Вы находитесь на странице: 1из 10

ISSN 02805316

ISRN LUTFD2/TFRT- -7596- -SE

Modelling of the ETH Helicopter


Laboratory Process

Magnus Gfvert

Department of Automatic Control


Lund Institute of Technology
November 2001

1. Introduction
This report contains derivation of the dynamics of the ETH helicopter laboratory process, see Figure 1, using the Euler-Lagrange approach. The process is designed at the Automatic Control Laboratory at ETH in Zrich,
see Mansour and Schaufelberger (1986) and Schaufelberger (1990). A description of the setup is found in Morari, M., W. Schaufelberger and A.
Glattfelder (1995). The process is of MIMO type with nonlinear dynamics, and static input nonlinearities, as will be shown below. The present
model is derived with the purpose of accurate simulation of the helicopter
process. It may also prove helpful for nonlinear controller design. Identification of a linear model, and a linear controller design is presented in
kesson, Gustafson and Johansson (1996).

Figure 1

The ETH helicopter laboratory process.

A schematic picture of the process is found in Figure 2. The helicopter


consists of a vertical axle (A), on which a lever arm (L) is connected by a
cylindric joint. The helicopter has two degrees of freedom: the rotation of
the vertical axle (angle ) with respect to the fixed ground, and the pivoting
of the lever arm (angle ) with respect to the vertical axle. Two rotors are
mounted on the lever arm: R1 and R2 , with the resultant aerodynamic
forces giving rise to moments in the and directions respectively. The
voltages u1 and u2 to the rotor motors are the inputs to the process. A
weight is mounted at an adjustable position on the lever arm towards
rotor R2 .

2. Kinematics
It is assumed that the mass distribution on the lever arm is restricted to a
straight line between the rotors, a distance h from the pivot point. Denote
by O P an origo on this line. Let [r x ( R), r y ( R), r z ( R)] denote a point P on
the lever arm parameterized in the distance R from O P , expressed in an
earth fixed reference system with origo in O and oriented with ez along the
3


l1

PSfrag replacements

R1

l2
R2

OP
L

ez
O

ey
ex

Figure 2

Helicopter process configuration.

center axle. Then


r x ( R) = R cos cos h sin cos

r y ( R) = R cos sin h sin sin


r z ( R) = R sin + h cos

(1)

The corresponding velocities are obtained from differentiation of (1) with


respect to time:
v x( R) = R sin cos R cos sin h cos cos + h sin sin
v y( R) = R sin sin + R cos cos h cos sin h sin cos
v z( R) = R cos h sin

(2)

The squared magnitude of the velocity of P is then given by v2( R) = v2x( R)+
v2y( R) + v2z ( R):



v2( R) = R2 2 + cos2 2 + h2 2 + sin2 2 2hR cos sin 2

(3)

3. Energy expressions
The kinetic and potential energies are derived from
Z
1
v2( R) dm( R)
T=
2
Z
V = g r z ( R) dm( R)
4

(4)
(5)

where g is the acceleration of gravity. With (3) inserted this yields for lever
arm
TL =



1 
1  2
+ cos2 2 JL h cos sin 2 mlc + h2 2 + sin2 2 m
2
2
(6a)

VL = mg sin lc + mg h cos

(6b)

and for the axle


1
JA2
2
VA = 0
TA =

(6c)
(6d)

* R
where the inertia of the lever arm JL = R2 dm( R), the center of gravity
* 1 R
* R
of the lever arm lc = m
R dm( R), the lever arm mass m = dm( R),
and the center axle inertia JA have been introduced. The total kinetic and
potential energies are

T = TL + TA

V = V L + VA

(7)
(8)

4. Equations of motion
Forming the Lagrangian

(9)

L=TV
the equations of motion are given by


d VL

dt V


d VL

dt V

VL
=
V
VL
=
V

(10)

Inserting (6) in (10) gives






2 cos sin + cos2 JL +2h sin2 cos2 cos sin mlc


+ h2 2 sin cos + sin2 m + JA = (11a)





+ cos sin 2 JL + h sin2 + cos2 2 + g cos mlc


+ h2 h2 sin cos 2 g h sin m = (11b)

These equations may be expressed on matrix form as


 
 

D ( , ) + C( , , , ) + g( , ) =

(12)

The fundamental property N ( , , , ) = D ( , ) 2C( , , , ) is fulfilled


with the skew symmetric matrix N ( , , , ). The matrices D (), C(), g(),
and N () are defined as in Figure 3.
5

Figure 3

Matrices for Equation (12).


0


2 cos sin JL 2h sin2 cos2 mlc 2h2 sin cos

g cos mlc mg h sin

JL + h2 m



cos sin JL + h sin2 cos2 mlc + h2 sin cos m


2 cos sin JL 2h sin2 + cos2 mlc + 2h2 sin cos m

*
N ( , , , ) =

g( , ) =

cos2 JL 2h cos sin mlc + h2 sin2 m + JA

*
C( , , , ) =


cos sin JL + h sin2 cos2 mlc + h2 sin cos m

cos sin JL + h sin2 + cos2 mlc h2 sin cos m

D ( , ) =

5. Rotors
The rotors are driven by DC-motors without current-control. The motor
operation is described by
d
ia = Ra ia k + u
dt
d
J = Td TL
dt

La

(13a)
(13b)

where Ra and La are the rotor-circuit resistance and inductance respectively, and k the motor constant. The driving moment is described by
Td = kia

(14)

TL = D e e

(15)

and the motor load is described by

where D is the aerodynamic torque coefficient, according to propeller Blade


Element Theory, see e.g. Weick (1926) or, if preferred, a modern textbook
on theory of flight. Combining (1315) in steady-state yields
u0 =

Ra D
0 e 0 e+ k 0  k 0
k

(16)

The approximation is found valid by examining experimental results in


Morari et al. (1995). The second order motor dynamics may then be approximated by first order dynamics as
1
1
d
1 = 1 +
u1
dt
T1
k1 T1
1
1
d
u2
R2 : 2 = 2 +
dt
T2
k2 T2
R1 :

(17a)
(17b)

where T1 and T2 are the time constants of the motors.


The resulting aerodynamic drag forces are given by F1 = C1 1 e 1 e and
F2 = C2 2 e 2 e with C1 and C2 being aerodynamics drag coeffients (Weick
1926). Each rotor affects the helicopter with a moment resulting from the
aerodynamic force, and with a moment that is the reaction moment from
the driving torque of the rotor motor. (The sign of the reaction moments
depend on the configuration of the rotor blades.) Gyroscopic effects of the
rotors are assumed to be small and are neglected. (Any gyroscopic moments resulting from the rotor rotations would mainly include components
perpendicular to the and rotation axis.) For R1 :

and for R2 :

1, = D1 1 e 1 ecos
1, = l1 C1 1 e 1 e

(18a)
(18b)

2, = l2 cos C2 2 e 2 e
2, = D2 2 e 2 e

(19a)
(19b)

= 1, + 2,
= 1, + 2,

(20a)
(20b)

These moments are combined to form

6. Simulation model
The complete set of equations describing the helicopter process is given by
(11) together with (17)(20). Rewriting these on state-space form gives

1
d  2
= cos JL 2h cos sin mlc + h2 sin2 m + JA
dt


2 cos sin JL 2h sin2 cos2 mlc 2h2 sin cos m
+ D1 1 e 1 ecos + l2 cos C2 2 e 2 e]

(21a)

d
=
(21b)
dt

1 
d 
= JL + h2 m
cos sin 2 JL h sin2 + cos2 2 mlc
dt

g cos mlc + h2 sin cos 2 m + mg h sin + l1 C1 1 e 1 e+ D2 2 e 2 e
(21c)
d
=
(21d)
dt
1
1
d
1 = 1 +
u1
(21e)
dt
T1
k1 T1
1
1
d
u2
(21f )
2 = 2 +
dt
T2
k2 T2

7. Equilibrium points
Equations (11), (17) (20) may be solved for stationary points ( 0 , 0, u1,0 , u2,0 )
by setting = = = = 1 = 2  0:
0 = D1 1,0 e 1,0 ecos + l2 cos C2 2,0 e 2,0 e

mg (lc cos 0 h sin 0 ) = l1 C1 1,0 e 1,0 e+ D2 2,0 e 2,0 e

(22a)
(22b)

For the unforced system with =  0 then = 0 , = 0 are


equilibrium points, with arbitrary 0 and tan 0 = lc /h. Since the tan1
function is periodic there are infinitely many equilibrium points S0 . In particular there is one 0 [ /2, /2), and one 0 [ , /2) [ /2, ).
Stability for the lever arm dynamics around ( 0 , 0 ) may be investigated
by regarding the resulting simplification and Taylor expansion of (11b):

mg
mg
[h sin lc cos ] =
[h cos 0 + lc sin 0 ] + O ( 2)
JL
JL

mg cos 0 2
=
h + lc2 + O ( 2) (23)
hJL
*

with = 0 . Thus the stationary point ( 0 , 0 ) with 0 [ /2, /2)


is stable for h <S0, and unstable for h > 0, and the stationary point with
0 [ , /2) [ /2, ) is unstable for h < 0, and stable for h > 0. The
resulting dynamics is a pendulum equation.
8

8. Parameters
Parameters for a real helicopter process are presented in Morari et al.
(1995). Geometric and interial parameters are presented directly. Motor
and rotor properties are presented in graphs resulting from experiments.
The corresponding parameters presented here are computed from the graphs.
Description

Parameter

Value

Unit

Arm length to R1

l1

0.1995

Arm length to R2

l2

0.1743

Mass of lever arm bar

ml

0.280

0.0298

mw

0.158

lw

0.090

m1

0.3792

Mass of rotor R2

m2

0.1739

Time constant for rotor R1

T1

1.1

Time constant for rotor R2

T2

0.33

Motor constant for rotor R1

k1

1.00 102

Motor constant for rotor R2

k2

1.39 102

Aerodynamic drag for rotor R1

C1

2.50 105

Aerodynamic drag for rotor R2

C2

1.58 106

Aerodynamic torque for rotor R1

D1

2.90 107

Aerodynamic torque for rotor R2

D2

1.76 107

[m ]
[m ]
[kg]
[m]
[kg]
[m]
[kg]
[kg]
[s]
[s ]
[Vs/rad]
[Vs/rad]
[Ns2 /rad2]
[Ns2 /rad2]
[Nms2 /rad2 ]
[Nms2 /rad2 ]

Pivot height
Mass of weight
Distance to weight
Mass of rotor R1

(nominal1 )

Table 1

Helicopter model parameters.

The total mass of the lever arm is


m = m l + m1 + m 2 + m w

(24)

The moment of inertia for the lever arm is the sum of the moment of inertia
for the solid lever bar and for the point masses of the rotors and the weight:
JL =

ml l13 + l23
2
+ m1 l12 + m2 l22 + mw lw
3 l1 + l2

(25)

The moment of inertia for the vertical axle may be neglected:


JA  0

(26)

The center of gravity is


lc =
1

m l ( l 1 l 2 ) + m 1 l 1 m2 l 2 mw l w
m

(27)

May be varied in the range 0.0705 0.119 [m].

9. Bibliography
kesson, M. and E. Gustafson and K. H. Johansson (1996): Control Design
for a Helicopter Lab Process, Preprints 13th World Congress of IFAC,
San Francisco, California.
Mansour, M. and W. Schaufelberger (1989): Software and laboratory experiments using coomputers in control education. IEEE Control Systems Magazine, 9:3, pp. 1924.
Morari, M., W. Schaufelberger and A. Glattfelder (1995): Klassische Regelung
eines Helikoptermodells, Fachpraktikumversuch A50, Institut for Automatik, ETH, Zrich, Rev. 6, Nov. 1996.
Schaufelberger, W. (1990): Educating future control engineers. Proc. of
IFAC World Congress. Tallin, Estonia, pp. 3951.
Weick, F. E. (1926): Propeller design I: practical application of the blade
element theory, NACA TN 235, Langley Memorial Aeronautical Laboratory, Washington.

10

Вам также может понравиться