An experimental robot load identification method for industrial application

Similar documents
Combining Internal and External Robot Models to Improve Model Parameter Estimation*

Research Article Dynamic Model Identification for 6-DOF Industrial Robots

Selection of Servomotors and Reducer Units for a 2 DoF PKM

Manipulator Dynamics 2. Instructor: Jacob Rosen Advanced Robotic - MAE 263D - Department of Mechanical & Aerospace Engineering - UCLA

In this section of notes, we look at the calculation of forces and torques for a manipulator in two settings:

Parameter identification of robotic systems with series elastic actuators

Identification of Industrial Robot Parameters for Advanced Model-Based Controllers Design

Optimal Excitation and Identification of the Dynamic Model of Robotic Systems with Compliant Actuators

Robotics I. Test November 29, 2013

Dynamic model of robots:

Identification of the payload inertial parameters of industrial manipulators

Artificial Intelligence & Neuro Cognitive Systems Fakultät für Informatik. Robot Dynamics. Dr.-Ing. John Nassour J.

Adaptive Robust Tracking Control of Robot Manipulators in the Task-space under Uncertainties

q 1 F m d p q 2 Figure 1: An automated crane with the relevant kinematic and dynamic definitions.

Identifying the Dynamic Model Used by the KUKA LWR: A Reverse Engineering Approach

Virtual Passive Controller for Robot Systems Using Joint Torque Sensors

Dynamic model of robots:

Robotics I. Classroom Test November 21, 2014

Robot Dynamics II: Trajectories & Motion

Dynamics. 1 Copyright c 2015 Roderic Grupen

Balancing of an Inverted Pendulum with a SCARA Robot

Experimental Joint Stiffness Identification Depending on Measurements Availability

Robust Control of Cooperative Underactuated Manipulators

Lecture Schedule Week Date Lecture (M: 2:05p-3:50, 50-N202)

IROS 16 Workshop: The Mechatronics behind Force/Torque Controlled Robot Actuation Secrets & Challenges

Lecture «Robot Dynamics»: Dynamics 2

Positioning Controller for Mechanical Systems with a Mini Harmonic Drive Servo Actuator

(W: 12:05-1:50, 50-N202)

DYNAMIC MODELLING AND IDENTIFICATION OF A CAR. Gentiane Venture* Wisama Khalil** Maxime Gautier** Philippe Bodson*

Rotational & Rigid-Body Mechanics. Lectures 3+4

Lecture «Robot Dynamics»: Dynamics and Control

Robot Control Basics CS 685

A FORCE BALANCE TECHNIQUE FOR MEASUREMENT OF YOUNG'S MODULUS. 1 Introduction

Chapter 6: Momentum Analysis

Game Physics. Game and Media Technology Master Program - Utrecht University. Dr. Nicolas Pronost

Automatic Control Motion planning

Lesson Rigid Body Dynamics

Natural Frequency Analysis of Spring-Manipulator System for Force Generation Utilizing Mechanical Resonance

Exponential Controller for Robot Manipulators

Robotics I. Figure 1: Initial placement of a rigid thin rod of length L in an absolute reference frame.

Introduction to Robotics

Gordon G. Parker Mechanical Engineering Dept., Michigan Institute of Technology Houghton, MI

Nonlinear Identification of Backlash in Robot Transmissions

A Benchmark Problem for Robust Control of a Multivariable Nonlinear Flexible Manipulator

Dynamics. describe the relationship between the joint actuator torques and the motion of the structure important role for

A MOTORIZED GRAVITY COMPENSATION MECHANISM USED FOR THE NECK OF A SOCIAL ROBOT

A Physically-Based Fault Detection and Isolation Method and Its Uses in Robot Manipulators

LUMPED MASS MODELLING WITH APPLICATIONS

Experimental identification of the static model of an industrial robot for the improvement of friction stir welding operations

Trajectory Planning from Multibody System Dynamics

Funnel control in mechatronics: An overview

Research Article On the Dynamics of the Furuta Pendulum

GAIN SCHEDULING CONTROL WITH MULTI-LOOP PID FOR 2- DOF ARM ROBOT TRAJECTORY CONTROL

Case Study: The Pelican Prototype Robot

Manufacturing Equipment Control

Design and Control of Compliant Humanoids. Alin Albu-Schäffer. DLR German Aerospace Center Institute of Robotics and Mechatronics

Robotics I Midterm classroom test November 24, 2017

DYNAMICS OF SERIAL ROBOTIC MANIPULATORS

Differential Kinematics

NONLINEAR FRICTION ESTIMATION FOR DIGITAL CONTROL OF DIRECT-DRIVE MANIPULATORS

Multibody simulation

THE ESTIMATION OF MASS MOMENT OF INERTIA ON AN EQUIPMENT - MODELLING AND OPTIMISATION

1820. Selection of torsional vibration damper based on the results of simulation

ENGG 5402 Course Project: Simulation of PUMA 560 Manipulator

FREEFLYING ROBOTS - INERTIAL PARAMETERS IDENTIFICATION AND CONTROL STRATEGIES

D DAVID PUBLISHING. Design of Torque Balancing Mechanisms. 1. Introduction. Bruno Zappa, Vittorio Lorenzi, Paolo Righettini and Roberto Strada

Introduction to Haptic Systems

NMT EE 589 & UNM ME 482/582 ROBOT ENGINEERING. Dr. Stephen Bruder NMT EE 589 & UNM ME 482/582

Nonlinear PD Controllers with Gravity Compensation for Robot Manipulators

MODELING AND IDENTIFICATION OF A MECHANICAL INDUSTRIAL MANIPULATOR 1

6. 3D Kinematics DE2-EA 2.1: M4DE. Dr Connor Myant

Dynamic Features of a Planetary Speed Increaser Usable in Small Hydropower Plants

Design and Control of Variable Stiffness Actuation Systems

Introduction to Control (034040) lecture no. 2

Laws of gyroscopes / cardanic gyroscope

Seul Jung, T. C. Hsia and R. G. Bonitz y. Robotics Research Laboratory. University of California, Davis. Davis, CA 95616

Mechatronic System Case Study: Rotary Inverted Pendulum Dynamic System Investigation

Controlling the Apparent Inertia of Passive Human- Interactive Robots

Robust linear control of an active suspension on a quarter car test-rig

ROBUST CONTROL OF A FLEXIBLE MANIPULATOR ARM: A BENCHMARK PROBLEM. Stig Moberg Jonas Öhr

A Sliding Mode Controller Using Neural Networks for Robot Manipulator

Robotics I. February 6, 2014

Modelling of Ball and Plate System Based on First Principle Model and Optimal Control

S 5 S 3 S 4 S 1 S 2 P 6 P 5 P 3 TCP

Video 8.1 Vijay Kumar. Property of University of Pennsylvania, Vijay Kumar

MODELING AND EXPERIMENTAL DESIGN FOR THE ON-ORBIT INERTIAL PARAMETER IDENTIFICATION OF FREE-FLYING SPACE ROBOTS

Design of Extended Kalman Filters for High Performance Position Control of Electrical Drives

Chapter 6: Momentum Analysis of Flow Systems

Robot Manipulator Control. Hesheng Wang Dept. of Automation

Dynamic Tests on Ring Shear Apparatus

Mechanical Engineering Department - University of São Paulo at São Carlos, São Carlos, SP, , Brazil

Dynamics modeling of an electro-hydraulically actuated system

A Search for Consensus Among Model Parameters Reported for the PUMA 560 Robot

Dynamic Modeling of Rotary Double Inverted Pendulum Using Classical Mechanics

Impedance control using a cascaded loop force control

Translational and Rotational Dynamics!

Lecture 14: Kinesthetic haptic devices: Higher degrees of freedom

Structural Dynamics Lecture 2. Outline of Lecture 2. Single-Degree-of-Freedom Systems (cont.)

Servo Motors Classification Based on the Accelerating Factor

Toward Torque Control of a KUKA LBR IIWA for Physical Human-Robot Interaction

Transcription:

An experimental robot load identification method for industrial application Jan Swevers 1, Birgit Naumer 2, Stefan Pieters 2, Erika Biber 2, Walter Verdonck 1, and Joris De Schutter 1 1 Katholieke Universiteit Leuven, Division PMA, Celestijnenlaan 300B, B-3001 Heverlee, Belgium, jan.swevers@mech.kuleuven.ac.be 2 AMATEC Robotics GmbH, Landsberger Strasse 63a, D-82110 Germering, Germany Abstract. This paper discusses a new experimental robot load identification method that is used in industry. The method is based on periodic robot excitation and the maximum likelihood estimation of the parameters, techniques adopted from [1]. This method provides (1) accurate estimates of the robot load inertial parameters, and (2) accurate actuator torques predictions, both of which are essential for the acceptance of the results in an industrial environment. The key element to the success of this method is the comprehensiveness of the applied model, which includes beside the dynamics resulting from the robot load and motor inertia, the coupling between the actuator torques, the mechanical losses in the motors and the efficiency of the transmissions. Experimental results on a KUKA industrial robot equipped with a calibrated test load are presented. 1 Introduction The ever increasing market demands push manufacturers to use their robots at the limits of performance, and often beyond these limits. Approximately 50% of the robots are statically or dynamically overloaded 1. This occurs especially in spot welding applications and handling tasks. The overloading often causes premature failure of the robot. The reason for this overloading is that the path planning for these robots is often based on a dynamic robot model that includes inaccurate estimates of the inertial parameters of the load. Tool manufacturers often cannot provide these estimates at all or at least not with sufficient accuracy. Reducing the cycle time by optimizing the robot trajectory is an important issue. This optimization has to be based on a model of the robot dynamics, taking into account accurate estimates of the inertial parameters of the payload it is carrying (e.g. a spot welding tool). The most important aspect of these models is their ability to accurately predict the required actuator torques based on the desired robot motion. Experimental robot identification techniques are a time and effort efficient way to obtain accurate estimates of the inertial parameters of the robot load. 1 data from AMATEC Robotics GmbH, member of the KUKA Group

2 Jan Swevers et al. They estimate the parameters of dynamic robot models based on motion and actuator torque data measured during well-designed identification experiments, i. e. robot motions along optimized trajectories, e.g. [2 5,1]. The actuator torque data are obtained through actuator current measurements, i.e. no additional sensors are required. This paper discusses an experimental robot load identification approach that borrows its experiment design and maximum likelihood parameter estimation approach from [1]. This load identification approach is already recognized by industry as an effective means of robot load identification that is frequently used [6]. For it to have accomplished this result, the method has to satisfy the following two conditions: (1) provide a model that allows accurate prediction of the actuator torques, and (2) provide accurate estimates of the inertial parameters of the load. It will be shown that the first objective is more easy to satisfy than the second, i.e. a model which provides acceptable actuator torques predictions does not necessarily provide accurate parameter estimates. This last issue is important because the obtained model will only be accepted in an industrial environment if the values of the estimated parameters correspond to expected values of the inertial parameters 2. The key element to obtain sufficiently accurate parameter estimates is the comprehensiveness of the applied dynamic model, i.e. which dynamic effects and losses are considered or neglected. The outline of the paper is as follows. Section 2 describes the model used for robot load identification, and discusses how the above mentioned dynamic effects related to the motors and transmissions are taken into account for the considered test case. Section 3 briefly summarizes the experiment design and parameter estimation approach, which are based on [1]. Section 4 describes the test case, discusses and evaluates the experimental results. 2 Generation of a dynamic robot model for load identification The dynamic model of a n DOF robot manipulator derived according to the Newton-Euler formalism [7], is a dynamic relation between the gear torques τ G, i.e. the torques that drive the links, the motion of the links, i.e. the joint positions q G, velocities q G, and accelerations q G, and the 10 inertial parameters of each link, represented by vector p (mass, coordinates of the center of gravity (COG), and the inertia tensor of each link related to its COG): τ G = f(q G, q G, q G, p). (1) This model is nonlinear in the parameters p. All variables that appear in this text are expressed at the link side. 2 In fact, this experience was the main motivation for AMATEC Robotics GmbH to start with the development of the presented identification method

Experimental robot load identification 3 Since the robot load is connected to the last robot link (link n), the load is considered as part of this link. Typically for a KUKA robot used for spot welding operations, the mass of link n = 6 itself is 3 to 10 kg, compared to the mass of the load it can carry: 100 to 350 kg. In general, the mass of the last link itself is smaller than the accuracy that can be obtained, which corresponds to a maximum relative error of 5%. 2.1 Additional effects Model (1) does not consider any losses in the motors and the transmissions, nor coupling between the actuator torques. These effects are described in this section. Coupling between the actuator torques The axes of the wrist of the considered robot are mechanically coupled in the sense that the motion of each wrist link is determined by more than one wrist actuator. This coupling is described as a fixed linear transformation, which is known with high accuracy, between the joint and actuator space. Actuator dynamics A considerable part of the actuator torques is consumed by actuator friction and by the accelerating and decelerating rotor inertia. Rotor inertia and friction are included in the robot model. The friction model includes viscous and Coulomb friction. Efficiency of the transmission The mechanical transmission between the actuator and link is a complex system, with many parts moving at different angular velocities. This makes it difficult to model the losses in the transmission by means of a friction model. A more appropriate approach is to consider the efficiency of the transmission, which depends in a nonlinear way on the joint velocity q Gi and torque τ Gi and is modelled using only one parameter R zi per joint : τ Ti = g(τ Gi, q Gi, R zi ). τ Ti and τ Gi are the torques at the input and output of the transmission of joint i, respectively. Relation between motor current and motor torque The KUKA KR150 is equipped with permanent magnet synchronous motors. The weakly nonlinear relation between the torque and current of these motors has been identified experimentally using a third order polynomial model. 2.2 Total robot model The resulting dynamic robot model, which includes all above mentioned effects, can be summarized as follows: τ M = F (q M, q M, q M, θ), (2) F () is the vector function that summarizes equation (1) and the above mentioned effects. It relates the measured variables, i.e. the motor torques (cur-

4 Jan Swevers et al. rents) τ M and motor angles, velocities and accelerations, to the model parameters : θ = [ p R v R c J M R z ] T, (3) with R v, R c, J M, and R z the m-vectors containing for all axes, the motor viscous and Coulomb friction coefficients, the rotor inertia, and the efficiency parameters of the transmission respectively. θ contains 14 m parameters. m is the number of axes that are actuated during the excitation experiment. Model (2) is nonlinear in the parameters θ (3). Due to motion constraints (e.g. each robot rotation axis has only one degree of freedom) and the measurement constraints (only the joint motion and the actuator currents (torques) are measured) not all 10 inertial parameters of each link are identifiable: some are identifiable in combinations and other parameters do not appear in the dynamical equations. The construction of a base set of independent inertial parameters, i.e. a complete and minimal parameterization, can be performed a priori, from the kinematic structure of the robot. It is assumed that this parameter reduction has been performed. The resulting base set of inertial parameters for the considered robot load identification problem is further discussed in section 4. The rotor inertia of all actuators and the inertial parameters of all robot links except link 6, since this link is considered as part of the load, are constant and can be considered identical for all robots of the same type. The ones that influence the robot dynamics can be determined a priori using series of identification experiments, and considered as known for the robot load identification, reducing the number of unknown parameters in θ, and the uncertainty on the estimates of these parameters [8]. The Coulomb and viscous friction parameters, and the parameters that model the efficiency of the transmissions are not constant and therefore considered unknown. They are robot and temperature dependent. 3 Experiment design and parameter estimation issues This section briefly discusses experiment design and parameter estimation, which are based on [1], in view of robot load identification. 3.1 Experiment design The excitation trajectory for each joint is a finite Fourier series. The fundamental pulsation is common for all joints, yielding a periodic robot excitation, which: is advantageous because it allows time-domain data averaging, which yields data reduction and improves the signal to noise ratio of the experimental data. This is especially important for the motor current measurements since they are very noisy.

Experimental robot load identification 5 allows estimation of the variance of the measurement noise by calculating the sample variance [1]. This information is valuable for the maximum likelihood parameter estimation discussed below. allows an analytic calculation of the joint velocities and accelerations from the measured joint angles [1]. The degrees of freedom of this experiment design approach are the length of the Fourier series and the Fourier coefficients [1]. They can be fixed e.g. through optimization, as it is done in [1], or by means of trial-and-error. 3.2 Parameter estimation The robot load identification method discussed in this paper is based on the maximum likelihood estimation method presented in [1]. This method yields the parameter vector θ that maximizes the likelihood of the measurement data. It assumes that the measured actuator torques are corrupted by independent zero-mean Gaussian noise, and that the joint angle measurements, and consequently also the joint velocities and accelerations, are free of noise. Swevers et al. [1] show that under these assumptions, the maximum likelihood parameter estimation corresponds to a weighted least squares optimization problem. The weights are the variance estimates of the actuator torques. Here, the optimization problem is however nonlinear in the parameters because of the nonlinear dependency of the dynamic robot model on the parameters θ (3). It can be solved using iterative search methods such as the Gauss-Newton or Levenberg-Marquardt method. The application of these methods requires an initial estimate of the unknown model parameters. The Jacobian of F w.r.t. the unknown parameters of θ has to be of full rank [8]. 4 Experimental validation This section discusses the implementation and experimental validation of the above mentioned robot load identification procedure on a KUKA KR150 with a calibrated load. Section 4.1 describes the considered test case. Section 4.2 discusses the experiment design and parameter estimation. Section 4.3 summarizes the experimental results. 4.1 Description of the test case The test case consists of a KUKA KR150 industrial robot equipped with a calibrated test load. This robot, which is used in industry mainly for spot welding applications, has 6 rotational joints. Figure 1 shows this robot with a spot-welding tool (left) and the considered calibrated test tool. Table 1, second column, shows the 10 inertial parameters of the test tool. These parameter values result from the CAD-model of the test load and from weighing the different parts of the load.

6 Jan Swevers et al. Fig. 1. Top: KUKA KR150 industrial robot with spot-welding tool. Bottom: Calibrated test load 4.2 Practical implementation of the experiment design and parameter estimation for the KUKA KR150 The unknown parameters for the considered robot load identification are: the 10 inertial parameters of the load, the Coulomb and viscous friction parameters, and the parameters that model the efficiency of the transmissions. The rotor inertias and the inertial parameters of the links that influence the robot dynamics are considered as a priori known (see discussion at the end of section 2). This load identification approach is compared with two other approaches: one approach fixes the friction and efficiency parameters, and one approach uses a barycentric parameterization of the robot dynamics (see section 4.3). Analysis of the rank of the Jacobian of F (2) w.r.t the unknown parameters shows that all inertial parameters can be uniquely identified if axes 3 to 6 are actuated in the excitation experiment. This yields 22 unknown parameters (10 inertial parameters of the load, 2 friction parameters and 1 efficiency parameter for each axis). If only the wrist axes are moved, the number of unknown parameters reduces to 19. However, it is then not possible to identify the mass of the load: the rank of the Jacobian is 18.

Experimental robot load identification 7 The parameters of the Fourier series based excitation trajectories have been selected such that the trajectories for the wrist axes cover the whole working envelope of the robot wrist, taking into account constraints on joint velocities and accelerations. A small movement for axis 3 is sufficient since this motion is only required for the identification of the mass of the load. Figure 2 shows one of the sets of trajectories that has been used for the robot load identification. 1.7 Axis 3 3 Axis 4 1.5 Axis 5 3 Axis 6 1.65 2 1 2 Angle [rad] 1.6 1.55 1 0 1 0.5 0 0.5 1 0 1 1.5 2 1 2 1.45 0 20 40 time [s] 3 0 20 1.5 40 0 20 40 time [s] time [s] 3 0 20 40 time [s] Fig. 2. One of the sets of excitation trajectories for axes 3, 4, 5, and 6 The experiments (periodic excitation and steady state response measurements), and the data processing (averaging, variance estimation, and calculation of joint velocities and accelerations) are performed according to the procedures discussed in [1]. The unknown model parameters are estimated based on the model (2), using the noise-free estimates of joint angles, velocities and accelerations, q Mi (k), q Mi (k), and q Mi (k) respectively, and averaged actuator torque measurements τ Mi (k), for i = 3, 4, 5, 6. This estimation corresponds to the nonlinear least squares optimization problem that considers noise on the actuator data measurements only. This optimization is solved using iterative search methods starting from an initial guess of the unknown parameters. This initial guess is provided by approach 1 discussed in section 4.3. 4.3 Validation of the experimental results Tables 1 and 2 give an overview of the obtained experimental results. The second column of table 1 shows the exact values of the inertial parameters of the calibrated load. Columns 3 to 5 compare the results obtained using 3 alternative approaches: Approach 1: The model used in this approach is based on the barycentric parameters of links 3 to 6 [9], and includes the motor friction coefficients (viscous and Coulomb) and the rotor inertias as unknown. This

8 Jan Swevers et al. model is linear in the parameters, which simplifies their estimation considerably [1]. The efficiency of the transmissions is fixed to 1. The inertial parameters of the test load are then derived from the estimated barycentric parameters and from accurate a priori estimates of inertial parameters of the links. Approach 2: The approach uses model (2). The rotor inertias and inertial parameters of the links are fixed to a priori estimated values. The motor friction parameters are considered unknown. The efficiency of the transmissions is fixed to 1. Approach 3: This approach corresponds to approach 2, except that the efficiency of the transmissions are now considered as unknown. The inertial parameters of the test load are estimated according to these 3 approaches using data resulting from 14 different experiments, yielding 14 sets of estimates for each approach. Columns 3 to 5 of table 1 show the average errors with respect to the exact values (column 2) and the standard deviation of these sets of 14 estimates. Table 2 shows, for one of the experiments, the root mean square actuator torque estimation errors resulting from the models obtained with the 3 different estimation approaches. Comparison of these results shows that: The load parameters provided by the first approach are not accurate enough, because (1) the efficiencies of the transmissions cannot be included in the linear approach (2) some parameters are only identifiable in combination with others, and (3) the used identification trajectories only guarantee a sufficient excitation for the load parameter identification, not for all other parameters of the robot. The model however yields accurate actuator torque predictions. This can be understood from the fact that the model contains a large number of parameters. These parameters are the degrees of freedom of the parameter estimation which is a least squares minimization of the actuator torque prediction. Approach 3 yields the best results with respect to parameter accuracy. In addition, the actuator torque prediction accuracy of the models resulting from approach 3 and 1 are comparable, and better than that of the models resulting from approach 2. This shows that the inclusion of the efficiency of the transmissions in the model and in parameter estimation is required. Remarks 1. Including a priori knowledge of certain parameters, e.g. rotor inertias, inertial parameters of links 3 to 5, in the parameter estimation problem reduces the uncertainty on the identified load parameters [8]. This has been verified experimentally. Estimation of all parameters in model (2)- (3) yields larger deviations on the load parameter estimates than the ones shown in table 1.

Experimental robot load identification 9 2. The accuracy of the load mass estimation presented in this paper cannot always be reached. Nevertheless, experiments on a large number of industrial robots have shown that the relative error is always smaller than 5%, and consequently, that it is justified to assume the inertial parameters of the link 6 are negligible with respect to those of a typical robot load. Table 1. Overview of experimental results: comparison of the accuracy of the inertial parameter estimates obtained using 3 different approaches Average estimation error Parameter Exact value Approach 1 Approach 2 Approach 3 m [kg] 125.46 13.24 ± 1.1 0.81 ± 1.0 0.85 ± 0.9 c x [mm] 88 1 ± 1.4 1 ± 2 1 ± 1 c y [mm] 32 3 ± 1.1 3 ± 1 2 ± 0.9 c z [mm] 140 4 ± 4.8 6 ± 4 4 ± 5 I xx [kgm 2 ] 38.21 5.96 ± 1.0 1.13 ± 0.8 0.70 ± 0.6 I yy [kgm 2 ] 6.55 5.10 ± 0.6 1.23 ± 0.7 1.72 ± 0.7 I zz [kgm 2 ] 37.13 4.09 ± 0.3 0.46 ± 0.6 0.39 ± 0.2 I xy [kgm 2 ] 7.28 1.06 ± 0.4 0.66 ± 0.6 0.26 ± 0.3 I xz [kgm 2 ] 1.49 0.23 ± 0.2 0.23 ± 0.3 0.34 ± 0.4 I yz [kgm 2 ] 1.23 0.12 ± 0.2 0.46 ± 0.5 0.18 ± 0.1 Table 2. Overview of experimental results: comparison of the RMS actuator torque prediction errors for the 3 different approaches/models (for one experiment) RMS estimation errors Actuator torque range Approach 1 Approach 2 Approach 3 Axis 3 [Nm] 3100Nm 120Nm 84 111 109 Axis 4 [Nm] 1200Nm 880Nm 48 52 42 Axis 5 [Nm] 1200Nm 700Nm 41 48 40 Axis 6 [Nm] 850Nm 770Nm 38 43 39

10 Jan Swevers et al. 5 Conclusion The presented experimental robot load identification method is based on the robot identification method described in [1]: the experiment design and maximum likelihood parameter estimation approaches are adopted. The modelling approach however is different: instead of using a barycentric parameter description of the robot dynamics, the model depends on the inertial parameters of the load. In addition, the model takes into account the dynamics resulting from the robot inertia, coupling between the actuator torques and losses in the motors and the transmissions. The rotor inertias and inertial parameters of the links are set to a priori determined values. Experimental results show that this modelling approach and the inclusion of these effects are essential to obtain sufficiently accurate parameter estimates. The obtained parameter and actuator torque prediction accuracy of the resulting dynamic model satisfy the requirements imposed by industrial users of robots: this new method is therefore recognized by industry as an effective means of robot load identification. Acknowledgement Concerted Research Action GOA/99/04 is gratefully acknowledged. The scientific responsibility is assumed by its authors. The work of Walter Verdonck is sponsored by a specialization grant of the Flemish Institute for Support of Scientific and Technological Research in Industry (IWT) References 1. Swevers, J., Ganseman, C., Tükel, D. B., De Schutter, J., and Van Brussel, H. (1997) Optimal Robot Excitation and Identification, IEEE Transactions on Robotics and Automation, Vol. 13(5), 730-740. 2. Atkeson, C.G., An, C.H., and Hollerbach, J.M. (1986) Estimation of Inertial Parameters of Manipulator Loads and Links, The International Journal of Robotics Research, Vol. 5(3), 101-119. 3. Gautier, M., and Khalil, W. (1992) Exciting Trajectories for the Identification of Base Inertial Parameters of Robots, The International Journal of Robotics Research, Vol. 1(4), 362-375. 4. Olsen, H.B., and Bekley, G.A. (1986) Identification of Robot Dynamics, Proceedings of the IEEE International Conference on Robotics and Automation, San Francisco, USA, 1004-1010. 5. Presse, C. and Gautier, M. (1993) New Criteria of Exciting Trajectories for Robot Identification, Proceedings of the IEEE International Conference on Robotics and Automation, Atlanta, USA, 907-912. 6. Hirzinger, G., Fischer, M., Brunner, B., Koeppe, R., Otter, M., Grebenstein, M., and Schäfer, I. (1999), Advances in Robotics: the DLR experience, The International Journal of Robotics Research, Vol. 18(11), 1064-1087.

Experimental robot load identification 11 7. Craig, J. J. (1989) Introduction to Robotics, Addison-Wesley Publishing Company, New-York. 8. Schoukens, J., and Pintelon, R. (1991) Identification of linear systems, a practical guideline to accurate modeling, Pergamon Press, ISBN 0-08-040734-X. 9. Maes, P., Samin, J.C., and Willems, P.Y. (1989) Linearity of multibody systems with respect to barycentric parameters : dynamic and identification models obtained by symbolic generation, Mechanics of Structures and Machines, Vol. 17(2), 219-237.