 2008.
12
Slegers, N., and Costello, M., Model Predictive Control of a Parafoil and Payload System, Journal of Guidance,Control,
and Dynamics, Vol.28, No.4, 2005, pp.816-821.
13
Akasaka, T., Beppu, G., and Azuma, A., An Investigation of Paragliders Turn Mechanism, Journal of the Japan society
for aeronautical and space science, Vol.47, No.540, 1999, pp.9-15 (in Japanese).
14
Nathan Slegers and Mark Costello, Aspects of Control for a Parafoil and Payload System, Journal of Guidance,Control,
and Dynamics, Vol.26, No.6, 2003, pp.898-905.
15
Kondo, H. and Ochi, Y. A Design Method of Feedback Controller based on Integral-type Optimal Servomechanism,
AIAA Guidance, Navigation, and Control Conference and Exhibit, Chicago, IL, Aug. 2009 (to appear).
16
Vinnicombe, G, Frequency Domain Uncertainty and the Graph Topology, IEEE Transactions on Automatic Control, Vol.
38, No. 9, 1993, pp. 1371-1383.
17
Smith, H. W. and Davison, E. J., Design of Industrial Regulators, Proceedings of IEE, Vol. 119, 1972, pp. 1210-1216.
18
Astrom, K. J. and Hagglund, T. The Future of PID Control, Control Engineering Practice, Vol. 9, No. 11, 2001, pp.
11631175.
19
Zhou, K., Essentials of Robust Control, Prentice-Hall, Inc., Upper Saddle River, New Jersey, 1998. pp. 349-373.
20
Vinnicombe, G., Uncertainty and Feedback, Imperial College Press, London, 2001, pp. 115-117.
21
Buskes, G. and Cantoni, M., Reduced order approximation in the -gap metric, Proceedings of the 46th IEEE Conference
on Decision and Control, 2007, New Orleans, LA, pp. 4367-4372.
22
Azuma, A., Characteristics of paraglider, JASPA, 1996 (in Japanese).
23
Kothare, M. V., Campo, P. J., Morari, M., and Nett, C. N., A Unified Framework for the Study of Anti-windup Design,
Automatica, Vol. 30, No. 12, 1994, pp. 1869-1883.
P

13
American Institute of Aeronautics and Astronautics

Potrebbero piacerti anche