Vous êtes sur la page 1sur 21

Dead-zone Non-Linearities

And
Their Effect on Vehicle Lateral Dynamics
By
Sean Brennan
Submitted
May 1999
To
Professor Andrew Alleyne
For the Class
ME 493
1
2 Table of Contents
2 TABLE OF CONTENTS ..................................................................................................................................1
3 OUTLINE OF PAPER.......................................................................................................................................2
4 PUBLISHED WORK CONCERNING DEAD-ZONES...........................................................................3
4.1 MOTIVATION AND EXAMPLES........................................................................................................................ 3
4.2 RELATED NON-LINEARITIES........................................................................................................................... 4
4.3 SITUATIONS WHERE DEAD-ZONES ARE INTRODUCED INTENTIONALLY................................................. 4
5 DYNAMIC DESCRIPTIONS OF DEAD-ZONES ....................................................................................5
5.1 A MATHEMATICAL DESCRIPTION OF A DEAD-ZONE.................................................................................. 5
5.2 DERIVATION OF THE DESCRIBING FUNCTION.............................................................................................. 5
6 VEHICLE DYNAMICS MODEL...................................................................................................................7
6.1 FIRST MODEL: LATERAL DYNAMICS ALONE............................................................................................... 7
7 LIMIT CYCLE ANALYS IS ............................................................................................................................9
7.1 ANALYSIS OF STABILITY OF BICYCLE MODEL VIA DESCRIBING FUNCTIONS......................................... 9
7.2 NYQUIST PLOT OF VEHICLE SYSTEM AND PREDICTION OF A LIMIT CYCLE......................................... 10
7.3 SIMULATION RESULTS................................................................................................................................... 12
8 COMPENSATION OF DEAD-ZONE NONLINEARITIES ................................................................14
8.1 INVERSE OF DEAD-ZONE ............................................................................................................................... 14
8.2 DITHERING OF INPUT ..................................................................................................................................... 16
8.3 BIASING THE CONTROL ................................................................................................................................. 17
9 CONCLUSIONS................................................................................................................................................19
10 BIBLIOGRAPHY..............................................................................................................................................20
11 APPENDICES ....................................................................................................................................................20
2
3 Outline of Paper
In this paper, the dead-zone non-linearity is described and introduced, and its effect
on vehicle dynamics is studied via simulation studies and experiments on scale vehicles.
To begin the introduction, a discussion of published work concerning dead-zone
nonlinearities is given. Included in this discussion is a physical description of a dead-
zone, and examples are given where dead-zones might arise. A description of the
published theoretical framework used to analyze dead-zone and related nonlinearities in
dynamic systems is provided, and situations are described where these nonlinearities are
intentionally introduced. After the initial discussion of published work, a more rigorous
mathematical model for a dead-zone is described, and a derivation of a describing
function for this dead-zone model is provided.
To analyze a dead-zone in regard to vehicle dynamics, a basic vehicle dynamics
model is introduced describing the relationship between lateral position and steering
input. A dead-zone is introduced as an input non-linearity. That is, the steering linkage
that relays the driver steering command to the vehicle dynamics model is assumed to
contain a dead-zone. Nyquist plots of the theoretical model are used for the prediction of
limit cycles. These predictions are validated using simulation studies. Experimental
work is then presented that verifies the limit cycle results.
Finally, compensation of dead-zone nonlinearities is discussed. The first method of
compensation described in the literature uses an inverse technique. Simulation results
showing the effectiveness of this technique are presented, and a discussion on adaptive
routines is given. A second, older technique of oscillating the input by introduction of a
sinusoidal input is presented. Again simulation results are given showing the effects of
this method. Finally, a third technique that biases the control input is presented. A
conclusion then summarizes the main points of the paper.
3
4 Published Work Concerning Dead-zones
4.1 Motivation and Examples
If we examine the plot of the input of a system versus the output, we may observe
the following relationship.
Figure 1: A simple graphical representation of a dead-zone.
The term dead-zone usually refers to the flat zone in the above figure in which a
change in input results in no change in output. Without providing a more formal
mathematical definition of a dead-zone, we consider physical examples of dead-zones to
motivate this analysis. Hydraulic systems often have small dead-zones in order to
prevent fluid leakage and reduce wear and tear [1, 2]. Models of limb motion in animals
and humans often incorporate dead-zone nonlinearities to prevent constant muscle
actuation [2]. In DC motors, a dead-zone effect is often observed due to coulomb friction
at low frequencies [2]. Steering systems on older vehicles and boats may display a dead-
zone characteristic as the rudder or steering wheel linkages become worn [1]. Even the
ubiquitous potentiometer will display dead-zone like characteristics for small increments
of the potentiometer shaft [1].
For systems with a dead-zone type of non-linearity, one might ask how the
controller performance would be affected? For a linear controller, the answer is that the
performance will degrade significantly [1]. It is important to realize that dead-zone
nonlinearities significantly affect the stability and performance of a system. Many drivers
know the additional difficulty of driving a vehicle where the steering wheel is slightly
loose. With the above potentiometer example, the dead-zone not only negates the
infinite resolution of the potentiometer, but it means that the potentiometer sensitivity
to input is very non-linear. This may in turn affect the stability of a circuit incorporating
the potentiometer as feedback [1].
Input
Output
4
4.2 Related Non-Linearities
A dead-zone non-linearity has no memory of previous operating conditions, but
often nonlinearities are encountered do have this memory. These nonlinearities include
backlash and hysteresis. More complex nonlinearities can be obtained by using
combinations of dead-zone, backlash, and hysteresis. The ability of some nonlinearities
to remember previous operating conditions introduces dynamics into the system, and
hence create additional control difficulties. For this reason, the focus of this paper will
remain solely on dead-zone nonlinearities and how these might affect a sample vehicle
system.
Because dead-zone nonlinearities do not have any memory of previous conditions,
a framework is easily developed to analyze dead-zone nonlinearities. Several authors,
most notably Gang Tao and Petar Kokotovic, have extensively studied dead-zone and
related nonlinearities in terms of the stability of closed loop systems. A focus of current
research is the stability of linear systems with an adaptive controller that will identify
the nonlinear section of the closed loop. To perform this identification, an estimate of the
nonlinear term is first formed and an inverse of the nonlinearity is fed back to the input of
an adaptive controller. The controller not only controls the plant, but also has an update
scheme to update the nonlinear estimate. Stability of the controller and convergence of
the updating scheme is proven via Lyapunov-type analysis [3]. This framework for
controlling and adapting to dead-zone nonlinearities has also been extended to backlash
[4], hysteresis [5], and general non-smooth nonlinearities at both the input and output of
linear dynamic systems [6].
4.3 Situations Where Dead-zones Are Introduced Intentionally
Henry Paynter, working on Analog Computers at MIT, wrote an article in the
Lightning Empiricist nearly 40 years ago discussing how hysteresis was the natural
antagonist of noise or oscillation in a slow servo [1]. For systems that have some type of
persistent oscillation near the origin, a dead-zone may be introduced to ignore this
input. Often, dead-zones are introduced into adaptive control systems to prevent the
adaptive algorithm from drifting or becoming unstable when there is little persistence of
excitation. Although in this limitation does not guarantee asymptotic convergence of the
5
adaptive routine, the use of this type of nonlinearity can provide a guarantee of BIBO
stability that would not otherwise be available [7].
5 Dynamic Descriptions of Dead-zones
5.1 A Mathematical Description of a Dead-Zone
We attempt to obtain a mathematical representation of a dead-zone by describing
the dead-zone in terms of the figure shown below.
Figure 2: A formal definition of a dead-zone.
The output, u(t), is related to the input, v(t), by the following expression:
( )
( )

'


< <

l l l
r l
r r r
b ) t ( v , b ) t ( v m
b ) t ( v b , 0
b ) t ( v , b ) t ( v m
) t ( u (1)
Note that if the dead-zone is symmetric about the y axis, then b
l
= b
r
= b, m
l
= m
r
= m.
5.2 Derivation of the Describing Function
In this section, a describing function for a dead-zone is derived. We first assume
that the dead-zone is symmetric about the y-axis. If we make this assumption, we see
that the dead-zone is an odd function, and hence the describing function is found by
solving the harmonic balance equation:
( ) 0 1 ) a ( j G + (2)
We define the describing function (a) as:
( )

d sin sin a
a
2
) a (
0
(3)
Input
Output
b
r b
l
0
m
r
m
l
6
Note that if a < b, then (a) = 0. For a b, we note that over the interval 0 to pi the sine
term looks like the figure below.
Figure 3: A diagram explaining the method used to derive the describing function.
If we notice the symmetry about the line = /2, the describing function integral
becomes
( )

d sin sin a
a
4
) a (
2 /
0
(4)
Note that the term * is given by:

,
`

.
|


a
b
sin *
1
(5)
( )

d sin b sin a m
a
4
d sin 0
a
4
) a (
2 /
*
*
0
(6)
]
]
]

2 /
*
2 /
*
2
d sin b d sin a
a
m 4
) a ( (7)
2 /
*
cos b cos sin
2
a
2
a
a
m 4
) a (

]
]
]

(8)
]
]
]

* cos b * cos * sin


2
a
2
* a
2
cos b
2
cos
2
sin
2
a
4
a
a
m 4
) a ( (9)
]
]
]

,
`

.
|

,
`

.
|

,
`

.
|

,
`

.
|
+
,
`

.
|



a
b
sin cos b
a
b
sin cos
2
b
a
b
sin
2
a
4
a
a
m 4
) a (
1 1 1
(10)
]
]
]

,
`

.
|

,
`

.
|
+
,
`

.
|



a
b
sin cos
a
b
a
b
sin
m 2
m ) a (
1 1
(11)

sin()
0

/2
line of symmetry

b
a
7
]
]
]
]

,
`

.
|
+

,
`

.
|



2
1
a
b
1
a
b
a
b
sin
m 2
m ) a ( (12)
A plot of this describing function is plotted below assuming b to be 0.5 and m to be 1.
We see that as the amplitude increases, the describing function asymptotically approaches
unity as expected, because at larger and larger amplitudes the influence of the dead-zone
becomes more and more diminished.
0
0.25
0.5
0.75
1
0 2 4 6 8
a

(
a
)
Figure 3: A plot of the describing function for a dead-zone
Before examining the effect of the dead-zone nonlinearity on a vehicle, a dynamic
representation of a vehicle must first be introduced.
6 Vehicle Dynamics Model
6.1 First Model: Lateral Dynamics Alone
The motivation of this paper was the study of the effect of vehicle systems and how
dead-zone nonlinearities affect their performance. Central to this study is the choice of
the vehicle model. The vehicle model relating driver steering input to lateral position can
be described with the transfer function [8]:
8
( )
( )
( )
r
C
r
x
f
C
f
x
r
C
f
C s
r
C
2
r
x
f
C
2
f
x m
r
C
f
C
z
I
2
s
2
mV
z
I
L
r
C
f
C
2
V s Lx
r
C
f
VC
2
s
f
C
2
V
2
s
1
s
f
s Y
2 2
r z
mV L V
I

+
,
`

.
|

+
,
`

.
|

+

+

+


,
`

.
|
(13)
where C
af
, C
ar
are the cornering stiffness of the front and rear tires, L
f
and L
r
are the
lengths from the center of gravity to the front and rear axles. I
z
is the moment of inertia of
the vehicle about the z-axis, which is the axis perpendicular to the road surface, m is the
mass of the vehicle, and V is the velocity of the vehicle. For this study, the parameters
for the vehicle were the measured values from a scale vehicle. These values are as
follows:
Values
Velocity 3.0 m/s
Mass 4.045 kg
Z-Axis Moment of Inertia, I
z
0.1193 kg m
2
Distance from C.G. to Front Axle, x
1
0.1387 m
Distance from C.G. to Rear Axle, x
2
0.1893 m
Front Cornering Stiffness C
1
30.36 N/rad.
Rear Cornering Stiffness, C
2
30.36 N/rad.
The substitution of the above parameters results in the transfer function:
( )
( ) 157.5462 42.1344 4.3216
2745.1 173.2 32.6
s
2
s
s
2
s
2
s
1
s
f
s Y
+ +
+ +

(14)
( )
( )
46 . 6 3 s 750 . 9
2
s
204 . 635 s 08 . 40
2
s 7.54
2
s
1
s
f
s Y
+ +
+ +

(15)
If we use P control, we can represent the closed loop system (without a nonlinearity):
Figure 4: A representation of the closed loop
The controller limits can be obtained from a root locus plot of the system with the
characteristic equation:
ref
controller plant
-
+
9
46 . 6 3 s 750 . 9
2
s
204 . 635 s 08 . 40
2
s 7.54
2
s
1
p
K 1 0
+ +
+ +

,
`

.
|
+
(16)
The root locus plot:
Figure 5: The root locus plot showing the pole trends for different P gains.
We find that the system is stable for all gains, K
p
> 35.
A second model was considered that incorporated not only lateral dynamics but also
actuator dynamics of the system. After including the actuator dynamics, a second set of
oscillatory poles are introduced into the system dynamics. It was found after significant
trial and error that the system was not easily stabilizable using only PD type control on
the lateral position. Although a formal Routh-type of analysis my find some regions of
stable gains, trial and error approach shows that this region will be very small . Hence,
some type of state feedback is necessary to stabilize the system using yaw angle of the
vehicle. The use of full state feedback introduces a large amount of additional
complexity that is not the goal of this analysis.
7 Limit Cycle Analysis
7.1 Analysis of Stability of Bicycle Model via Describing Functions
If a dead-zone is introduced into the vehicle dynamics as a dead-zone in the steering
linkage, then the nonlinearity would appear in block diagram form as shown below:
10
Figure 6: A representation of the closed loop
One problem with this representation is that the nonlinearity is in a non-standard form.
We can derive a standard form with the nonlinearity solely in the feedback loop:
u
c
= C(r-y)
= C*r C*P*u
= C*r C*P*Phi(u
c
)
Where C is the linear controller and P is the plant. This can be represented in the
standard form as:
Figure 7: An alternate representation of the closed loop so that the nonlinearity is in the standard form.
From the above figure, we can see that the loop stability and dynamic
performance can be analyzed if we lump the plant and controller transfer functions
together into one term, hereafter referred to as G(s).
7.2 Nyquist Plot of Vehicle System and Prediction of A Limit Cycle
To study the stability of a vehicle with a dead-zone in the steering linkage, we
assume that the dead-zone is an odd function as derived in the previous sections. If we
make these assumptions, then we can examine the Nyquist plot of G(s), and examine
ref
controller plant
-
+

u
c u y
ref
controller
plant
-

u
c
u
y
controller
+
+
G(s)
11
whether or not it intersects the plot of -1/Phi(a) in the same domain, where a can be any
value. Shown below is a sample Nyquist plot of the system with a gain equal to 100.
The values of Phi can range from 0 to 1, and hence any intersection of the Nyquist plot
below with the negative real axis between 1 and infinity will result in a probable limit
cycle. For systems with finite gain margins, we note that the system will only be stable if
the magnitude is less than 1 when the phase is 180 degrees. On the Nyquist plot, this
corresponds to crossing the imaginary axis between 1 and infinity. Hence, ALL
LINEAR STABLIZING CONTROLLERS WITH FINITE GAIN MARGINS
HAVE THE POSSIBILITY OF PRODUCING A LIMIT CYCLE. Realizing that the
bicycle model dynamics will always fit this criterion, it is no longer of concern whether
or not a limit cycle will occur. The question becomes whether or not the limit cycle
amplitude will be significant to the point of deteriorating the vehicle performance.
Figure 8: A sample Nyquist plot of the system with a gain of 100.
To predict the limit cycle amplitude, we utilize the Nyquist plot of G(s). Using a
numerical intersection technique, we determine the frequency and magnitude where the
Nyquist plot crosses the real axis. As an example, the above plot clearly shows an
intersection, we can predict the frequency and amplitude of the limit cycle. The
numerical approximation found that the frequency of the crossover was 11.8738
radians/sec, and the magnitude was 2.9103. We then ask what amplitude a will satisfy
the harmonic balance equation (2). Using the describing function, we then solve for the
limit cycle amplitude of G(s) using a numerical bisection technique. For a proportional
12
controller with gain 100, the amplitude was found to be 0.92. Note that this is the
amplitude of the output of the G(s) which corresponds to the output of the controller, u
c
.
We can obtain the output magnitude of the plant output y by dividing the amplitude of
u
c
by the proportional gain.
7.3 Simulation Results
A simulation was conducted using a dead-zone nonlinearity with the bicycle model
plant described above. The dead-zone was chosen to be b = 0.5 with m = 1. This
corresponds to 28 degrees amplitude of slack in the steering linkage, obviously an
extreme case. The choice of this amplitude was chosen in order to produce a clear limit
cycle. More reasonable dead-zones are simulated in later portions of this paper. A plot
of the simulation output shows a frequency of approximately 11.7 radians/sec and
amplitude of 0.097. The frequency of oscillation is quite close to our predicted value of
11.8 radians/sec. The amplitude of the oscillation is compared below to the value
predicted above, and is clearly quite close.
Figure 9: A sample Nyquist plot of the system with a gain of 100.
If we plot a comparison of the theoretical frequency and amplitude to simulation results
for various gains, we can obtain a better understanding on how we might expect the limit
cycle to change with respect to gain:
13
Figure 10: Theoretical and simulation amplitude and frequency (right) of limit cycles for various gains.
Using the above plots as a reference, key trends relating to the dead-zone behavior
can be discussed. The most important point is that theoretically, the frequency of the
limit cycle is fixed, regardless of controller gains. This agrees with intuition because the
dead-zone has no memory and no dynamic component. The frequency of the dead-zone
in theory is solely determined by the crossover frequency where G(s) has a phase angle of
180 degrees. This frequency is not affected by changing a proportional gain (which
only affects the magnitude plots), and hence we see a predicted line from theoretical
results. The simulated results do not necessarily agree simply because the higher order
terms of the harmonic balance equation are neglected.
If we now vary the limit cycle range (the parameter b), we may examine the effect
of the dead-zone size on the limit cycle behavior. Conducting a simulation study, the
figures below were obtained. Again, the predicted frequency of the limit cycle is not
affected by the dead-zone size. However, the amplitude of the limit cycle scales linearly
with the size of the dead-zone. Again, this agrees with intuition as we may imagine the
front wheels bouncing back and forth through the dead zone in order to stabilize the
vehicle.
Theoretical
Simulation
Simulation
Theoretical
14
Figure 11: Theoretical and simulated limit cycle characteristics for various dead-zone sizes, b.
8 Compensation of Dead-zone Nonlinearities
8.1 Inverse of Dead-zone
In the examples given at the beginning of this paper, it was stated that boats often
display a dead-zone characteristic. Rather than moving the rudder back and forth, boats
with multiple engines may power one motor more than the other to introduce a steer
input, and then move the rudder far enough away to compensate [1]. Clearly, this is not a
very efficient method of driving a ship, but if the dead-zone is small this method is
clearly usable. If we examine the operating point of the system, we see that for small
perturbations about the operating point that the system is linear. If the perturbation re-
enters the non-linear region however, the linear performance will be destroyed. A better
method to compensate for dead-zone is described in [2]. This method involves
compensating for the dead-zone nonlinearity by introducing a coulomb-type nonlinearity.
The placement of this nonlinearity is important, as shown in the following simulations.
The following simulations were conducted to show the effect of introducing a
coulomb friction type of nonlinearity. One simulation was conducted with the
compensation correctly introduced, while the other shows that the incorrect compensation
results in no improvement in performance. The controller gain was chosen to be 100, and
the dead-zone and coulomb friction amplitude were chosen to be 0.5 radians.
Theoretical
Simulation Theoretical
Simulation
15
Figure 12: Incorrect (left figure and diagram) and correct (right) methods of compensating for a dead zone
by introducing a coulomb type nonlinearity.
One possibility in compensating a dead-zone is the use of an incorrect compensation.
Simulations were conducted where the coulomb friction nonlinearity amplitude was low
and high by a factor of 20%. The figures below show the effect of this compensation
mismatch.
Figure 13: Simulated system responses when the coulomb amplitude was mismatched low (left) and high
(right) by a factor of 20% .
16
If we examine more closely the case where the compensation is too high, we see that the
significant difference between correct compensation and overcompensation is simply in
the settling behavior.
Figure 14: A closer examination of the settling of the overcompensated case.
If we compare the plots of overcompensation to the plots of perfect compensation, we
find that there is little difference in the system responses. We conclude that, because the
difference between overcompensation and correct compensation is small, there may exist
a severe difficulty in compensating a dead-zone correctly using an adaptive routine. Tao
and Kokotovic describe methods to perform an adaptive identification of the dead-zone
nonlinearity, and some claims about convergence have been proven for ideal systems [3].
However, we can see from these simple simulation studies that in practice the adaptation
will have a very difficult task identifying the difference between overcompensation and
exact compensation. This is especially true if we consider that the linear model is not
exact for any real-world systems.
8.2 Dithering of Input
One technique described in [1] to compensate for dead-zone is to introduce a high
frequency input to the system with amplitude approximately equal to the dead-zone. This
17
particular method is popular when there may be several types of non-linearities cascaded
or distributed throughout a system. The block-diagram structure for the simple bicycle
model appears as:
Figure 15: A block diagram representation of a system where dithering is used to overcome dead-zone.
The system response for a 100 radian/sec sinusoidal input is:
Figure 16: The system response with 100 radian/sec dither input.
Clearly from the simulation responses, this technique does not work very well, and in fact
can give worse performance than with the dead-zone alone.
8.3 Biasing the Control
One final technique described earlier to compensate for dead-zone is to use a second
control input to introduce a constant disturbance of the system. In effect, this method
simply biases the input of the dead-zone nonlinearity so that the operating point of the
system at equilibrium is outside of the dead-zone.
18
Figure 17: The operating point using the biased controller.
This method assumes that the plant has two inputs available, and that the operator is
satisfied with operating the systems at a steady-state equilibrium point away from the
origin. Also, the response of the system will also be asymmetric; that is, the controller
behavior for positive disturbances will not be the same as the controller behavior with
negative disturbances. We can also conclude that some disturbances, if sufficiently large,
will generate limit cycles. The block diagram form used to test this method is shown
below:
Figure 18: The block diagram of a bicycle model system using two control inputs.
The system response shows that this method does give the predicted converge, albeit at
an equilibrium point away from the origin.
Input
Output
Operating Point
19
Figure 19: The system response using the bias technique.
Note that if we introduce a disturbance into the system, we can easily obtain limit cycle
behavior again. If the disturbances can be kept small and the bias term is acceptable, we
can see that this technique may be useful to prevent limit cycle behavior.
9 Conclusions
A very simple examination was given regarding the effect of dead-zones on a simple
bicycle model using a simple controller. A fair agreement between theoretical and
simulation results was demonstrated over a wide range of controller gains and dead-zone
sizes. Three different methods of compensating for dead-zone are described and
examined via simulation studies. For the most popular technique, inverse compensation,
it was shown that overcompensation and exact compensation reveal very similar system
behavior. It was shown that adaptive routines, although convergent theoretically, will
likely give poor adaptation behavior in practice. A dithering compensation technique
was tested, and showed poor performance. A bias technique was also tested and gave
good convergence. However, for the bias technique we needed to assume that there are
20
two control inputs available and the bias term is large enough that system disturbances
will not push the equilibrium point into the dead-zone.
10 Bibliography
[1] B. Pease, What's all this dead-zone stuff, anyhow, in Electronic Design, vol. 43,
1995, pp. 123-4.
[2] G. Tao and P. V. Kokotovic, Adaptive Control of Systems with Actuator and
Sensor Nonlinearities, vol. 1, 1 ed. New York: John Wiley & Sons, Inc., 1996.
[3] G. Tao and P. V. Kokotovic, Adaptive control of plants with unknown dead-
zones, IEEE Transactions on Automatic Control, vol. 39, pp. 59-68, 1994.
[4] G. Tao and P. V. Kokotovic, Continuous-time adaptive control of systems with
unknown backlash, IEEE Transactions on Automatic Control, vol. 40, pp. 1083-7, 1995.
[5] G. Tao and P. V. Kokotovic, Adaptive control of plants with unknown
hystereses, IEEE Transactions on Automatic Control, vol. 40, pp. 200-12, 1995.
[6] G. Tao, Adaptive control of systems with nonsmooth input and output
nonlinearities, IEEE Transactions on Automatic Control, vol. 41, pp. 1348-52, 1996.
[7] Z. Ding, Robust adaptive control of nonlinear output-feedback systems under
bounded disturbances, IEE Proceedings: Control Theory & Applications, vol. 145, pp.
p.323-9, 1998.
[8] H. Peng and M. Tomizuka, Preview Control for Vehicle Lateral Guidance in
Highway Automation, ASME JDSMC, vol. 115, pp. 679-686, 1993.
11 Appendices
Matlab Files Used to Simulate System

Vous aimerez peut-être aussi