You are on page 1of 17

D.

Erickson
M. Weber
I. Sharf
Department of Mechanical Engineering
McGill University
Montreal, PQ Canada
inna.sharf@mcgill.ca

Abstract
In this paper, we review and compare four algorithms for the identification of contact stiffness and damping during robot constrained
motion. The intended application is dynamics modeling and simulation of robotic assembly operations in space. Accurate simulation
of these tasks requires contact dynamics models, which in turn use
contact stiffness and damping to calculate contact forces. Hence,
our primary interest in identifying contact parameters stems from
their use as inputs to simulation software with contact dynamics capability. Estimates of environmental stiffness and damping are also
valuable for force tracking and stability of impedance controllers.
The algorithms considered in this work include: a signal processing
method, an indirect adaptive controller with modifications to identify
environment damping, a model reference adaptive controller and a
recursive least-squares estimation technique. The last three methods
have been proposed for real-time implementation in impedance and
force-tracking controllers. The signal processing scheme uses a frequency estimate calculated with fast Fourier transform of the force
signal and is an off-line method. The algorithms are first evaluated
using numerical simulation of a benchmark test. Experiments conducted with a robotic arm contacting a flexible wall provide a further
demonstration of their performance. Our results indicate that the indirect adaptive controller has the best combination of performance
and ease of use. In addition, the effect of persistently exciting signals
is discussed.

KEY WORDScontact parameters, identification, estimation, contact dynamics, environment stiffness and damping,
impedance control, adaptive control, recursive least-squares

1. Introduction
Robotic tasks can be classified into two categories: unconstrained and constrained motion. Unconstrained motion occurs when the manipulator is instantaneously free to move in
The International Journal of Robotics Research
Vol. 22, No. 1, January 2003, pp. 41-57,
2003 Sage Publications

Contact Stiffness
and Damping
Estimation for
Robotic Systems

any direction without contacting the environment. Examples


of such tasks include spray painting and visual inspection.
Constrained motion occurs when the manipulator interacts
with its environment through a point or multiple points of
contact. Tasks such as grinding, cutting and assembly demonstrate constrained motion.
Modeling and control of constrained robotic operations
present a number of challenges. Accurate simulation of these
tasks requires the contact dynamics capability in formulating
and solving the motion equations. From the control standpoint, it is widely acknowledged that a force control or combined position/force control strategy is required to ensure desired interaction with the environment. One of the most popular control strategies for constrained tasks is impedance control (Hogan 1985).

1.1. Impedance Control


Impedance control utilizes a single control law which attempts
to regulate both position and force by specifying a dynamic
relationship between them. This relationship is chosen to be
a second-order linear impedance because such systems are
well understood and simple to control. A standard impedance
control law is shown in eq. (1), where Mt , Bt , Kt are the target
impedance mass, damping and stiffness, xr is the reference
trajectory, x is the actual trajectory of the end-effector, and Fe
is the external force applied to the end-effector:
Mt (x x r ) + Bt (x x r ) + Kt (x xr ) = Fe .

(1)

The target impedance parameters are specified by the user


and have the dimensions of the task space. They cause the
manipulator to exhibit the dynamics of a multi-directional
massspringdamper system. The target impedance matrices
are typically chosen to be diagonal, resulting in uncoupled
response along each principle direction of x.
41

42

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003

1.2. Environment Model


For the purposes of the present paper, it is necessary to distinguish two approaches to environment modeling. The first
is used for impedance and force control and it represents the
environment as an n-dimensional spring (n 6), or a spring
and damper, generally referred to as impedance. Thus, when
the manipulator contacts the environment, a reaction force
results which can be defined as
Fe = Ke (xe x) + Be (xe x ).

(2)

Here, Ke and Be are n-dimensional matrices representing the


stiffness and damping characteristics of environment, while
xe represents the environment location. It is usual, however, to
treat each Cartesian variable independently or, in other words,
to assume that the environment impedances in different directions are uncoupled (Seraji and Colbaugh 1997; Singh and
Popa 1995). In this case, representation (2) is replaced by n
scalar equations of the form:
Fe = Ke (xe x) + Be (xe x).

(3)

This one-dimensional version of the environment model is


shown schematically in Figure 1, where xe , Ke and Be are
scalars representing the environment location before contact,
stiffness and damping, respectively. Implicit in the models of
eqs. (2) and (3) is the fact that the interaction between the
robot and the environment is confined to a single point or a
small region, such as the case of robotic grinding operations.
Other common applications of this model are for robot collisions with a wall or motion along the wall. Note that some
researchers also include environment inertia in their models
(missing in our eq. (2)). One of the identification methods
explored in this paper (Love and Book 1995) can directly estimate the inertia of the environment. For many applications,
such as space, it is reasonable to assume that the inertia of
the environment is known. In terrestrial applications, environments tend to be stationary or quasi-static, in which case
the inertia term can be neglected.
In the context of contact dynamics modeling, the robot
environment interaction is distributed over a finite number
of contact points between the robot payload and the fixture.
At each point, a contact force can be defined in terms of its
normal and tangential components as
Fc = Fn n + Ft t.

(4)

One common model used to define the normal component


of the contact force is the spring-dashpot model (Gilardi and
Sharf 2002) of the form:

Fn = Ke + Be .

Fig. 1. Environment model.

(5)

In the above, the contact parameters, stiffness Ke and damping Be , relate the local deformation and its rate to the normal

force at each contact point between the contacting objects. In


the context of contact dynamics modeling, at each contact
point can be calculated from the geometry and location of the
two mating objects. Note that we have retained the subscript
e in denoting contact parameters for uniformity of notation.
In the simplest contact scenario, that of one-point contact
between robot end-effector and fixture, Fc of eq. (4) becomes
the external end-effector force Fe and the contact model in the
normal direction (eq. (5)) is completely analogous to the onedimensional environment model of eq. (3) with = xe x.
In this case, the problems of contact parameter and environment impedance estimation are identical. For general threedimensional contacts with multiple points of contact between
the payload and the fixture, and assuming identical contact
parameters at all contact points, the net contact force can be
written as the sum of individual contact forces
Fe = Ke

N

i=1

i ni + B e

N

i=1

i ni +

N


Fti ti ,

(6)

i=1

where N is the number of contact points and Fti is determined


by the particular friction model. It is noted that the contact parameters employed in eqs. (5) and (6) are by definition scalar
quantities, regardless of the complexity of contact geometry.
For both representations of the environment, in general, the
stiffness and damping parameters are poorly known, while environment location or the contact geometry are either known
or can be determined. This implies that the parameter estimation problem can be formulated as a linear identification
problem of the form
Fe = T ,

(7)

where contains the contact parameters to be estimated and ,


called the regressor, is strictly a function of contact geometry.

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation


The above is true for one-point (one-dimensional) contact or
for general three-dimensional scenarios with multiple points
of contact. In the former case, the regressor is = [xe x, xe
T ) and the vector of parameters is = [Ke , Be ]T .
x]
T (or [, ]
The reader is referred to Weber et al. (2002) for derivation of
eq. (7) for complex contact geometries. In this case, the main
complication is that friction identification is now coupled with
contact stiffness and damping estimation, thus making the
solution prone to approximations and assumptions of friction
modeling.
1.3. Motivation
For certain tasks such as space-based manipulation or nuclear waste remediation, the need for high-fidelity simulation
is well understood. To simulate robotenvironment contact
tasks, the simulation software must incorporate a contact dynamics model, such as in eq. (5), and apply appropriate solution methodology. One such software package is MDSF (Ma
et al. 1997) developed by MacDonald Dettwiler Space and
Advanced Robotics, Ltd.1 A recently completed experimental validation of MDSFs contact dynamics capability (Van
Vliet et al. 2000) has demonstrated that for high-fidelity simulation, good estimates of contact parameters must be known.
The identification of environment characteristics is also an
integral component of force-tracking impedance controllers.
One drawback of position-based impedance control is its inability to directly control the interaction force Fe . Instead,
the contact force results from the target impedance values,
reference trajectory and environment properties. A few researchers have designed force-tracking impedance controllers
which command the robot to exert a specified force on the
environment. For example, Seraji and Colbaugh (1997) developed direct and indirect adaptive control to achieve force
tracking by estimating the environment stiffness and location.
Performance was demonstrated in one dimension by using a
seven-degrees-of-freedom (7-DoF) robot to exert a desired
force normal to the reaction surface, while tracking a desired
trajectory tangent to the surface. Singh and Popa (1995) also
showed that model reference adaptive control (MRAC) could
be applied to impedance force control to achieve force tracking. It was found that, like the algorithm of Seraji and Colbaugh, an estimate of the environment stiffness was required.
Environment estimation can also be used to improve stability during contact. Love and Book (1995) showed how estimates of environment impedance parameters could be used
to design an impedance controller that was stable both during and at the onset of contact. They also showed that, in the
absence of accurate estimates, unstable contact could result.
Their algorithm uses a recursive least-squares (RLS) technique applied to a discrete form of the environment dynamics
equations proposed by An et al. (1988).
1. Formerly SPAR Aerospace, Robotics Division.

43

1.4. About This Paper


In this paper we compare the existing environment parameter estimation algorithms (Seraji and Colbaugh 1997; Singh
and Popa 1995; Love and Book 1995), as well as an original method which uses a signal processing approach, all
applied to the problem of contact or environment stiffness
and damping estimation. We outline some important changes
made to improve the accuracy of parameter estimation of the
indirect adaptive controller. Our comparison is based both
on numerical performance of the methods and implementation on a real system. In addition, an important aspect of
parameter estimationthe requirement for persistent excitation of the reference signalis discussed and demonstrated
experimentally.
Following other authors, our brief presentation of the different methods is limited to one-dimensional (for the environment impedance model) or one-point contact (in the
case of contact model) situations. It is emphasized, however, that although the bench test considered in this paper represents a simple contact geometrysingle-point, onedirectional contactour findings have direct relevance to
multiple-point contacts for complex mating bodies. As noted
earlier, the identification problem for such scenarios can still
be formulated as a linear relationship between the net contact
(end-effector) force and the scalar contact parameters to be
identified. With the exception of the signal processing technique, the other methods investigated here can be directly
applied to identify contact parameters for three-dimensional
contact geometries, using appropriate definitions of regressor
and parameter vector .

2. Environment Estimation Algorithms


2.1. Signal Processing Method
The signal processing method of environment estimation was
developed from the theory of second-order, linear, time invariant systems. A robot that is controlled using impedance
control exhibits a second-order dynamic relationship between
the position of the end-effector and the external force applied
to it. The characteristics of this relationship are governed by
the target impedance values (Mt , Bt , Kt ), through which the
user can specify the dynamic behavior of the manipulator
(Hogan 1985). Also recall that the environment is assumed
to behave as a linear impedance according to its stiffness and
damping coefficients (Ke , Be ). When these two systems come
into contact, a new second-order system is formed (see Figure 2) which is composed of the impedance characteristics of
both the controller and environment. If the impedance characteristics of the combined system (impedance controller and
environment) can be determined, then those of the environment can be calculated.

44

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003

Fig. 2. Second-order system formed by environment and impedance controlled robot.

2.1.1. Determination of System Natural Frequency and


Damping Ratio
The impedance properties of the environmentrobot system
can be determined from the step response of this system. This
is achieved by commanding a step in the location of the robot
end-effector (x) during contact, and measuring the resultant
force (Fe ). Assuming an underdamped response, the damped
natural frequency d can be determined from the fast Fourier
transform (FFT) of the force signal. To determine the damping
ratio , we have used the settling time method which gives
the following relation for Ts to achieve convergence to within
5% of the steady-state value

The solution for is therefore found by counting the number


of visible cycles (including fractional parts) before the measured contact force response converges to within 5% of the
steady-state value, and substituting the result into eq. (11).
2.1.2. Determination of Environment Stiffness and Damping
Coefficients
The stiffness and damping coefficients of the environment
can be extracted from the known values of d and . For the
robotenvironment system in Figure 2, the equivalent stiffness, damping and mass are
Keq

Kt + K e

(12)

(8)

Beq

where n is the undamped natural frequency (Dorf and Bishop


1995). The settling time is then given by

Meq

=
=

Bt + B e
Mt .

(13)
(14)

n Ts

= 0.05,

Ts =

2.996
n

(9)

and the number of cycles of the response before the settling


time is reached is

2.996 1 2
#cycles =
.
(10)
2
From the above, the damping ratio is obtained as
=

0.4768
#cycles + 0.2274
2

(11)

In terms of these parameters, the natural frequency and damping ratio are given by

Keq
d
n =
=
(15)
2
Meq
1

Beq

.
2 Keq Meq

(16)

With the values for d and determined from the force response and equations (12)(16), we obtain the environment
stiffness and damping in terms of the known values:
Ke

Be

n2 Mt Kt

2 (Kt + Ke )Mt Bt .

(17)
(18)

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation


It is important to note the advantages and disadvantages
of this method. To its credit, the signal processing method
requires very little sensory data. Only the (normal) contact
force needs to be measured and it is common for robots that
perform contact tasks to have a force sensor mounted at the
end-effector. This algorithm then processes the force measurements off-line, after the experiments are completed. Unlike
the other methods described below, this scheme does not require measurements of the environment deflection or velocity.
Interestingly, this method can be best described as a hybrid between basic frequency-domain and time-domain techniques,
as it makes use of both types of information. One disadvantage
is that, in order to measure both d and , the force response
must be underdamped. This implies that the target impedance
values must be carefully chosen to give the desired behavior.
Frequency-domain methods in general are not subject to this
limitation, but for our problem they require kinematic data
(i.e. position or velocity). Also, the method cannot be easily
extended to identify contact parameters for general contact geometries because it makes use of the equivalent stiffness and
damping for the robotenvironment system. These cannot be
easily defined for three-dimensional contacts with multiple
contact points, as in the environment model of eq. (6). Finally, the algorithm assumes that the manipulator dynamics
are represented perfectly by the target impedance which is not
achievable in practice.

The indirect adaptive controller proposed by Seraji and Colbaugh (1997) was intended to achieve force tracking within
impedance control. This implies that the measured force (Fe )
should converge to a desired reference force (Fr ). To achieve
this goal, a trajectory generator was developed which modified the reference trajectory (xr ) on-line, during the contact.
The algorithm uses estimates of the environment stiffness and
location (K e , xe ) to calculate xr in the impedance control law
(1) from
xr = xe +

dynamics equations (2), which therefore cannot be written in


the standard linear in parameters form. In light of this, the
indirect adaptive controller was reformulated to estimate environment stiffness and damping (K e , B e ), but not location
(xe ). This is in line with our previous comments that location of the environment is either known (as may be the case
in terrestrial applications) or can be measured, similarly to
the location of robot end-effector already required for most
methods considered here. With this modification, the actual
contact force remains as in eq. (2), the trajectory generator
becomes
xr = xe +

1
Fr
K e

1
Fr .
K e

(19)

The environment estimates are adapted during each time step


according to an update law derived using a Lyapunov-based
approach. This method uses the popular regressor form of the
environment impedance equations. The results presented by
Seraji and Colbaugh (1997) showed that force tracking was
achieved, but convergence of the environment estimates to the
correct values could not be guaranteed. This was attributed to
lack of a persistent excitation condition.
Our experience with numerical simulation, however, revealed that the addition of persistent excitation still did not result in parameter convergence. The cause for this is the choice
of stiffness and environment location as adjustable parameters. These parameters are multiplied together in the contact

(20)

and the estimated force becomes


F = K e (xe x) B e x

(21)

where we make use of constant xe . The force estimate can


now be written in the linear form by defining the regressor
= [xe x, x]
T and the vector of parameter estimates
= [K e , B e ]T



 K e

F = xe x, x
= T .
(22)
B e
Subtracting eq. (3) from eq. (22) yields
F = T

2.2. Indirect Adaptive Control

45

(23)

where F = F Fe and = .
The adaptation law can be formed using a Lyapunov technique (Slotine and Li 1991) by defining the energy function
V , where  is a positive definite matrix:

V = T  .

(24)

If the parameter adaptation law is chosen as


=  1 F

(25)

it can be shown, using eqs. (24) and (25), that the derivative
of the energy function is
V = 2 T T
which is negative definite in terms of provided

In T dt In
for 0 < < .

(26)

(27)

The above condition on the regressor is in fact the definition


of persistent excitation (Anderson et al. 1986). Since eq. (26)
is negative definite in terms of , it can be concluded that the
estimate error 0 as t , or that the environment
estimates (K e , B e ) converge to the actual values (Ke, Be ).

46

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003

If the actual parameter values are assumed to be constant,


then = and the adaptation law (25) can be used to generate new parameter estimates during each time step by numerical integration. The new estimates are then used to update
the trajectory generator (20). The resulting impedance controller achieves both force-tracking and accurate parameter
estimates, the latter in the presence of persistent excitation.
The implementation of the parameter estimation in the indirect adaptive controller requires data on the actual position and velocity of the end-effector (x, x),
and the interaction force (Fe ). In practice, accurate measurement of absolute
kinematics quantities at the robot tip is difficult to achieve and
this poses a challenge in the application of the method. To use
the algorithm, the user must specify the gain matrix  for the
adaptation law (25) and the reference force signal Fr . This
force input must be persistently exciting if accurate estimates
for both the environment stiffness and damping are required.
2.3. Model Reference Adaptive Control
MRAC was used by Singh and Popa (1995) to investigate
some fundamental issues of force control, namely explicit
force control, general impedance control and force-tracking
impedance control. Of main interest to us is the latter since it
is the impedance control combined with explicit force control
that involves environment parameter estimation. In fact, the
work of Singh and Popa demonstrated that to achieve a forcetracking impedance controller, knowledge of the environment
contact parameters was required.
The MRAC controller presented by Singh and Popa is significantly more complicated than the indirect adaptive controller proposed by Seraji and Colbaugh (1997). The full controller is not described in detail here, but a brief explanation
of the method is given below. Similarly to eq. (3) the MRAC
controller is based on the model of the environment as a linear
spring in parallel with a viscous damper. The major components of the controller include the following:

The state dynamics are driven by the model-following control law which involves state feedback (x, x)
and an appropriately chosen auxiliary signal. The virtual trajectory dynamics
are specified by the user, and this signal must be persistently
exciting to ensure convergence of the model-matching parameters (Singh and Popa 1995). Furthermore, it is stated that
environment parameter convergence can be guaranteed if the
desired contact force is time-varying (persistently exciting).
The experimental results presented by Singh and Popa
show accurate estimation of the environment stiffness and
damping for a variety of environments. Simultaneously, a desired contact force is achieved.
2.4. Recursive Least Squares
Love and Book (1995) have successfully demonstrated that
contact stability can be improved if estimates of the environment impedance parameters are known. They modeled
the environment as a locally stationary massspringdamper
system:
Fe = Me x + Be x + Ke (x xe ).

(28)

By defining x = x xe and remembering that the environment is stationary (xe = xe = 0), eq. (28) can be rewritten
as
Fe = Me x + Be x + Ke x.

(29)

The above differs slightly from the equation used by Love and
Book (1995), but instead reflects the original work described
by An et al. (1988). The bilinear transformation is then used
to transform eq. (29) into its discrete-time counterpart

2

2
2
1 z1
Fe = Me
T
1 + z1
(30)



2
1 z1
+ Ke x
+ Be
T
1 + z1

1. The plant model represents the actual dynamics of the


manipulator in contact with environment. It is assumed
that the system is linear and the state-space parameters
are in general unknown.

where T is the sampling period. We can expand eq. (30) into


a polynomial in terms of z. By recognizing that z1 represents
a shift of one step in the time domain, and letting k describe
the time-step index, the corresponding difference equation is

2. The reference model represents the target impedance


with the reference signal specified by the virtual trajectory generator.

Fe[k] + 2Fe[k1] + Fe[k2]




2


2
2
=
Me
+ Be
+ Ke x[k]
T
T

2
2
x[k1]
+ 2 Ke Me
T


2


2
2
+ Ke x[k2]
+ Me
Be
T
T

3. The MRAC parameter estimator provides the adaptation law for system parameters that match the plant
model to the reference model.
4. The environment parameter estimator provides the
adaptation law for the environment stiffness and
damping.

Ax[k] + Bx[k1] + Cx[k2]

(31)

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation

47

which can be cast into the regressor form as


T
[k]
y[k] = [k]

(32)

Fe[k] + 2Fe[k1] + Fe[k2]

x[k] xe
x[k1] xe
x[k2] xe

T
A B C
.

(33)

with
y[k]

[k]

[k]

(34)
(35)

The RLS solution for [k] takes the following form (Ljung
1987)


T
(36)
[k1]
[k] = [k1] + L[k] y[k] [k]

Fig. 3. Manipulator and flexible wall used for the benchmark


test.

where
L[k]

P[k]

P[k1] [k]
T
+ [k]
P[k1] [k]

T
P[k1] [k] [k]
P[k1]
1
.
P[k1]
T

+ [k]
P[k1] [k]

(37)
(38)

The initial guess for the adaptation gain matrix P and the
weighting factor (0 1) must be specified by the user.
Once [k] has been calculated (and subsequently A, B, C), the
desired parameter estimates at the kth sample period can be
recovered using

2
1 T
T
Me =
(A + C B) , Be =
(A C) ,
4 2
4
Ke =

1
(A + B + C) .
4

(39)

Similarly to the indirect adaptive scheme, the least-squares


algorithm requires measurements of contact force Fe and the
corresponding end-effector position x. It is noted that unlike
the other two methods considered here, the RLS solution is
geared to identifying impedance parameters within the framework of pure impedance control, without tracking a desired
force signal. Formulated as a linear identification problem, the
success of the parameter estimation is subject to the standard
persistent excitation condition as stated in eq. (27).

3. Description of Benchmark Test


3.1. Physical Setup
To compare the four schemes presented above, a benchmark
test was developed. This test was analyzed using both a numerical simulation in MATLAB, as well as experiments conducted with the Planar Robotics Testbed at the University of
Victoria. The testbed includes a custom manipulator assembled using modular links and joint motors, as well as a flexible
wall fixture (see Figure 3).

Fig. 4. Detail of flexible wall fixture.

The flexible wall represents a relatively simple environment designed specifically for our experiments as a first step
to evaluating the performance of different techniques. In particular, the flexible wall is attached to a set of linear bearings, and can thus translate in one direction. Behind the wall
are a set of coil springs and oil-filled dampers. These components are also modular, and can be added or removed to
change the mechanical impedance of the wall (see Figure 4).
For the benchmark test presented here, the stiffness of the
environment was defined by the two springs in parallel, with
the effective stiffness of Ke = 4800 N m1 . This represents
a relatively soft environmentthe choice made partially because of the limited accuracy of the wall displacement sensor.
In practical applications, we encounter much stiffer environments which makes the identification task more challenging.
The damping coefficients of the dampers are not accurately
known but are estimated to give approximate environment
damping of Be = 200 kg s1 .

48

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003


(Ke , Be ), only one non-zero frequency is required (Landau
et al. 1998). Therefore, a persistently exciting input to the
MRAC and indirect adaptive schemes was defined by adding
a small-amplitude sinusoid to Fd to give

0.7
Reference trajectory
Flexible wall

0.6

0.5

Fr = Fd + 8 sin(20t).

(40)

xr (m)

0.4

For the RLS scheme, the persistent excitation must be added


directly to the desired trajectory, giving

0.3

xr = xd + 0.02 sin(20t),

0.2

0.1

10

15

time (s)

Fig. 5. Trajectory of manipulator, showing contact with


flexible wall.

Instrumentation on the testbed includes: joint encoders on


the manipulator, a planar force/torque sensor mounted at the
end-effector and a linear displacement potentiometer directly
connected to the flexible wall. The force sensor measures
the contact force at the robot end-effector. The potentiometer
gives a measurement of the wall deflection from which the
end-effector velocity is calculated using a finite difference
scheme (for the indirect adaptive and MRAC methods). We
note that, although end-effector force sensors are common
on robot manipulators, an absolute position sensor capable
of measuring the environment deflection in three dimensions
may be difficult to implement. This is especially true for stiff
environments.
3.2. Trajectory and Inputs
The manipulator was commanded to follow a reference trajectory using impedance control. The reference trajectory used
in numerical simulations is shown in Figure 5 and was defined
by a fifth-order polynomial for the approach and withdrawal
phases of the maneuver. Part of the trajectory lies inside the
flexible wall, resulting in robotenvironment contact for several seconds. Note that during the contact phase, a desired
force (Fd = 50 N) was specified for the indirect adaptive and
MRAC controllers, with the reference trajectory modified accordingly (see eq. (20) for the indirect adaptive controller).
As discussed earlier, accurate damping estimation can be
guaranteed only with a persistently exciting reference signal.
A common approach to create such a signal is by adding several sinusoidal signals with distinct frequencies. It has been
shown that the number of required frequencies in the reference signal depends on the number of parameters to be determined. For a system involving two unknown parameters

(41)

where xd is the desired (reference) trajectory before the addition of persistent excitation. The amplitudes of these signals were chosen to give a good signal-to-noise ratio, while
remaining within the physical capabilities of the manipulator and environment. The frequency of the sinusoid is sufficiently far from zero while still within the bandwidth of the
manipulator.
When the choice of the excitation frequency is not evident or to increase the temporal bandwidth of the excitation,
a pseudo-random binary sequence (PRBS) can be used. This
signal resembles a square wave with a constant amplitude, but
has a randomly varying period (Landau et al. 1998). As such,
it has a constant spectral density over a broad bandwidth of
frequencies. Experimental results obtained with PRBS excitation are presented in Section 5.
The target impedance coefficients used in the indirect adaptive and RLS algorithms were chosen to be: Mt = 5 kg,
Bt = 100 kg s1 and Kt = 500 N m1 . These nominally
cause the manipulator-environment system to have a natural frequency of n = 10 rad s1 and critical damping. The
MRAC algorithm was found to be unstable during contact
when these target impedance values were used, so this algorithm was individually tuned, resulting in: Mt = 1 kg,
Bt = 650 kg s1 and Kt = 1e5 N m1 . Finally, to produce an
underdamped response for the signal processing method, we
used: Mt = 20 kg, Bt = 50 kg s1 and Kt = 2000 N m1 .
Each algorithm was given the same initial conditions for
the parameter estimates: K e[0] = 3000 N m1 and B e[0] =
50 kg s1 . Other initial guesses were also attempted with, overall, not significantly different outcomes. Note that, in practice,
the initial guess for environmental stiffness should error on
the soft side in order to avoid excessive transients in the contact forces. The adaptation gain matrix used in the indirect
and MRAC schemes was chosen through trial and error to
be  = diag{5.0e4, 2.5e3}. The RLS scheme required an
initial value for P[0] = In , where is a large constant (Landau et al. 1998). For our simulations and experiments we set
P[0] = 105 In . The weighting factor (), which is often chosen to be slightly less than one, was set to = 0.98 (Landau
et al. 1998). Using these values, simulation and experimental results of the benchmark test were obtained both with and
without persistent excitation, as presented below.

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation

49

4. Simulation Results

4.2. Results Without Persistent Excitation

The benchmark test was first implemented as a numerical


simulation using MATLAB. To model the noise inherent in
the force sensor used in our experiments, a random signal with
an amplitude of 1 N was added to the calculated contact
force. On the other hand, perfect end-effector position and
velocity measurements were assumed in the simulations.

For practical applications, persistent excitation may prove to


be difficult or hazardous. For stiff environments, the amplitude
of possible motions may be too small to have a positive effect
on convergence, while for fragile environments the time varying force signal may cause damage. Therefore, it was deemed
important to study how the contact parameter estimation algorithms perform in the absence of persistent excitation. The
same benchmark test was performed, but the superimposed
sinusoidal signal was removed from Fr or xr in eqs. (??) and
(??).
The measured force response for the signal processing
method during contact is given in Figure 7. The contact force
demonstrates the desired second-order underdamped behavior, as shown. For this simulation, the response converges to
within 5% of the steady-state value in approximately 1.5 cycles, which results in a calculated damping ratio of = 0.303.
The frequency response of this signal shows a single dominant
frequency at d = 2.93 Hz (see Figure 8) which corresponds
to the undamped natural frequency n = 3.07 Hz. Following
the method outlined in Section 2, the stiffness and damping
estimates were found to be 5441 N m1 and 173 kg s1 , resulting in a parameter error of 13% for both values.
Results from the indirect adaptive, MRAC and RLS methods are given in Figure 9, and demonstrate some interesting
trends. The stiffness estimates still converged to nearly the
correct value, with errors for the indirect, MRAC and RLS
schemes equal to 2.4%, 0.1% and 1.7% respectively. As before, the RLS algorithm demonstrated significant overshoot,
but converged rapidly to nearly the actual value. The response
of the indirect and MRAC controllers was smooth, rapid and
accurate. Thus, the absence of persistent excitation does not
significantly affect the stiffness estimation when applied to
simulated results.
Substantial degradation, however, is apparent in the damping estimates. As can be seen from the damping (lower) plot in
Figure 9, the RLS estimate converged slowly, and exhibited a
large amount of noise in the response. The contact phase was
sufficiently long however (3.3 s) to allow for estimation of
damping to occur with the final estimate error of 2.5%. Thus,
the RLS algorithm demonstrates damping convergence even
in the absence of the sinusoidal persistent excitation signal
because the trajectory is slowly time-varying. For the indirect
adaptive and MRAC schemes, it is evident that adaptation occurs exclusively during the transient phases of the maneuver:
the onset and conclusion of contact. Due to the force-tracking
nature of these algorithms, it is only during these times that
the manipulator has a non-zero velocity normal to the wall.
As evidenced by the accurate estimates of Be calculated with
the indirect and MRAC algorithms (3.0% and 0.0% errors,
respectively), adaptation during these brief moments is possible. However, the convergence rate is highly dependent on
the adaptation gains, and thus convergence to actual values

4.1. Results With Persistent Excitation


Sinusoidal excitation was added to the reference signal of the
indirect adaptive, MRAC and RLS algorithms (recall that the
signal processing method uses a trajectory with a step input, so
this algorithm should not be used with persistent excitation).
For these three methods, the prescribed trajectory caused the
manipulator to be in contact with the wall for just over 3 s, from
t = 5.9 s to t = 9.2 s. During this time, all three algorithms
showed a definite parameter adaptation of the stiffness and
damping estimates towards the actual values (see Figure 6).
The stiffness plot showed that the indirect adaptive controller
and the MRAC controller resulted in a very similar response.
In particular, both methods converged within 1 s and produced
stiffness estimates with less than 0.1% error (as defined by
the value at the last instant of contact). Although the RLS
algorithm demonstrated significant overshoot, convergence
was also rapid and accurate, with a final stiffness estimate
error of 1.3%.
The plot of damping estimation demonstrates good convergence for each algorithm (see the lower plot in Figure 6). The
response was, in general, rapid with little overshoot. In addition, all algorithms resulted in damping parameter estimates
with less than 1.0% error. Through many simulations, it was
apparent that the damping estimation was dramatically improved if the persistent excitation signal was increased (either
in amplitude or frequency). This is reasonable since either
change would result in faster manipulator reference velocities (xr ) and therefore more rapid convergence to the actual
parameter value. Based on these simulations, it can be concluded that in the presence of persistent excitation all three
algorithms are capable of very accurate stiffness and damping
estimation.
It is noted that the rate of parameter convergence is highly
dependent on the choice of adaptation gains. These gains are
chosen by the user (usually by trial and error) which implies
that the parameter convergence rate can be controlled to some
extent. The indirect adaptive controller and MRAC controller
use identical parameter adaptation laws and gain matrices,
allowing for direct comparisons of these algorithms. The RLS
scheme, however, is fundamentally different in the manner in
which it forms parameter estimates and therefore cannot be
directly compared to the other schemes.

50

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003


10000

Stiffness (N/m)

8000
6000

Actual

4000
Indirect
MRAC
RLS

2000
0

10

15

time (s)

400

Damping (kg/s)

300
Actual

200
100

Indirect
MRAC
RLS

0
100

10

15

time (s)

Fig. 6. Simulation results of stiffness and damping estimation with sinusoidal excitation.

45

1600

Calculated force
5% bounds

40

1400

30

1200

25

1000

Magnitude

Force (N)

35

20
15

800
600

10

400

200

0
5
10.5

11

11.5

12

time (s)

Fig. 7. Simulated force response during contact for the signal


processing method.

10

20
30
Frequency (Hz)

40

Fig. 8. Frequency response of contact force signal.

50

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation

51

10000

Stiffness (N/m)

8000
6000

Actual

4000
Indirect
MRAC
RLS

2000
0

10

15

time (s)

400

Damping (kg/s)

300
Actual

200
100

Indirect
MRAC
RLS

0
100

10

15

time (s)

Fig. 9. Simulation results of stiffness and damping estimation without persistent excitation.

cannot be guaranteed before the velocity of the contact point


approached zero.

5. Experimental Results
The simulation results presented above indicate that all four
algorithms are capable of accurate environment parameter
estimation. Using the signal processing method, acceptable
stiffness and damping estimation were possible without persistent excitation. For the indirect adaptive, MRAC and RLS
algorithms, stiffness estimation is very accurate either in
the presence or absence of persistent excitation, but reliable
damping estimation cannot be guaranteed without additional
excitation.
Experimental validation of the indirect adaptive, RLS and
signal processing algorithms was performed by implementing
the benchmark test on the Planar Robotics Testbed described
in Section 3. As demonstrated in simulations of Section 4, the
MRAC algorithm generated very similar parameter estimates
and responses to those of indirect adaptive method since the
two share identical parameter adaptation laws and gain matrices. At the same time, the indirect adaptive controller was
found to be intuitive and easily implemented, while the MRAC

method was considerably more complex. For these reasons,


we concluded that the MRAC algorithm offers no significant
advantage over the indirect adaptive controller and therefore
it was not investigated with the Robotics Testbed.
5.1. Data Collection and Processing
All experiments were performed with a sampling rate of
500 Hz. The measured data were filtered by using the Butterworth filter of second order with cut-off frequency set to
50 Hz. These filter parameters were optimized manually to
minimize the errors in stiffness and damping estimates. The
speed of the wall was calculated by using the finite-difference
scheme on the wall deflection and filtering the result. Finally,
some static friction is present at the bearings of the wall fixture and it was determined that a force of 6 N is required to
move the wall from its stationary position. This value was
subtracted from the force measurements to give the contact
force employed in the estimation algorithms.
5.2. Results With Persistent Excitation
Indirect adaptive and RLS parameter estimation results are
presented for the sinusoidal and PRBS excitations in the input signal. In each case, 10 experiments were performed to

52

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003


0.7

0.6

0.5

xr (m)

0.4

0.3

IA trajectory

0.2

0.1
Wall
0

0.1

RLS trajectory

10

15

20

time (s)

Fig. 10. Reference trajectories used for the indirect adaptive and RLS experiments with sinusoidal excitation.

arrive at the average parameter estimates reported below. The


contact force and wall deflection profiles measured in one
of the tests with sinusoidal excitation are shown in Figure 11.
The reference trajectories used for the RLS and indirect adaptive experiments with sinusoidal excitation are plotted in Figure 10. These were designed to produce comparable deflections and forces for a more truthful evaluation of the RLS and
indirect adaptive estimation results. The corresponding profiles of the stiffness and damping estimation are illustrated in
Figure 12.
Stiffness estimation for both algorithms is fairly good, with
errors of 12.6% for the indirect adaptive controller and 3.5%
for the RLS technique. (Note that, as in simulation results,
these errors are calculated for the estimates at the last instant
of contact.) Convergence was rapid in each case (510 periods
of oscillation), with the indirect adaptive controller exhibiting a smooth response while the RLS scheme again showed a
large initial overshoot. The RLS estimate also trailed during
the withdrawal from the wall. These differences in performance concur with the differences in trajectories for the two
maneuvers as can be seen in Figure 10.
Damping estimation does not exhibit clear convergence,
although the end-of-contact values are reasonable for both
methods. The average errors for damping are 15.7% and
22.0% for the indirect adaptive and RLS controllers, respectively. To verify convergence of damping estimation, experiments were conducted with a longer contact maneuver. The

corresponding results are shown in Figures 13 and 14. These


reconfirm convergence of the stiffness estimates observed before, and demonstrate convergence of damping albeit to a
value higher than expected.
A set of results obtained by using PRBS excitation described in Section 3.2 is included in Figure 15. The excitation signal was generated by using 10 binary registers which
were shifted at a frequency of 20 Hz. The feedback signal
was created from a combination of the 7th and 10th register values (as suggested in Landau et al. 1998). The resulting
signal, composed of zeros and ones, was scaled and shifted
to produce an excitation signals of 8 N and 0.02 m for
the force and motion inputs, respectively. The results in Figure 15 exhibit similar characteristics to those obtained with
the sinusoidal excitation (Figure 12), but with a noisier convergence. Corresponding results for the longer maneuver are
shown in Figure 16. These exhibit similar features to PRBS
results in Figure 15, but with a more definitive convergence
of estimation.
5.3. Results Without Persistent Excitation
Experiments without the addition of an excitation signal were
performed with the signal processing, indirect adaptive and
RLS methods. As before, 10 experiments were carried out for
each procedure to generate an average stiffness and damping estimate. The values obtained with the signal processing
method were 3029 N m1 and 35 kg s1 for stiffness and

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation

20
0
Force (N)

20
Indirect
RLS

40
60
80
100

10
time (s)

12

14

16

18

20

10
time (s)

12

14

16

18

20

x 10

Displacement (m)

0
2
Indirect
RLS

4
6
8
10

Fig. 11. Experimental profiles of force and wall deflection with sinusoidal excitation.

8000
Indirect
RLS

Stiffness (N/m)

7000
6000

Actual

5000
4000
3000
2000

10
time (s)

12

14

16

18

20

10
time (s)

12

14

16

18

20

400

Damping (kg/s)

Indirect
RLS
300
Actual

200

100

Fig. 12. Experimental results of stiffness and damping estimation with sinusoidal excitation.

53

54

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003

20

Force (N)

0
20
40
60
80

Indirect
RLS

100

x 10

10

time (s)

15

20

25

Displacement (m)

0
2
4
6
8

10

Indirect
RLS
0

10

time (s)

15

20

25

Fig. 13. Experimental profiles for long contact maneuver of force and wall deflection with sinusoidal excitation.

10000
Indirect
RLS
Stiffness (N/m)

8000

6000
Actual
4000

2000

10

time (s)

15

20

25

Damping (kg/s)

400

300
Actual

200

100
Indirect
RLS
0
0

10

time (s)

15

20

25

Fig. 14. Experimental results for long contact maneuver of stiffness and damping estimation with sinusoidal excitation.

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation

8000
Indirect
RLS

Stiffness (N/m)

7000
6000

Actual

5000
4000
3000
2000

10
time (s)

12

14

16

18

20

10
time (s)

12

14

16

18

20

400

Damping (kg/s)

Indirect
RLS
300
Actual

200

100

Fig. 15. Experimental results of stiffness and damping estimation with PRBS excitation.

10000

Stiffness (N/m)

Indirect
RLS
8000

6000
Actual
4000

2000

10

time (s)

15

20

25

400
Damping (kg/s)

Indirect
RLS
300
Actual

200

100

10

time (s)

15

20

25

Fig. 16. Experimental results for long contact maneuver of stiffness and damping estimation with PRBS excitation.

55

56

THE INTERNATIONAL JOURNAL OF ROBOTICS RESEARCH / January 2003

9000
Indirect
RLS

Stiffness (N/m)

8000
7000
6000

Actual

5000
4000
3000
2000

10
time (s)

12

14

16

18

20

10
time (s)

12

14

16

18

20

400

Damping (kg/s)

Indirect
RLS
300
Actual

200

100

Fig. 17. Experimental results of stiffness and damping estimation without persistent excitation.

damping, resulting in errors of 37% and 82%, respectively


(Erickson 2000). The substantial discrepancies from the actual parameter values can be explained by the fact that the
impedance of the arm is not exactly represented by the target
impedance. Also, finite sampled data and inaccuracies inherent in the estimation of the damping ratio limit the accuracy
of the signal processing technique.
Stiffness and damping estimation for the indirect adaptive
and RLS methods without persistent excitation is shown in
Figure 17. The indirect adaptive controller resulted in stiffness estimates with an average error of 8.9% while the RLS
algorithm was more substantially amiss and demonstrated an
average error of 19.2%. The cause for the discrepancy between the converged stiffness of the indirect adaptive method
and the steady-state value predicted during mid-contact of
the RLS maneuver has not been uncovered.
The results of the damping estimation were in line with our
expectations. In the absence of persistent excitation, accurate
damping estimation should not be possible with either algorithm. This was in fact observed with the indirect adaptive and
RLS methods which resulted in damping errors of 60% and
85%, respectively. However, an attempt at adaptation is visible during the initial and, in the case of RLS, final transients,
when the wall velocity is non-zero. Since the adaptation rate is
gain-dependent, it cannot be concluded that the actual damping value will always be obtained during this limited time.

6. Conclusions
The analysis and results presented in this paper are intended to
compare contact parameter identification algorithms. Specifically, a signal processing method, indirect adaptive controller,
MRAC controller and RLS estimator were used to determine
the stiffness and damping of the environment during robot
constrained motion. The signal processing method represents
an original contribution, while the remaining three schemes
have been proposed by other authors (Seraji and Colbaugh
1997; Singh and Popa 1995; Love and Book 1995). However,
substantial modifications were made to the original indirect
adaptive controller, and a proof of parameter convergence was
given.
The signal processing method uses frequency-domain and
time-domain information, and accordingly was implemented
off-line. It has one substantial advantage: only force measurements are required to extract the desired information. The
other three methods are time-domain algorithms and were implemented on-line. Their data requirements include the force,
deflection and, with the exception of RLS, velocity at the contact point. Accurate deflection or position measurements are
difficult in practice, but many identification methods in the
literature require it. In our future work, we will address this
issue by including environment location in the parameters to
be estimated.

Erickson, Weber, and Sharf / Contact Stiffness and Damping Estimation


The simulation results obtained with the four algorithms
indicate that all can generate accurate estimates of both Ke
and Be in a short time. With the exception of the signal processing method, persistent excitation was found to be required
for accurate estimation of both stiffness and damping. Without persistently exciting signals, stiffness estimation was still
possible (in fact, it was more accurate), but damping estimation was not reliable. Three of the algorithms considered were
tested experimentally by using a 3-DoF robotic arm contacting a flexible wall. The results of these tests indicate that the
indirect adaptive and RLS methods are capable of generating
very good estimates for the wall stiffness and damping (in the
presence of persistent excitation), within several periods of
oscillation. Without persistent excitation, some damping estimation may occur during contact transients, but this depends
on the maneuver and gain selection. The indirect adaptive solutions tended to be smoother, showing a more definite convergence. This, however, is likely to be due to the force-tracking
feature of the controller, rather than an inherent property of
the estimator. We also observe that the indirect adaptive controller was found to be intuitive and easily implemented, more
so than the RLS and MRAC schemes.
Estimates calculated with the signal processing method
were less accurate because of the discrepancy between real
and target impedances of the manipulator and limited sampled
data. Even then, we believe the method may still provide a viable option for practical applications, if the real impedance
of the arm can be modeled. Our future work in the area of
contact parameter identification will involve investigation of
frequency-domain methods and evaluation of these and timedomain techniques for contacts involving complex geometries
and stiff environments.

References
An, C., Atkinson, C., and Hollerbach, J. 1988. Model-Based
Control of a Robot Manipulator. Cambridge, MA: MIT
Press.
Anderson, B., Bitmead, R., Johnson Jr, C., Kokotovic, P., Kosut, R., Mareels, M., Praly, L., and Riedle, B. 1986. Stability of Adaptive Systems: Passivity and Averaging Analysis.
Cambridge, MA: MIT Press.

57

Dorf, R., and Bishop, R. 1995. Modern Control Systems. 7th


edition. Reading, MA: Addison-Wesley.
Erickson, D. 2000. Contact stiffness and damping estimation
for constrained robotics systems. Masters thesis, University of Victoria, Victoria, Canada.
Gilardi, G., and Sharf, I. 2002. Literature survey of contact
dynamics modeling. Journal of Mechanism and Machine
Theory, 37:12131239.
Hogan, N. 1985. Impedance control: An approach to manipulation: Parts i-iii. ASME Journal of Dynamic Systems,
Measurement and Control, 107:124.
Landau, I., Lozano, R., and MSaad, M. 1998. Adaptive Control. London: Springer-Verlag.
Ljung, L. 1987. System Identification: Theory for the User.
Englewood Cliffs, NJ: Prentice-Hall.
Love, L., and Book, W. 1995. Environment estimation for enhanced impedance control. In Proceedings of the IEEE
International Conference on Robotics and Automation,
pp. 18541859.
Ma, O., Buhariwala, K., Roger, N., MacLean, J., and Carr,
R. 1997. MDSFa generic development and simulation
facility for flexible, complex robotic systems. Robotica,
15:4962.
Seraji, H., and Colbaugh, R. 1997. Force tracking in
impedance control. International Journal of Robotics Research, 16(1):97117.
Singh, S., and Popa, D. 1995. An analysis of some fundamental problems in adaptive control of force and impedance
behavior: Theory and experiments. IEEE Transactions on
Robotics and Automation, 11(6):912921.
Slotine, J.-J., and Li, W. 1991. Applied Nonlinear Control.
Englewood Cliffs, NJ: Prentice-Hall.
Van Vliet, J., Sharf, I., and Ma, O. 2000. Experimental validation of contact dynamics simulation of constrained
robotic tasks. International Journal of Robotics Research,
19(12):12031217.
Weber, M., Ma, O., and Sharf, I. 2002. Identification of contact dynamics model parameters from constrained robotic
operations. Presented at DETC02 ASME 2002 Design Engineering Technical Conferences and Computers and Information in Engineering Conference, Montreal, Canada,
September.

You might also like