Flobaroid PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/315716936

FloBaRoID - A Software Package for the Identification of Robot Dynamics


Parameters

Conference Paper · June 2017


DOI: 10.1007/978-3-319-61276-8_18

CITATIONS READS
7 227

4 authors, including:

Stefan Bethge Nikos G Tsagarakis


Istituto Italiano di Tecnologia Istituto Italiano di Tecnologia
4 PUBLICATIONS   9 CITATIONS    473 PUBLICATIONS   9,557 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

CogIMon – Cognitive Interaction in Motion View project

WALK-MAN: Whole-body Adaptive Locomotion and Manipulation View project

All content following this page was uploaded by Stefan Bethge on 17 February 2020.

The user has requested enhancement of the downloaded file.


FloBaRoID — A Software Package for the
Identification of Robot Dynamics Parameters

Stefan Bethge, Jörn Malzahn, Nikolaos Tsagarakis, Darwin Caldwell

Italian Institute of Technology, Genova, 16163 Italy,


stefan.bethge, jorn.malzahn, nikos.tsagarakis, darwin.caldwell@iit.it

Abstract. This paper presents the open software package FloBaRoID


(FLOating BAse RObot dynamical IDentification), which aims to pro-
vide a package implementing all necessary methods to identify robot
dynamics parameters starting from a kinematic model. The package fea-
tures data acquisition and preprocessing, extraction of identifiable base
dynamics parameters, and finding physically consistent dynamics param-
eters for stable control and simulation. The paper details each of these
steps and exemplifies the software usage with experimental results for
the 7-DOF robot Kuka LWR 4+.

Keywords: robot modelling, dynamics, parameter identification, soft-


ware

1 Introduction
Robot dynamics simulations and model based controllers require precise model
knowledge. The Newton-Euler or Lagrange formalisms allow to derive the equa-
tions of motion that describe the dynamics. The inertial parameters appearing in
this model are typically not provided by robot manufacturers. For custom built
robot prototypes they can be extracted from thoroughly maintained CAD data.
Although for the actuators and the structural robot components this is feasible,
it is much more difficult and time consuming to precisely match the mass dis-
tributions including electrical wires and electronics boards. Manufacturing and
assembly tolerances contribute additional inaccuracies along with unknown fric-
tion parameters. Especially robot prototypes are subject to frequent hardware
revisions and in situ bugfixes or extensions that aggravate CAD maintenance. For
these systems, an experimental identification from motion and force or torque
measurements is an alternative and can also be used to refine CAD models to
improve the force/torque prediction accuracy.
In the past decades, a substantial amount of research on this topic has been
conducted and the reader is referred to [1] for a recent literature overview. The
most popular approach is the linear regression approach[2] on which FloBaRoID
is based. The related closed source project SYMORO+ allows calculating iden-
tification models only for fixed-base dynamics and does not include methods to
obtain any parameters. The open source successor OpenSYMORO adds floating-
base dynamics. Both expect the kinematic model defined in Denavit-Hartenberg
2 Stefan Bethge et al.

parameters, which are not practically suited for all robot types, such as in the
case of e.g. Walk-Man[3].
The aim of FloBaRoID is to provide a generic tool that implements state
of the art dynamic parameter identification techniques using e.g. linear convex
optimization[4] for arbitrary fixed and floating-base robots. FloBaRoID is open
source to allow open community research also beyond the state of the art1 . Flo-
BaRoID can generate optimized identification trajectories (but also works with
suboptimal excitations), pre-processes the sensor readings and finds inertial pa-
rameters ensuring physical consistency. It supports general kinematic models
and reads robot descriptions in the Unified Robot Description Format (URDF),
supports floating-base dynamics for mobile robots and includes external con-
tact forces. The software is implemented in the Python programming language.
For the numerical fixed and floating-base dynamics computations, the iDynTree
library is used[5]. While FloBaRoID can greatly simplify the dynamics iden-
tification task, it cannot provide a one-click solution. The user is required to
understand the mathematical nature and limitations of the inertial parameter
identification problem for multi-body systems. Therefore this paper comprises
a brief introduction into the topic and describes how each step is handled by
FloBaRoID.
The paper is organized as follows: Section 2 gives an overview of the pa-
rameter identification problem while Sections 3, 4 and 5 describe trajectory op-
timization, data preprocessing, and identification methods. Section 6 evaluates
the performance of exemplary identification experiments.

2 Problem Definition
For a system with nlinks bodies connected by ndof joints, the equations of motion
for the rigid multi-body dynamics have the form:
X
τ + ρ = H(q)q̈ + C(q, q̇)q̇ + g(q) + d(q̇) + JlT flext , (1)
l∈L

with:
q : joint position vector (R6+ndof )
τ : joint torque vector (R6+ndof )
ρ : measurement error (R6+ndof )
H : joint-space inertia matrix (R6+ndof ×6+ndof )
C(q, q̇)q̇ : Coriolis and centrifugal torques (R6+ndof , C ∈ R6+ndof ×ndof )
g(q) : Gravity torques (R6+ndof )
d(q̇) : Frictional damping torques (Rndof )
Jl : Jacobian of the contact frame l (R6×6+ndof )
flext : Wrench exerted by external environment on link l (R6 )

1
Source and Documentation is available at https://github.com/kjyv/flobaroid
FloBaRoID 3

The equations can be brought into a parameter linear form[2] with the re-
gressor matrix Y (q, q̇, q̈) and the inertial parameter vector x:
X
τ + ρ = Y (q, q̇, q̈)x + JlT flext (2)
l∈L

and:

Y (q, q̇, q̈) : dynamics regressor (R6+ndof ×(10·nlinks +2·ndof ) ) ,


x : inertial parameter vector (R10·nlinks +2·ndof ) .

For simplicity, the arguments are omitted when it is obvious in the following.
The 10 standard inertial parameters for each link i consist of the link mass (mi ),
the three values of the first moment of mass (cix , ciy , ciz ), the six independent
entries for the 3 × 3 link inertia tensor expressed at the origin of the link frame
(Iixx , Iixy , Iixz , Iiyy , Iiyz , Iizz ). Additionally for each joint j attached between
link j − 1 and j, there is one direction dependent constant friction parameter
(F cj ) and one parameter for velocity dependent friction (F vj )[6]. The parameter
identification problem is then:
! 2
X
min τ − JlT flext − Y x . (3)

x
l∈L

3 Optimized trajectories
The trajectories executed by the robot for experimental identification should be
designed such that (3) is a well-conditioned problem. Under the assumption that
measurement error ρ displays a zero-mean Gaussian probability distribution, (2)
can be solved in the least squares sense to obtain an accurate estimate of the iner-
tial parameters x (in practice however only when Y has full rank, cf. Section 5).
Based on a suitable trajectory parametrization, the trajectory generation can be
formulated as an optimization problem. FloBaRoID uses a smooth periodic rep-
resentation of joint space trajectories by means of parametrized Fourier-series[7].
A global particle swarm optimizer (ALPSO), optionally followed by a local gradi-
ent based solver (IPOPT), searches for optimal trajectory parameters. Values for
the trajectory boundaries in terms of joint angle limits, maximum velocity and
torque limits (using torques estimated from CAD parameters as upper bound)
are taken from URDF to constrain the search space and reduce computation
time.
Calculating the dynamics for each step of the parameter optimization is done
using iDynTree, which does not handle external contacts and as such the opti-
mization can only be used for fixed-based robots at the moment. Finding op-
timal trajectories for floating-base robots such as humanoids is a problem of
much higher complexity and is not easily integrated into a generic package. A
possible approach is described in [8]. However, using data from existing balanced
movements, parameters for floating-base robots can also be identified.
4 Stefan Bethge et al.

FloBaRoID interfaces with the robot middlewares YARP and ROS, so that
trajectories can be executed on a real robot with minimal efforts. The open
architecture of FloBaRoID generally allows the extension to other robot control
interfaces upon demand.

4 Preprocessing

The data to be acquired comprises joint positions and torques as well as con-
tact forces, if available. Numerical differentiation yields velocity and acceleration
signals. Zero phase forward-backward digital filtering using a Butterworth filter
reduces signal and differentiation noise. The filter cut-off frequencies may need
to be changed from the defaults in order to not filter out signal frequencies or to
reduce noise further, as the optimal values depend on the individual robot mass,
operating velocities and sensor properties. For floating base robots, the base link
orientation, linear acceleration and rotational velocity readings are derived from
inertial measurement unit data. The rotational acceleration is derived through
numerical differentiation and filtering as described above. The base link linear
velocity does not influence the joint torques[9] and can be ignored.

5 Identification

A straight forward method to solve (3) is the ordinary least squares estimation
(OLS) method based on the pseudoinverse Y + of the regressor matrix Y in (2).
For most practical robots, this is however not directly possible for two reasons:

1. Not all of the standard inertial parameters actually influence the dynamics.
They are called non-identifiable parameters and correspond to zero columns
in Y .
2. Other parameters cannot be identified independently as their columns in Y
are linear dependent. These parameters can only be identified together and
individual values are unknown.

Both reasons cause the regressor matrix to be rank deficient and the problem is
singular.

5.1 Base Parameter Identification

To resolve this issue, [10] introduces the base regressor Yb with:

Yb = Y B . (4)

The base regressor matrix Yb is the surjective projection of the standard regressor
matrix Y to a linear independent subspace using a projection matrix B. There-
fore, Yb comprises a non-unique choice of linearly independent columns of Y . The
FloBaRoID 5

corresponding non-unique choice of base parameters is xb with dim xb ≤ dim x.


After projection, (2) can be restated and solved for the set of base parameters:
X
xb = Yb+ (τ + ρ) + Yb+ JlT flext , (5)
l∈L

minimizing the torque prediction error:


2
minkτ − Yb xb k . (6)
xb

5.2 Standard Parameter Identification


The identified base parameter vector xb is sufficient to predict joint torque read-
ings. However, in order to update URDF files and improve simulators, it is
beneficial to retrieve a full set of standard inertial parameters x. Since the pro-
jection (4) is not bijective, there is no unique solution for x and among the
possible solutions, not all are physically consistent.
A standard parameter vector x could be obtained from:
x = Bxb (7)
and this vector would still minimize the torque prediction error (3). However it
likely includes physically inconsistent parameter values such as negative masses,
center of mass coordinates far outside the link body or inertia values that do not
yield positive definite link inertia matrices. Such values threaten model based
controller and simulation stability.

5.3 Physical Consistency Through Constrained Optimization


Improving the data quality and operation space coverage for instance through
trajectory optimization, better sensors and Weighted Least Squares techniques
(WLS, [11]) based on the expected parameter value magnitude or estimated co-
variance, can reduce the likelihood of arriving at inconsistent parameter values.
Nevertheless, the only way to guarantee physical consistency of the obtained pa-
rameter values is to translate physical consistency into mathematical constraints,
while solving (6) as a constrained optimization problem.
FloBaRoID implements a method using semidefinite programming (SDP)[4],
which solves the constrained optimization problem directly for (3) in the stan-
dard parameter space. This way, physically consistent solutions x are found that
minimize the estimation error. Alternative successful non-linear optimization ap-
proaches are reported in the literature but they cannot formally guarantee to
find a global minimum of the estimation error[4].
Using the projection matrix B that combines the standard columns to the
base columns, FloBaRoID identifies physically consistent standard parameters
directly using the SDP program:
minimize u
(u,Bxstd )

subject to Uτ (u, Bxstd )  0 ,


6 Stefan Bethge et al.

with
(τ − Ybase Bxstd )T
 
u
Uτ (u, xstd ) = .
τ − Ybase Bxstd 1
The matrix Uτ is the Schur complement of:

u ≥ (τ − Ybase Bxstd )T (τ − Ybase Bxstd ) , (8)


2
which is equal to minimizing u subject to u ≥ kτ − Ybase Bxstd k .
As the non-identifiable parameters are not expressed in the measurement
data, it is in general hard to find the exact parameters of the physical bodies,
regardless of the measurement data quality. Results can be improved depending
on the chosen constraints, thereby also influencing non-identifiable parameters.
In addition to physical consistency, other constraints such as boundaries using
previously known values from CAD data, directly measured parameters, sym-
metries or otherwise known values can be included as well.

6 Application Examples
This section exemplifies the dynamic parameter identification with FloBaRoID
in two stages based on the Kuka LWR 4+ robot arm. The first stage is an
identification performed in an ideal simulation environment. The second example
is based on experimental data acquired on the real hardware.

6.1 Simulations
Two URDF files with standard parameters of a Kuka LWR 4+ arm are used for
simulations. One is part of a package that is available from Centre E. Piaggio,
Pisa2 , which also allows to connect the Kuka FRI library to standard ROS tools.
This URDF represents the “real” parameters to be identified in the simulation.
A second URDF file with quite different mass distribution from the CoR-Lab at
University of Bielefeld3 is used. It serves as prior knowledge during the identifica-
tion, simulating knowledge from CAD data which allows to formulate parameter
constraints in addition to the physical consistency constraints.
As it appears, both URDF files do not contain identified parameters, because
many of the values have exact decimals. Regardless of the correctness with re-
spect to the real hardware, the simulation objective is to demonstrate how well
the difference in model parameters can be recovered through identification. The
example leverages the fact that the simulated sensor readings are bias and noise
free. With optimal excitation trajectories, the base parameters can be identi-
fied up to machine precision. The accuracy regarding the identified standard
parameters depends on the constraints and the distance of the prior parameter
values to the actual ones, because there are theoretically infinitely many consis-
tent solutions as described in Section 5.2. For this example, the link masses are
2
https://github.com/CentroEPiaggio/kuka-lwr/
3
https://github.com/corlab/cogimon-gazebo-models
FloBaRoID 7

Table 1: Identified standard parameters of the simulated Kuka LWR 4+. Non-
identifiable parameters are marked with *. These have no effect on dynamics
and are determined only to satisfy consistency constraints.

Parameter Real Identified Parameter Real Identified Parameter Real Identified


m1 2.7000 2.6581* I3xx 0.0738 0.0671 I5yz 0.0000 0.0014
c1x 0.0000 −0.0000* I3xy 0.0000 0.0000 I5zz 0.0092 0.0066
c1y −0.1620 −0.0265* I3xz 0.0000 0.0000 m6 1.6000 1.6999
c1z 0.3510 0.3232* I3yy 0.0497 0.0449 c6x 0.0000 −0.0000
I1xx 0.0738 0.8044* I3yz −0.0211 0.0029
c6y 0.0000 0.0074
I1xy 0.0000 0.0000* I3zz 0.0257 0.0239
c6z 0.1000 0.0969
I1xz 0.0000 −0.0000* m4 2.7000 2.4156
I6xx 0.0104 0.0351
I1yy 0.0497 0.8041* c4x 0.0000 −0.0000
I1yz 0.0211 0.0033* c4y −0.1620 −0.0422 I6xy 0.0000 −0.0000
I1zz 0.0257 0.0272 c4z 0.1890 0.2148 I6xz 0.0000 −0.0000
m2 2.7000 2.8714* I4xx 0.0414 0.0678 I6yy 0.0104 0.0325
c2x 0.0000 0.0000 I4xy 0.0000 −0.0000 I6yz 0.0000 −0.0000
c2y 0.1620 0.0431* I4xz 0.0000 −0.0000 I6zz 0.0042 0.0068
c2z 0.1890 0.3149 I4yy 0.0173 0.0419 m7 0.3000 0.3399
I2xx 0.0414 0.0999 I4yz 0.0113 0.0113 c7x 0.0000 0.0000
I2xy 0.0000 0.0000 I4zz 0.0257 0.0276 c7y 0.0000 0.0000
I2xz 0.0000 0.0000 m5 1.7000 2.0427 c7z 0.0000 −0.0000
I2yy 0.0173 0.0772 c5x 0.0000 0.0000 I7xx 0.0500 0.0277
I2yz −0.0113 −0.0113 c5y 0.0000 −0.0074 I7xy 0.0000 −0.0000
I2zz 0.0257 0.0243 c5z 0.2108 0.0620 I7xz 0.0000 −0.0000
m3 2.7000 2.8194 I5xx 0.0369 0.0342 I7yy 0.0500 0.0277
c3x 0.0000 −0.0000 I5xz 0.0000 0.0000 I7yz 0.0000 0.0000
c3y 0.1620 0.0422 I5yy 0.0287 0.0287
I7zz 0.0500 0.0500
c3z 0.3510 0.1220 I5xy 0.0000 −0.0000

constrained to an interval of ±50% around the a priori masses, while the center
of mass coordinates are constrained to the enclosing hulls of CAD meshes of
each link. To excite the dynamics, the robot performs an optimized trajectory
motion. A different trajectory from random parameters is used as validation
dataset. Table 1 lists the identification result. Although not all parameters get
identified exactly, the squared distance of the a priori parameter vector to the
ground truth parameters is decreased from 1.6353 to 0.3552. At the same time
the torque prediction normalized RMS error is reduced from 0.907% to 0.022%
on the identification data set and to 0.04% on the validation data set.

6.2 Experiments

For the experimental example with real measurements, the robot is controlled to
follow the same optimized trajectory as before. As the true inertial parameters
of the robot are unknown, only the torque prediction error on the validation
data set can be evaluated as a measure of accuracy. In addition, the identified
standard parameters are referenced against the identified parameters reported
in [12], for which however the authors cannot confirm physical consistency.
8 Stefan Bethge et al.

The identified standard parameters reduce the normalized RMS torque pre-
diction error from 0.90% with the parameters from [12] to 0.29% on the identifi-
cation data set, while the validation error for a random trajectory was reduced
from 0.81% to 0.36%. The resulting torque estimations are shown in Figure 1
and identified parameters in barycentric form are given in Table 2.
While the measured data is predicted accurately for smooth motions, direc-
tion changes can produce physical effects that are not modeled such as static
friction or backlash. The maximum amplitude of the errors is within ±2.23 Nm
or less for each joint, which corresponds to 1.6% or less of the maximum motor
torques of each joint, cf. fig. 1. The differently large errors possibly result from
differently large torques produced by the trajectory so that measurement and
identification errors play a larger part. Also these joints might simply have expe-
rienced more wear and exhibit a higher amount of unmodelled effects, e.g. joint 5.
Hence, modelling these effects may increase the overall estimation accuracy for
this particular robot further, the overall error is however already small.

40
10
τ1 (Nm)

τ2 (Nm)

τ3 (Nm)
5
20
0 0
0
−20
−10
−5 −40
0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10

10 1 1
τ4 (Nm)

τ5 (Nm)

τ6 (Nm)

0 0
0
−1 −1
−10 −2 −2

0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10

0.5
Maxima of absolute errors, τ1 : 1.36 Nm (0.38%),
τ7 (Nm)

0
τ2 : 2.23 Nm (0.63 %), τ3 : 1.69 Nm (0.84 %), τ4 :
−0.5 2.16 Nm (1.08 %), τ5 : 1.31 Nm (0.64 %), τ6 : 1.21
−1 Nm (1.60 %), τ7 : 0.43 Nm (0.57 %).
0 2 4 6 8 10

Fig. 1: Plots of estimated (orange) and measured torques (green) and estimation
error (blue, dashed) for each joint using optimized trajectories on a physical
Kuka LWR 4+.

7 Conclusions and Outlook


The paper presents the robot dynamics identification software package Flo-
BaRoID. It briefly introduces the reader into the topic of robot dynamics identi-
fication and exemplifies the software usage based on the dynamics identification
of a Kuka LWR 4+ robot arm. The experimental identification example is ref-
erenced against the identified parameters reported by [12]. It has to be noted
FloBaRoID 9

Table 2: Identified standard parameters including link side Coulomb and viscous
friction for Kuka LWR 4+ compared to values from [12]. Parameters can also
be found as URDF within the FloBaRoID git repository.

Parameter Prior Identified Parameter Prior Identified Parameter Prior Identified


m1 4.1948 2.6994* I3yz −0.0006 0.0013 I6xz −0.0000 0.0121
c1x −0.0216 0.0000* I3zz 0.0500 0.0001 I6yy −0.0000 0.0017
c1y 0.0100 −0.0579* m4 2.3847 2.6948 I6yz −0.0011 −0.0029
c1z −0.0377 0.1300* c4x −0.0015 −0.0002 I6zz 0.0035 0.0049
I1xx 0.0100 0.0191* c4y −0.0000 0.0155 m7 0.6501 0.4250
I1xy 0.0100 −0.0000* c4z 0.1486 0.1915 c7x −0.0005 −0.0113
I1xz 0.0100 −0.0000* I4xx 0.0124 0.0670 c7y 0.0015 0.0399
I1yy 0.0019 0.0040* I4xy 0.0000 0.0237 c7z 0.0716 −0.0310
I1yz 0.0100 0.0000* I4xz −0.0005 −0.0121 I7xx 0.0000 0.0033
I1zz 0.0100 0.0000 I4yy 0.0073 0.0193 I7xy −0.0000 0.0106
m2 4.2997 2.7000* I4yz −0.0005 −0.0095 I7xz 0.0000 0.0028
c2x 0.0003 0.0339 I4zz 0.0100 0.0047 I7yy 0.0000 0.0339
c2y −0.0041 0.0555* m5 1.7036 2.8900 I7yz −0.0000 0.0090
c2z 0.0824 0.1221 c5x −0.0004 0.0041 I7zz 0.0001 0.0024
I2xx 0.0474 0.2033 c5y −0.0554 0.0138 F c0 0.0000 0.9442
I2xy −0.0000 −0.0048 c5z −0.0101 0.0027 F c1 0.0000 0.0805
I2xz 0.0001 −0.0170 I5xx 0.0063 0.0432 F c2 0.0000 0.0401
I2yy 0.0500 0.1668 I5xy −0.0002 0.0329 F c3 0.0000 −0.3905
I2yz −0.0009 0.0904 I5xz 0.0000 0.0042 F c4 0.0000 −0.4683
I2zz 0.0016 0.0500 I5yy 0.0012 0.0250 F c5 0.0000 −1.2011
m3 3.6585 2.5608 I5yz −0.0050 0.0032 F c6 0.0000 −0.2563
c3x 0.0003 0.0088 I5zz 0.0071 0.0004 F v0 0.0000 0.6737
c3y 0.1137 −0.0388 m6 0.4001 0.3300 F v1 0.0000 0.7468
c3z −0.0001 0.0368 c6x 0.0021 −0.0321 F v2 0.0000 0.3076
I3xx 0.0470 0.0023 c6y 0.0586 −0.0129 F v3 0.0000 0.8421
I3xy −0.0003 −0.0052 c6z −0.0448 −0.0025 F v4 0.0000 0.8381
I3xz 0.0000 −0.0005 I6xx 0.0005 0.0295 F v5 0.0000 0.2333
I3yy 0.0008 0.0120 I6xy 0.0000 −0.0071 F v6 0.0000 0.1931

that the difference in the prediction error norms between the two parameter
sets is likely partially due to parameter scattering among the different units of
the same type. The units under consideration are produced at different times in
different lots and have an entirely different operation history. However, the op-
timization method used in [12] does not guarantee finding global optima which
the difference in accuracy can possibly also be attributed to.
The simulation and experimental example given in this paper demonstrate
the potential and challenges associated with state of the art robot dynamics
identification. The targeted scope of FloBaRoID is to support the identification
of fixed-base and of tree-like floating-base robots such as humanoids with a high
number of degrees of freedom. For such robots, additional challenges occur. For
instance, classical excitation trajectory optimization approaches may lead to a
loss of balance during the experiment. The integration of balancing constraints
in the optimization as well as elasticity in the equations of motion are planned
for future versions.
10 REFERENCES

References

[1] J. Wu, J. Wang, and Z. You, “An overview of dynamic parameter identi-
fication of robots,” Robotics and Computer-Integrated Manufacturing, vol.
26, no. 5, pp. 414–419, 2010.
[2] C. G. Atkeson, C. H. An, and J. M. Hollerbach, “Estimation of inertial
parameters of manipulator loads and links,” The International Journal of
Robotics Research, vol. 5, no. 3, pp. 101–119, 1986.
[3] N. G. Tsagarakis, D. G. Caldwell, A. Bicchi, F. Negrello, M. Garabini,
W. Choi, L. Baccelliere, V. Loc, J. Noorden, M. Catalano, et al., “Walk-
man: A high performance humanoid platform for realistic environments,”
Journal of Field Robotics (JFR), 2016.
[4] C. D. Sousa and R. Cortesão, “Physical feasibility of robot base inertial
parameter identification: A linear matrix inequality approach,” The Inter-
national Journal of Robotics Research, vol. 33, no. 6, pp. 931–944, 2014.
[5] F. Nori, S. Traversaro, J. Eljaik, F. Romano, A. Del Prete, and D. Pucci,
“Icub whole-body control through force regulation on rigid non-coplanar
contacts,” Frontiers in Robotics and AI, vol. 2, p. 6, 2015.
[6] P. Dupont, “Friction modeling in dynamic robot simulation,” Proceedings.,
IEEE International Conference on Robotics and Automation, pp. 1370–
1376, 1990.
[7] J. Swevers, C. Ganseman, D. Bilgin Tükel, J. De Schutter, and H. Van
Brüssel, “Optimal robot excitation and identification,” IEEE Transactions
on Robotics and Automation, vol. 13, no. 5, pp. 730–740, 1997.
[8] V. Bonnet, P. Fraisse, A. Crosnier, M. Gautier, A. González, and G. Ven-
ture, “Optimal Exciting Dance for Identifying Inertial Parameters of an
Anthropomorphic Structure,” IEEE Transactions on Robotics, vol. 32, no.
4, pp. 823–836, 2016.
[9] A. Del Prete, N. Mansard, O. E. Ramos, O. Stasse, and F. Nori, “Imple-
menting Torque Control with High-Ratio Gear Boxes and without Joint-
Torque Sensors,” International Journal of Humanoid Robotics (submitted
to), vol. 13, no. 1, pp. 1–29, 2015.
[10] H. Mayeda, K. Yoshida, and K. Osuka, “Base parameters of manipulator
dynamic models,” IEEE Transactions on Robotics and Automation, vol.
6, no. 3, pp. 312–321, Jun. 1990.
[11] G. Zak, B. Benhabib, R. G. Fenton, and I. Saban, “Application of the
Weighted Least Squares Parameter Estimation Method to the Robot Cal-
ibration,” Journal of mechanical design (1990), vol. 116, no. 3, p. 890,
1994.
[12] C. Gaz and A. D. Luca, “Extracting Feasible Robot Parameters from Dy-
namic Coefficients using Nonlinear Optimization Methods,” Proceedings.,
IEEE International Conference on Robotics and Automation, pp. 2075–
2081, 2016.

View publication stats

You might also like