Vous êtes sur la page 1sur 112

Analytical and Parametric Study of

Damped Vibrations in Gough-


Stewart Platforms

Behrouz Afzali-Far

LICENTIATE DISSERTATION

Machine Elements, Mechanical Engineering,


Lund University, Sweden.
March 2014
Analytical and Parametric Study of
Damped Vibrations in Gough-
Stewart Platforms

Behrouz Afzali-Far
Copyright Behrouz Afzali-Far

Division of Machine Elements,


Department of Mechanical Engineering

ISRN LUNTFD2/TFME14/3007SE(1-112)

ISBN 978-91-7473-965-7 (Print)


ISBN 978-91-7473-966-4 (Pdf)

Printed in Sweden by Media-Tryck, Lund University


Lund 2014

A part of FTI (the Packaging and


Newspaper Collection Service)
Preface

The work presented in this licentiate dissertation was initiated with the col-
laboration of Lund Telescope Group AB and ever since has been carried out
at the Division of Machine Elements and Division of Mechanics, Department
of Mechanical Engineering, Lund University. This work has been supervised
by my academic supervisors, Per Lidstrm, Kristina Nilsson and Anette An-
dersson as well as my former industrial supervisors, Arne Ardeberg and Tor-
ben Andersson. I would like to thank them all for their valuable guidance
throughout the project. Especially, I wish to express my sincere appreciation
to Per Lidstrm, my main supervisor, without whom this dissertation would
not have been possible. I would also like to acknowledge Lars Vedmar and
Mats Andersson for their support of this project.

Lund, March 2014


Behrouz Afzali-Far

1
2
Abstract

This work establishes a comprehensive and fully parametric model for the
damped vibrations of Gough-Stewart Platforms (GSPs) at symmetric con-
figurations. It is noteworthy that in the literature a complete solution to
this problem has not been presented. This work has been carried out in three
stages which are the subjects of Papers I, II and III. After the industrial stage
of the work which is reflected in Paper I, we have generalized the problem
and established a fully parametric model of the damped vibrations of GSPs
in Paper II and finally extended the model by taking inertia of the struts into
consideration in Paper III. The final model is parametrically developed in
terms of all the design variables of the system and can be directly employed
for the analysis, optimization and control of GSPs.
In this work, we parametrically formulate the kinematic equations which
are eventually obtained in the form of a Jacobian matrix. The focus of this
work is on the Cartesian-space formulation in which Bryant angles, due to
their advantages over Euler angles, are chosen to represent the orientation of
the platform. The equations of motion are formulated and linearized based
on a Lagrangian dynamics approach where Rayleigh axial damping of the
struts is introduced to the system. Inertia, stiffness and damping matrices
are parametrically formulated. By introducing the inertia of the struts, the
inertia matrix turned out to be quite complicated to formulate. Interestingly,
despite its apparent symmetrical geometry, the equivalent inertia matrix is
obtained as a non-diagonal matrix. The eigenvectors and damped eigenfre-
quencies of the system are then parametrically established. In addition, the
decoupled vibrations are analytically investigated where it is shown that the
consideration of strut inertia may lead to significant changes of the decoupled
conditions. Finally, for a reference GSP, the vibrational behavior with respect
to different design variables are systematically studied.

3
4
List of papers

In the following papers the author Behrouz Afzali-Far has written the major
part of the text and performed the analytical and numerical calculations.

Paper I
Behrouz Afzali-Far, Per Lidstrm, Kristina Nilsson and Arne Ardeberg,
Analytical Stiffness Optimization of High-Precision Hexapods for Large Opti-
cal Telescope Applications,
Proceedings of the 25th Nordic Seminar on Computational Mechanics, (2012).

Paper II
Behrouz Afzali-Far, Per Lidstrm, Kristina Nilsson,
Parametric Damped vibrations of Gough-Stewart Platforms for Symmetric
Configurations,
(Technical Rep. ISRN LUTFD2/TFME-13/1014-SE (1-21), Division of Me-
chanics, Lund University, Submitted to Journal of Mechanism and Machine
Theory, 2013)

Paper III
Behrouz Afzali-Far, Anette Andersson, Kristina Nilsson and Per Lidstrm,
Parametric Damped Vibrations of Gough-Stewart Platforms Taking into Ac-
count Inertia of the Struts,
(Technical Rep. ISRN LUTFD2/TFME-14/1014-SE (1-21), Division of Ma-
chine Elements, Lund University, Submitted to Journal of Mechanism and
Machine Theory, 2014)

5
6
Contents

Notations 9

1 Introduction 11
1.1 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2 Industrial background of the project . . . . . . . . . . . . . . . 15

2 Parametric kinematics 17
2.1 Geometrical design space . . . . . . . . . . . . . . . . . . . . . 17
2.2 Configuration spaces . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Kinematic equation and angular velocity . . . . . . . . . . . . . 19
2.4 Jacobian matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3 Parametric equations of motion 23


3.1 Design variables . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Equations of motion . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3 Linearized equations of motion . . . . . . . . . . . . . . . . . . 25

4 Parametric modal analysis 27


4.1 Eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.2 Eigenfrequencies . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.3 Decoupled vibrations . . . . . . . . . . . . . . . . . . . . . . . . 29

5 Numerical evaluation 31
5.1 Characteristics of the reference GSP . . . . . . . . . . . . . . . 31
5.2 FEM simulation . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.3 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . 32

6 Summary and future work 37


6.1 Summary and conclusions . . . . . . . . . . . . . . . . . . . . . 37
6.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

7
8 CONTENTS

Bibliography 38

Paper I 45

Paper II 51

Paper III 79
Notations

The notations including the subscripts used in the present dissertation are
listed in this chapter. In addition, the special mathematical notations used
for the kinematic formulation can be found in the appendix of Paper II.

Subscripts
i Corresponding to each strut i = 1, ..., 6
k Corresponding to each Cartesian-space coordinate k = 1, ..., 6
st Strut
J Joint-space representation
C Cartesian-space representation
p Related to the platform
b Related to the base
0 Representing the neutral configuration

Notations
b Half angel between two neighboring joints on the base
p Half angel between two neighboring joints on the platform
= b + p
k Axial stiffness of the struts
c Axial viscous damping of the struts
Ratio defined by = kc
Gs Geometrical design space
Ds Design space

9
10 CONTENTS

(Ob , exb , eyb , ezb ) Base-frame with Ob as the origin


(Op , exp , eyp , ezp ) Platform-frame with Op as the origin
Ob , Op Origin of the frames
cp Platforms center of mass
mp Mass of the platform
c
Ixxp Platforms moment of inertia with respect to axis (cp , exp )
c
Iyyp Platforms moment of inertia with respect to axis (cp , eyp )
c
Izzp Platforms moment of inertia with respect to axis (cp , ezp )
Rb Radius of the base-joint circle
Ist Struts moment of inertia
R Radius ratio of the GSP
H Height ratio of the GSP
h Height ratio of the platforms center of mass
s Distance between dmst and base joints
B Base joints
P Platform joints
q Generalized coordinates
dmst Mass element, related to the struts
Mass density of the struts
l Length of struts
h Distance between Op and the platform-plane
H Initial distance between Ob and the platform-plane
T Kinetic energy
V Potential energy
D Damping dissipation
L Lagrangian
J Jacobian matrix
Angular velocity
l Length vector along struts
p Position vector of the platform joints relative to Op
p0 Projection of p on the platform plane
h Vector defined by h = r Op Op0
b Position vector of base joints relative to Ob
r Position vector of dmst relative to the base joints
t Position vector of Op relative to Ob
R Rotation tensor
Chapter 1

Introduction

1.1 Literature review


Robots are categorized by their kinematic architecture, where parallel kine-
matic robots are those with a closed kinematic chain. Gough-Stewart Plat-
forms (GSP), among the six Degrees Of Freedom (DOFs) parallel kinematic
robots, has attracted the most attention of researchers and engineers. This
has been the case ever since the concept of GSP was introduced by Gough and
Stewart in 1947 and 1965, respectively [13]. In some applications it is also
known briefly as hexapod or Stewart platform. It has the simplest symmetric
structure with parallel architecture which can provide six Degrees-Of-Freedom
(DOFs) by its six linear actuators. Due to its parallel architecture, inherently
it has some advantages over serial robots including higher stiffness, higher
accuracy, elimination of the cumulative error, etc. [4]. GSPs have been in-
creasingly employed in several modern applications such as robotic surgery,
astronomical telescopes and manufacturing technology [57], see Fig. 1.1.1.
Despite this wide range of applications, in most of the cases the core
design requirement is the dynamic stability which can be evaluated by modal
analysis. A GSP can be defined by its design variables which represent its
geometry, stiffness, damping and inertia and it is obvious that the modal
behavior of a GSP is a function of these design variables. Reversely, the design
variables of a GSP can be optimized based upon the requirements of modal
behavior at the initially symmetric configuration. Modal behavior is referred
to the eigenfrequencies and eigenvectors of a system. Due to the multiple
DOFs and the complexity of the system, developing a fully parametric modal
analysis is crucial in order to accomplish an optimized GSP corresponding to
the required properties of each mode. In addition to analysis and optimization

11
12 CHAPTER 1. INTRODUCTION

(a) A surgery robot [7] (b) A simulator [8]

(c) A collimation system of an astronomical (d) NEXLINE Piezo Hexapod [10]


telescope [9]

Figure 1.1.1: Some of the modern applications of GSPs

of GSPs, a parametric model can also be used in the model-based control


systems without any remodeling efforts.
To obtain such a parametric model, respectively kinematics, linearized
equations of motion and modal analysis should be carried out parametrically.
Generally, forward and inverse kinematics of GSPs have been relatively more
studied in the literature (see e.g. [1113]).
Regarding the dynamics of GSPs, the Lagrangian method, see [11, 14
16], and the Newton-Euler method, see [1720], are most frequently used
1.1. LITERATURE REVIEW 13

in the literature. However, due to the lack of strut flexibility or the lack of
parametrized equations of motion in terms of design variables in these models,
these references could not be directly used to carry out the modal analysis.
The linearized equations of motion are identified by stiffness, damping and
mass matrices. These matrices can be expressed in terms of the Cartesian-
space or joint-space coordinates, see Section 2.2. In the Cartesian-space, the
stiffness matrix is a function of the flexibility of the struts as well as the
geometry properties. Due to the complexity of the stiffness matrix for the
entire workspace, often a semi-analytical or a finite element approach is used.
The semi-analytical approach can be a formulation of the stiffness matrix
simply in terms of the Jacobian matrix without a parametric representation in
terms of the design variables [2124]. Parametric formulations of the stiffness
matrix for some special conditions can be found in [2527].
In the Cartesian-space, the derivation of the inertia matrix is challeng-
ing due to the influence of the strut inertia. To simplify the model, many
researchers have ignored the strut inertia and assumed the inertia matrix in
the Cartesian-space simply as a diagonal matrix containing only the inertia
properties of the platform [2832]. The inertia of the rigid struts has been
considered in some dynamic models (see e.g. [5, 14, 33, 34]), but due to the
lack of flexibility in the system, these models cannot be employed for a modal
analysis. Moreover, in some control models, the strut inertia is simply esti-
mated to be a diagonal matrix in the joint-space [3539]. Hence, generally
a parametric equivalent inertia matrix of the system is not addressed in the
literature.
Performing the analytical modal analysis is a demanding process which
requires major simplifications of the complex equations that are obtained.
Due to these complexities, a straightforward and simple formulation of the
vibrational problem can only be done for the neutral configuration. Gener-
ally, the dynamic performance of GSPs at the neutral configuration is of great
importance for the design optimization. The importance of the neutral config-
uration in optimization is further motivated by Jiang et al. [40]. Furthermore,
to obtain more realistic eigenfrequencies of GSPs, it is also essential to con-
sider damping in the system. In particular, the axial viscous damping of the
struts should be added into the model. This is due to the structural damping
of the system as well as additional damping elements in some applications
(e.g. shock absorbers) which could lead to a dramatic change in the vibra-
tional behavior. However, axial viscous damping has not been parametrically
considered in the literature.
Specifically regarding the modal analysis of GSPs, only a limited number
14 CHAPTER 1. INTRODUCTION

of papers have been published which can be categorized into finite element
[41], semi-parametric [31, 4247] and parametric [28, 32] approaches. None of
these present a parametric modal analysis taking into account the inertia or
damping of the struts. Among the semi-parametric approaches, Wang et al.
[47] take the inertia of the struts into account for a hydraulic GSP with a semi-
analytical approach where the equivalent inertia matrix and the eigenvalue
problem are solved numerically. In an effort to construct the equivalent inertia
matrix, Mahboubkhah et al. [43] expand the inertia matrix to a 2424 matrix
by considering three masses on each strut, while a parametric solution to the
eigenvalue problem is not presented in this work. However, to study the modal
properties of the six modes, obtaining an equivalent 66 inertia matrix seems
to be a much more efficient solution than introducing extra DOFs and adding
complications by a 24 24 inertia matrix.
Very few researchers formulate the vibrations of GSPs parametrically and
none of them consider the strut inertia or damping. Jiang et al. [28] solved
the eigenvalues of the inverse inertia matrix in joint-space and obtained the
undamped vibrations parametrically. Tong et al. [32] formulated the un-
damped eigenvalues for a class of decoupled modified GSPs. Hence, in the
literature, a complete parametric solution to the damped eigenvalue problem
for the neutral configuration of GSPs is not available and in particular a para-
metric formulation of damped eigenfrequencies in terms of design variables is
not addressed. Furthermore, neither a parametric modal analysis nor a for-
mulation of the equivalent inertia matrix is addressed in the literature where
the strut inertia is taken into account.
After the industrial stage of the present work (see Paper I), in Paper II
we have generalized the industrial problem and established a fully-parametric
and closed-form formulation for the vibrations of GSPs where we have addi-
tionally introduced strut damping in the system. Parametric expressions of
the damped eigenfrequencies and the corresponding eigenvectors as well as
the Jacobian, stiffness and damping matrices are developed, all parametri-
cally in terms of the design variables. Interestingly, despite the complexity
of the system, it is shown how well-structured algebraic expressions are ob-
tained using the Cartesian-space approach. Having analytically studied the
eigenvectors, the necessary and sufficient conditions for fully decoupled vi-
brations are formulated. Finally, using a reference GSP, the sensitivity of the
damped eigenfrequencies to stiffness and damping variations are investigated
accompanied by a cross-check with an FEM simulation.
In paper III, we develop a fully parametric solution to the damped vibra-
tions of GSPs taking into account the strut masses and moments of inertia.
1.2. INDUSTRIAL BACKGROUND OF THE PROJECT 15

For the first time, a combined 6 6 equivalent inertia matrix (mass matrix),
including both the inertia properties of the platform and the struts, is formu-
lated parametrically. Consequently, the eigenvectors and the damped eigen-
frequencies are also parametrically developed in the Cartesian-space. Fur-
thermore, the conditions for the decoupled vibrations and the effect of strut
inertia on these conditions are analytically investigated. With the aid of the
reference GSP, the influence of the strut inertia on its modal behavior is sys-
tematically investigated. Therefore, Paper II and Paper III may serve as a
comprehensive and accurate parametric model of the GSP vibrations that
can be directly used in the analysis and optimization of any application. In
addition, it can also be employed as an accurate basis of model-based control
systems.

1.2 Industrial background of the project


Cassegrain-type large optical telescopes have two mirrors; the secondary mir-
ror (M2) is the optical element which reflects the light collected by the primary
mirror (M1) to the scientific instrument [48]. Active collimation of M2 cor-
rects image-quality defects due to structure deformations caused by gravity
and thermal variations. Due to the tracking velocity of optical telescopes,
the bandwidth required is approximately 0.01 Hz [49]. The image-quality of
a stable telescope with a well-figured M1 surface and well-collimated M2 is
limited by the atmospheric disturbances.
Further improvement of the image-quality requires correction of at least
first-order defects due to the atmospheric disturbances as well as errors caused
by telescope structural vibrations. Such first-order defects of atmospheric
disturbances can be compensated by tilting M2 with a sufficiently high band-
width. In practice, tilting is responsible for a major part of the image degra-
dation caused by the atmosphere. Given the importance of tilting, major
improvements of observed image-quality can be achieved through two-axes
rotations (tip-tilt), first-order compensatory movements of M2 [50]. Success-
ful operation requires a very stiff yet light-weight system including the active
mechanism, M2 and its supporting structure. The tip-tilt movements are con-
trolled with a bandwidth of at least 10 Hz, in closed-loop mode via an optical
sensor. The corresponding angular range should be several arcsec.
Mainly due to precision and stiffness requirements for the collimation sys-
tem of M2, high-precision GSPs are used in some modern telescopes [51]. For
further improvements of image-quality, a stiff high-bandwidth GSP can pro-
vide three additional useful degrees of freedom (decentering in two axes and
16 CHAPTER 1. INTRODUCTION

focusing) compared with only using tip-tilt. Hence, due to the fact that the
control bandwidth of a GSP is mechanically restricted by its natural frequen-
cies, the optimization of the natural frequencies is of great importance.
This project was, based on a demand for a high-bandwidth hexapod of
a large optical telescope, initiated by Lund Telescope Group AB and Lund
University. This industrial problem is further explained in Paper I. It should
be noted that we have extended this industrial problem to a general study of
the damped vibrations in GSPs which is the subject of this dissertation.
Chapter 2

Parametric kinematics

2.1 Geometrical design space


A GSP has six joints on the planar base and six joints on the planar platform
with a tri-symmetrical arrangement (see Figs. 2.1.1, 2.1.2). A combination
of universal and spherical joints for the base and the platform respectively, is
the most common design of existing GSPs [52]. This combination is recog-
nized as a 6-Universal-Prismatic-Spherical (6-UPS). It should be noted that
the formulation of the kinematics presented in this chapter is independent of
the joint type. The geometry is parametrized using a set of geometrical design
variables which are practical from a design point of view. These design vari-
ables consist of the radius of the base (Rb ), the radius of the platform (Rp ),
the initial distance from the base to the platform (H) , the distance from the
platform-plane to the platforms center of mass (h), half of the angel between
the two neighboring joints on the base-plane (b ) and on the platform-plane
(p ). The initial length of the struts corresponding to the neutral configu-
ration of the GSP is denoted l0 . To simplify the equations, the following
parameters are defined

Rp H h l0
R = , H = , h = , l = , = b + p (2.1.1)
Rb Rb Rb Rb
To avoid singularity, the geometrical design space Gs is defined by (see
Figs. 2.1.1, 2.1.2)


 
6
Gs = (Rb , R , H , h , b , p ) R Rb > 0, R > 0, H > 0, b , p 0, <
3
(2.1.2)

17
g1 in paper
18 2 CHAPTER 2. PARAMETRIC KINEMATICS

c c c
Platform: m p , I xxp , I yyp , I zzp
z y
ep ep
x
c p , Op ep
h p1
P4 P3 P2
P1
Op ' p'1
1

P5 P6

l3 t l2 l1,0
H l4 l1

l5 l6
B3 B2

z
dmst ,1
y
eb eb s
b1 r1
2 in paper 1 B4
Ob x
eb
B5 Base B6 B1

Figure 2.1.1: Parametrization of a general GSP

eby
B2

B3
p
P2
P1

P3 B1
b
1200
P4
ebx
Rp B6
P6

P5
Rb
B4

B5

Figure 2.1.2: Top view of the general GSP


2.2. CONFIGURATION SPACES 19

2.2 Configuration spaces


The configuration of the platform can be expressed in Cartesian based or
joint based coordinates which are known as Cartesian-space and joint-space
representations, respectively. Cartesian-space coordinates qC is defined by
the three Cartesian translations as well as the three Bryant angles rotational
degrees of freedom as follows
 T
qC = tx ty tz (2.2.1)

Joint-space qJ is defined by the lengths of each strut li (i = 1, ..., 6) as


 T
qJ = l1 l2 l3 l4 l5 l6 (2.2.2)

The neutral configuration (initially symmetrical configuration) is expressed


in the Cartesian-space and joint-space as
 T
qC,0 = 0 0 H + h 0 0 0 (2.2.3)
and
 T
qJ,0 = l0 l0 l0 l0 l0 l0 (2.2.4)
respectively.

2.3 Kinematic equation and angular velocity


The base-frame and the platform-frame are two right-handed orthonormal
coordinate systems. The base-frame consists of eb = (exb , eyb , ezb ) as its basis
system and Ob as its origin. Similarly, the platform-frame consists of ep =
(exp , eyp , ezp ) as its basis system and Op as its origin. B1, B2, . . . , B6 are the
points representing the joints which are fixed to the base and P 1, P 2, . . . , P 6
are the points on the platform. In the neutral configuration, the z-axes of
the base-frame and the platform-frame are collinear which means that the
platforms center of mass cp is located on this common axis.
As shown in Fig. 2.1.1, one can write the following equation

l i = t + p i bi (2.3.1)
Accordingly, the following matrix representation of the kinematic equation is
proven in Paper II
20 CHAPTER 2. PARAMETRIC KINEMATICS

[li ]e = [t]e + [R]e [pi ]e [bi ]e (2.3.2)


b b b p b

The rotation tensor R can be represented in many ways as a combination


of rotations around three independent axes. In this paper Bryant angles are
chosen to formulate the kinematics for two reasons. First, for the evaluation
of the rotational eigenmodes, the Bryant angles have a clear physical meaning
such that for decoupled vibrations, they correspond to the three distinct rota-
tional eigenmodes. The decoupled vibrations are further studied in Paper II
and Paper III. Second, Bryant angles do not have the singularity problems at
the neutral configuration of GSPs as Euler angles would. The rotation matrix
based on Bryant angles is


c c c s s
[R]e = c s + s s c c c s s s s c (2.3.3)

b
s s c s c s c + c s s c c

where , and are consecutive rotations around x, y and z axes, respec-


tively, starting from the base-frame.
The matrix representation of the angular velocity of the platform in the
platform-frame is obtained according to

T
[]e = [R]Te []e = [R]Te ax [R] e [R]Te
n o 
= p,x p,y p,z
p b b b b b
(2.3.4)
where the components of the angular velocity are obtained in Paper II as
follows

p,x = sin + coscos









p,y = cos cossin (2.3.5)




p,z = + sin

2.4 Jacobian matrix


The Jacobian matrix is defined by the linear approximation of the transfor-
mation between the Cartesian and joint spaces as follows

4qJ = J4qC (2.4.1)


2.4. JACOBIAN MATRIX 21

where 4qJ = qJ qJ,0 and 4qC = qC qC,0 . If qJ = qJ (qC ), then the Jacobian
matrix is written

qJ
J= R66 (2.4.2)
qC qC,0
The matrix elements jik of the Jacobian matrix J are given by

1 (li li )
jik = =
2l0

qC,k q
C,0


T T T T
1 [t]eb [t]eb + 2 [t]eb [R]eb [pi ]ep 2 [t]eb [bi ]eb 2 [bi ]eb [R]eb [pi ]ep



2l0 qC,k


qC,0
(2.4.3)

where l0 is the initial length of the struts. Eq. (2.4.3) is derived in Paper II.
Then, according to the kinematic equation Eq. (2.3.2), li is obtained by

li = |li | = (2.4.4)
p
li li
Consequently, l0 is calculated as
1


2
l0 = Rb 1 + 2R + 2H 2R cos( ) (2.4.5)
3
Eventually, the Jacobian matrix is obtained with the following structure

j11 j12 j13 j14 j15 j16
j21 j22
j13 j24 j25 j16

1 j31 j32 j13 j34 j35 j16
J= (2.4.6)

2l j j32 j13 j34 j35 j16

31
j21 j22 j13 j24 j25 j16
j11 j12 j13 j14 j15 j16
where it is proven that around the neutral configuration the Jacobian matrix
is bijective and differentiable, since

(Rb , R , H , h , b , p ) Gs det(J) > 0 (2.4.7)


See Paper II (Section 2.3) for more details including the parametrically for-
mulated elements of the Jacobian matrix.
22 CHAPTER 2. PARAMETRIC KINEMATICS
Chapter 3

Parametric equations of
motion

3.1 Design variables


The design space Ds of a general GSP, at its neutral configuration, is defined
by its design variables as follows

, Izz , ki , ci R34
n 
Ds = cp
d, mst,i , Ist,i , mp , Ixx cp cp
, Iyy

o
B
d Gs , mst,i , Ist,i cp
, mp , Ixx , Izz , k, c > 0
cp cp
, Iyy (3.1.1)
In the case of identical struts, the design space is reduced to

, Izz , k, c R14
n 
Ds = cp
d, mst , Ist , mp , Ixx cp cp
, Iyy

o
B
d Gs , mst , Ist cp
, mp , Ixx , Izz , k, c > 0
cp cp
, Iyy (3.1.2)
where the geometrical design space Gs is defined in Eq. (2.1.2).

3.2 Equations of motion


In this work, the equations of motion are formulated using Lagrangian dy-
namics. Having parametrically obtained the elastic potential energy V and
the kinetic energy T , the Lagrangian is calculated as

L=T V (3.2.1)

23
24 CHAPTER 3. PARAMETRIC EQUATIONS OF MOTION

ki

Bi Pi

ci
Figure 3.2.1: Stiffness and damping components of each strut i = 1,...,6

Consequently, having calculated the dissipation energy D as well, the para-


metric non-linear equations of motion is obtained by

d L L D
 
+ = 0, i = 1, .., 6 (3.2.2)
dt qi qi qi
The elastic potential energy V is given by the stiffness of the struts (ki
R+ , i = 1, .., 6, see Fig. 3.2.1) as follows
6
1X
V = ki 4li2 (3.2.3)
2 i=1
where 4li = li l0 , l0 is given in Eq.(2.4.5) and li is obtained using the
Jacobian matrix developed in Paper II.
The dissipation energy D is based on a viscous damping assumption, see
Fig. 3.2.1, as follows
6
1X
D= ci l2 (3.2.4)
2 i=1 i
The total kinetic energy of the system is T = Tp + Tst . The kinetic energy
of the platform Tp is given by
1 1
Tp = I cp + t tm (3.2.5)
2 2
and then it is written in terms of the design variables of the system as
1 2
tx + t2y + t2z m+

Tp (tx , ty , tz , , , , tx , ty , tz , , , ) =
2
1 cp 2 1 cp 2 1
( + sin)2
  
Ixx sin + coscos + Iyy cos cossin + Izz cp
2 2 2
(3.2.6)
3.3. LINEARIZED EQUATIONS OF MOTION 25

However, the formulation of the kinetic energy of the struts


6  
Tst = (3.2.7)
X
Tst,i |T otal
i=1

is much more complicated, which is parametrically developed in Paper III.

3.3 Linearized equations of motion


The linearized equations of motion which lead to the inertia matrix (mass
matrix), stiffness matrix and damping matrix are obtained using

2 Tp


M C,p = q2
C q
C,0 ,qC,0









2 Tst
M C,st = q2




C q

C,0 ,qC,0


(3.3.1)

2V
KC = q2





C q

C,0






2D
CC = q2



C qC,0 ,qC,0

In the following, only the obtained structures of these matrices are pre-
sented:

The equivalent inertia matrix including both the platform and the struts:

MC = M C,p + M C,st =

0 0 0 0

M Mx

M 0 My 0 0

M 0 0 0
(3.3.2)

M 0 0


0

sym. M
M
26 CHAPTER 3. PARAMETRIC EQUATIONS OF MOTION

The stiffness matrix:

0 0 0 kx 0

kxx

kyy 0 ky 0 0

kzz 0 0 0
KC = (3.3.3)

k 0 0


0

sym. k
k

The damping matrix:

0 0 0 0

cxx cx

cyy 0 cy 0 0

czz 0 0 0
CC = (3.3.4)

c 0 0


0

sym. c
c
These matrices are in the Cartesian-space. The components of these matrices
are parametrically formulated in Paper II and Paper III.
Chapter 4

Parametric modal analysis

4.1 Eigenvectors
To parametrically obtain the damped eigenfrequencies of the system, first
the eigenvectors are formulated using the undamped eigenvalue equations
MC 1 KC i2 I66 Xi = 061 (where I66 is the 6 6 identity matrix and
i are the undamped eigenfrequencies).
 Parametrically obtained
 eigenvectors
Xi form the modal matrix X = X1 X2 X3 X4 X5 X6 as follows

0 0 x,1 0 0

x,2
0 0 0 y,1 0 y,2

0 1 0 0 0 0
X= (4.1.1)

0 0 0 1 0 1

0 0 1 0 1 0

1 0 0 0 0 0
The six modes presented in Eq. (4.1.1) are the torsional, axial and four
tilting-lateral modes, respectively. Fig. 4.1.1 shows the mode shapes of GSPs.
In Paper II and Paper III, the eigenvectors are studied in detail where the
parametric formulations of each matrix component are also given.

4.2 Eigenfrequencies
Considering weak damping and by using a diagonalization technique (which
is presented in paper II), the damped eigenfrequencies are obtained as
s
2 4
d = 2 (4.2.1)
4

27
28 CHAPTER 4. PARAMETRIC MODAL ANALYSIS

(a) Torsional mode (- (b) Axial mode (z-mode) (c) Tilting-Lateral mode 1
mode) (x, 1-mode)

(d) Tilting-Lateral mode 2 (e) Tilting-Lateral mode 3 (f) Tilting-Lateral mode 4


(y, 1-mode) (x, 2-mode) (y, 2-mode)

Figure 4.1.1: Six mode shapes of GSPs


4.3. DECOUPLED VIBRATIONS 29

where the six undamped eigenfrequencies corresponding to the eigenvectors


in Eq. (4.1.1) are formulated as
q q
k


= M , z = kzz
M




r
M k +M kxx 2Mx kx x

x,1 =



2M M 2Mx
2








r
M k +M kyy 2My ky y

y,1 =

2M M 2My2 (4.2.2)





r
M k +M kxx 2Mx kx +x

x,2 =


2M M 2Mx2









r
y,2 = M k +M kyy 2My ky +y



2M M 2M 2
y

See Paper II (Section 2.3) and Paper III (Section 2.2) for further details.

4.3 Decoupled vibrations


The decoupled vibrations are defined by the existence of a fully decoupled
modal matrix. This means six pure modes corresponding to each one of
the Cartesian-space coordinates. To obtain such a decoupled modal matrix,
eigenvector components in the modal matrix given in Eq. (4.1.1) have to be
either zero or infinite. It means that when both x,1 and y,1 in the modal
matrix (corresponding modes are orthonormal) are infinitely large, then both
x,2 and y,2 are required to have a zero value and vice versa. Hence, the
necessary condition in order to achieve a fully decoupled modal matrix is (see
Paper III for the proof)

kxx kyy kx ky
= = = , Mx = My 6= 0 (4.3.1)
M M Mx My
that corresponds to a unique value of the strut inertia

B = l2 k l2 k
Ist 0 x
m = 6(ky +k
6(kx +kxx Rb h ) p
0 y
yy Rb h )
mp , h 6= 0 (4.3.2)

In Paper III, we also prove that the condition specified in Eq. (4.3.1) never
leads to a fully decoupled modal matrix unless the multiple eigenfrequencies
condition is satisfied. However, in the case that the inertia matrix is diagonal
30 CHAPTER 4. PARAMETRIC MODAL ANALYSIS

(Mx = My = 0), the fully decoupled vibrations are obtained if and only if
the stiffness matrix is diagonal (kx = ky = 0). A detailed analytical study
for each case is carried out in Paper II (Section 2.4) and Paper III (Section
2.3).
The analytical study presented in this work on the decoupled vibrations
can be employed in the design of high-precision GSPs in order to remove
the crosstalk between different degrees of freedom, and hence, improve the
precision of the system.
Chapter 5

Numerical evaluation

5.1 Characteristics of the reference GSP

In this work, with the aid of a reference GSP, the developed analytical model
is numerically evaluated. Selected data for the reference GSP are chosen to
be relevant for a collimation system of a large optical telescope, see Table 5.1.

Table 5.1: Characteristics of the reference GSP

Variable Description Value Unit


b Half-angel of the joints on the base 10 180

rad
p Half-angel of the joints on the platform 15 180

rad
k Axial stiffness of the struts 106 N/m
c Axial damping of the struts 103 Ns/m
mp Mass of the platform 102 kg
kgm2
c
Ixxp Moment of inertia of the platform around x 5, 7
kgm2
c
Iyyp Moment of inertia of the platform around y 5
kgm2
c
Izzp Moment of inertia of the platform around z 10
Rb Radius of the base joints circle 0.5 m
R Radius ratio (R = Rp/Rb ) 0.8 -
H Height ratio (H = H/Rb ) 1.8 -
h Height ratio of the center of mass 0.2 -
B
Ist Strut moment of inertia about base joints 0, 10
6 , 10 kgm2

31
32 CHAPTER 5. NUMERICAL EVALUATION

5.2 FEM simulation


The aim is to compare and cross-check the analytically obtained eigenfrequen-
cies and eigenvectors with the corresponding eigenfrequencies and eigenvectors
obtained by a totally different method (in this case an FEM simulation). The
reference GSP modeled in the FEM simulation has exactly the same charac-
teristics as in the analytical model. The FEM simulation is carried out using
ABAQUS . The comparison shows an excellent agreement between these
methods in terms of modal properties. For further details and the results of
this simulation see Paper II (Section 3).

5.3 Numerical results


The modal behavior of the reference GSP is systematically studied in Paper
II and Paper III. Figs. 4 and 5 of Paper III show the modal behavior of the
GSP versus geometrical variables. Here we additionally present Figs. 5.3.1
and 5.3.2 where the results concerning the modal behavior with respect to the
inertia, damping and stiffness are shown. In Fig. 5.3.1, all the six damped
eigenfrequencies are plotted according to Eq. 4.2.2 versus the inertia, damping
and stiffness variables for three different values of Ist
B . The first column of the

plots in Fig. 5.3.1 shows the vibrational behavior when there is no mass
assigned to the struts IstB = 0, while the second and third columns are for

Ist = 10/6 and Ist = 10, respectively.


B B

As it is expected, all the eigenfrequencies in Fig. 5.3.1 decrease with


c c
the increase of strut inertia. Although Ixxp 6= Iyyp , it is observed that nearly
everywhere the x, 1 and y, 1-modes are quite close in the plots in Fig. 5.3.1.
This is due to the fact that, according to Fig. 5.3.2, x, 1 and y, 1-modes
are mainly lateral dominant modes in the ranges shown in Fig. 5.3.1. In
lateral-dominant modes, the eigenfrequencies are not much sensitive to the
platforms moment of inertia. However, tilting-dominant modes are quite
c c
sensitive to the platforms moments of inertia. Hence, since Ixxp 6= Iyyp , as
expected x, 2 and y, 2-modes are clearly distinct in the plots.
The plots in 5.3.1 (a), ..., (d) show the modal behavior of the reference
GSP versus the inertia variables. Whereas, the plots in Fig 5.3.1 (e) show
the sensitivity of the six eigenfrequencies to the strut damping. The critical
damping is given by

2
ccr,i = k, i = 1, ..., 6 (5.3.1)
i
5.3. NUMERICAL RESULTS 33

and it shows that the higher the undamped eigenfrequencies, the less axial
damping lead to critical damping for the corresponding mode. Figure 5.3.1
(e) also illustrates that the sequence of the eigenmodes changes at every inter-
section point of the curves. Hence, it is important to note that the sequence
of the eigenfrequencies with respect to their values in damped and undamped
vibrations might be different.
The plots in Fig. 5.3.1 (f) show the damped eigenfrequencies with respect
to stiffness. Since the scale of stiffness is relatively large, the region which
corresponds to critical damping is not visible, in this region a similar type
of sequence change as observed in Fig. 5.3.1 (e) is present. Thus, for rel-
atively large stiffness values (when is approaching zero) the sequence of the
eigenfrequencies remains unchanged. This may also be expressed as
1 
|c=constant lim d,i |c=constant k (5.3.2)
k k

The corresponding eigenvectors of Fig. 5.3.1 are presented in Fig. 5.3.1,


where only the modal matrix components of the four coupled modes are
shown. In the plots in Fig. 5.3.1, if the poles of the eigenvector components
coincide with the zeros, a complete decoupled vibration is accomplished (a
pole corresponds to an infinity which is marked by and a zero corre-
sponds to a zero which is marked by o). In this example, it is shown that
for the given geometry complete decoupled vibrations are not feasible. How-
ever, in Paper II and Paper III we show how it is possible to obtain complete
decoupled vibrations by changing the geometry variables.
In Fig. 7 of Paper II, another important result is presented where it is
shown that by considering a constant , all the damped eigenfrequencies are
bound to an upper limit equal to

1 k
d = = (5.3.3)
c
Based on the obtained results, it is concluded that neglecting the strut
inertia may lead to unrealistic results. Therefore, it is crucial to consider the
strut inertia in order to obtain an accurate evaluation of GSPs.
34 CHAPTER 5. NUMERICAL EVALUATION

y,2-mode x,2-mode y,1-mode x,1-mode z-mode -mode

(a-1) (a-2) (a-3)


300 300 300

200 200 200


d

d
100 100 100

0 0 0
0 100 200 300 400 500 0 100 200 300 400 500 0 100 200 300 400 500
m (kg) m (kg) m (kg)
p p p
(b-1) (b-2) (b-3)
300 300 300

200 200 200


d

d
100 100 100

0 0 0
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
Icp
2
Icp
2
Icp (kgm )
2
(kgm ) (kgm )
xx xx xx
(c-1) (c-2) (c-3)
300 300 300

200 200 200


d

d
100 100 100

0 0 0
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
Icyyp
2
Icyyp
2
(kgm ) (kgm ) Icyyp (kgm2)
(d-1) (d-2) (d-3)
300 300 300

200 200 200


d

100 100 100

0 0 0
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
Iczzp (kgm2) Iczzp (kgm2) Iczzp (kgm2)
(e-1) (e-2) (e-3)
300 300 300

200 200 200


d

100 100 100

0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
c (Ns/m) 4
x 10 c (Ns/m) 4
x 10 c (Ns/m) 4
x 10
4 4 4
x 10 (f-1) x 10 (f-2) x 10 (f-3)
3 3 3

2 2 2
d

1 1 1

0 0 0
0 5 10 0 5 10 0 5 10
k (N/m) 9
x 10 k (N/m) 9
x 10 k (N/m) 9
x 10

Figure 5.3.1: The six damped eigenfrequencies of the reference GSP versus
the inertia, damping and stiffness variables; (a), (b), (c), (d), (e) and (f) show
c c c
the influence of mp , Ixxp , Iyyp , Iyyp , c and k respectively; all the variables are
studied for three different values of strut moment of inertia which are denoted
by: (1) Ist
B = 0, (2) I B = 10/6, (3) I B = 10.
st st
5.3. NUMERICAL RESULTS 35

x,2 y,2 y,1 x,1

(a-1) (a-2) (a-3)


100 100 100

0 0 0


-100 -100 -100
0 50 100 150 0 50 100 150 0 50 100 150
mp (kg) mp (kg) mp (kg)
(b-1) (b-2) (b-3)
100 100 100

0 0 0


-100 -100 -100
0 100 200 300 400 500 0 100 200 300 400 500 0 100 200 300 400 500
Icxxp (kgm2) Icxxp (kgm2) Icxxp (kgm2)
(c-1) (c-2) (c-3)
100 100 100

0 0 0


-100 -100 -100
0 100 200 300 400 500 0 100 200 300 400 500 0 100 200 300 400 500
Icp (kgm2) Icp (kgm2) Icp (kgm2)
yy yy yy
(d-1) (d-2) (d-3)
100 100 100

0 0 0

-100 -100 -100


0 100 200 300 400 500 0 100 200 300 400 500 0 100 200 300 400 500
Iczzp (kgm ) 2
Iczzp (kgm ) 2
Iczzp (kgm2)
(e-1) (e-2) (e-3)
100 100 100

0 0 0

-100 -100 -100


0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
c (Ns/m) 4
x 10 c (Ns/m) x 10
4 c (Ns/m) 4
x 10
(f-1) (f-2) (f-3)
100 100 100

0 0 0

-100 -100 -100


0 5 10 0 5 10 0 5 10
k (N/m) 9
x 10 k (N/m) x 10
9 k (N/m) 9
x 10

Figure 5.3.2: The components of the corresponding eigenvectors according to


Figure 6.3.1; all the variables are plotted for three different strut moment of
inertia which are denoted by: (1) Ist B = 0, (2) I B = 10/6, (3) I B = 10;
st st
shows the poles which leads to an infinity and o shows the zeros which
leads to a zero .
36 CHAPTER 5. NUMERICAL EVALUATION
Chapter 6

Summary and future work

6.1 Summary and conclusions


In this dissertation, a comprehensive parametric study of the modal behavior
of GSPs is carried out. It should be noted that in the literature of GSPs
neither a parametric study on the damped vibrations nor the consideration
of strut inertia has yet been addressed. To fill in this gap, we have developed
a fully parametric model in two stages which are the subjects of Paper II and
Paper III. In the following, the main results of the present dissertation are
listed:

This work establishes a comprehensive parametric model of the damped


vibrations in order to be used directly in analysis, optimization and
control of GSPs.

The kinematics, equations of motion, linearized equations of motion and


modal analysis of GSPs are parametrically developed.

The stiffness, damping and inertia matrices are parametrically formu-


lated.

The equivalent inertia matrix considering the effects of both the plat-
form and the struts is also formulated where it is proven that the inertia
matrix is not generally a diagonal matrix.

The eigenvectors and the damped eigenfrequencies are also paramet-


rically developed where it is found that there exist four tilting-lateral
modes.

37
38 CHAPTER 6. SUMMARY AND FUTURE WORK

The necessary and sufficient conditions for the decoupled vibrations are
analytically studied and formulated. In the case of a diagonal inertia
matrix, it is concluded that the decoupled vibrations occur if and only
if the stiffness matrix is diagonal. However, for a non-diagonal inertia
matrix, different conditions need to be satisfied which are formulated.

In this work, much effort has been devoted to obtain well-structured


expressions for all the parametrically formulated equations.

To demonstrate the analytical model developed in this work, the vi-


brational behavior of a reference GSP is systematically studied. An
FEM simulation is also carried out in order to cross-check the results
where an excellent agreement between the analytical and FEM results
is obtained.

6.2 Future work


In the future work, the intention is to develop an analytical optimization
method in order to achieve optimal dynamical performance at the neutral
configuration. Furthermore, a study of the joint-space formulation of the
GSP vibrations will be undertaken.
Modal properties at asymmetric configurations are of interest for the op-
timized GSP. This will require an investigation of the workspace influence on
spectral properties. Vibrational performance may be also be influenced by the
joints. In particular, for spherical and universal joints, evaluation of friction
is of importance. However, consideration of these details of the system makes
it inevitable to solve the eigenvalue problem numerically.
The possibility of developing new GSP applications, by the employment
of the mathematical model presented in this dissertation, may also be con-
sidered. Applications in which the struts have additional localized damping
elements are of particular interest. Hence, the focus could be on vibration-
isolation related areas.
Bibliography

[1] D. Stewart, A platform with six degrees of freedom, Proceedings of the


institution of mechanical engineers 180 (15) (1965) 371386.

[2] V. Gough, S. Whitehall, Universal tyre test machine, Proc. FISITA 9th
Int. Technical Congress.

[3] J.-P. Merlet, Parallel Robots, Vol. 208, 2006. doi:10.1007/1-4020-4133-0.

[4] Y. D. Patel, Parallel Manipulators Applications-A Survey, Modern Me-


chanical Engineering 02 (03) (2012) 5764. doi:10.4236/mme.2012.23008.

[5] Y. Ting, Y.-S. Chen, H.-C. Jar, Modeling and control for a Gough-
Stewart platform CNC machine, Journal of Robotic Systems 21 (11)
(2004) 609623. doi:10.1002/rob.20039.

[6] J. J. Zierer, J. H. Beno, D. a. Weeks, I. M. Soukup, J. M. Good, J. a.


Booth, G. J. Hill, M. D. Rafal, Design, testing, and installation of a high-
precision hexapod for the Hobby-Eberly Telescope dark energy experi-
ment (HETDEX), in: L. M. Stepp, R. Gilmozzi, H. J. Hall (Eds.), Proc.
of SPIE, Vol. 8444, 2012, pp. 84444O84444O12. doi:10.1117/12.926394.

[7] M. Wapler, V. Urban, T. Weisener, J. Stallkamp, M. Drr,


A. Hiller, A Stewart platform for precision surgery, Transactions of
the Institute of Measurement and Control 25 (4) (2003) 329334.
doi:10.1191/0142331203tm092oa.

[8] O. Stroosma, M. M. van Paassen, M. Mulder, Using the simona research


simulator for human-machine interaction research, in: AiAA Modeling
and Simulation Technologies Conference and Exhibit, 2003, pp. 5525
5525. doi:doi:10.2514/6.2003-5525.

[9] P. Schipani, S. DOrsi, D. Fierro, L. Marty, Active optics control of VST


telescope secondary mirror., Applied optics 49 (16) (2010) 3199207.

39
40 BIBLIOGRAPHY

[10] K. Mller, H. Marth, P. Pertsch, Piezo-Based, Long-Travel Actuators For


Special Environmental Conditions, PI Proceedings, Actuator 49 (721)
(2006) 15.

[11] K. Liu, F. Lewis, G. Lebret, D. Taylor, The singularities and dynam-


ics of a Stewart platform manipulator, Journal of Intelligent & Robotic
Systems 8 (3) (1993) 287308. doi:10.1007/BF01257946.

[12] S. Bandyopadhyay, A. Ghosal, An algebraic formulation of kine-


matic isotropy and design of isotropic 6-6 Stewart platform ma-
nipulators, Mechanism and Machine Theory 43 (5) (2008) 591616.
doi:10.1016/j.mechmachtheory.2007.05.003.

[13] B. M. St-Onge, Singularity Analysis and Representation of the Gen-


eral Gough-Stewart Platform, The International Journal of Robotics Re-
search 19 (3) (2000) 271288. doi:10.1177/02783640022066860.

[14] Z. Geng, L. S. Haynes, J. D. Lee, R. L. Carroll, On the dynamic


model and kinematic analysis of a class of Stewart platforms, Robotics
and Autonomous Systems 9 (4) (1992) 237254. doi:10.1016/0921-
8890(92)90041-V.

[15] J. Lee, Z. Geng, A dynamic model of a flexible stewart platform, Comput-


ers & Structures 48 (3) (1993) 367374. doi:10.1016/0045-7949(93)90313-
3.

[16] H. Abdellatif, B. Heimann, Computational efficient inverse dynam-


ics of 6-DOF fully parallel manipulators by using the Lagrangian
formalism, Mechanism and Machine Theory 44 (1) (2009) 192207.
doi:10.1016/j.mechmachtheory.2008.02.003.

[17] S. Pedrammehr, M. Mahboubkhah, N. Khani, Improved dynamic equa-


tions for the generally configured Stewart platform manipulator, Jour-
nal of Mechanical Science and Technology 26 (3) (2012) 711721.
doi:10.1007/s12206-011-1231-0.

[18] B. Dasgupta, T. Mruthyunjaya, Closed-Form Dynamic Equations of the


General Stewart Platform through the Newton-Euler Approach, Mecha-
nism and Machine Theory 33 (7) (1998) 9931012. doi:10.1016/S0094-
114X(97)00087-6.

[19] B. Dasgupta, T. Mruthyunjaya, A Newton-Euler formulation for the


inverse dynamics of the Stewart platform manipulator, Mechanism
BIBLIOGRAPHY 41

and Machine Theory 33 (8) (1998) 11351152. doi:10.1016/S0094-


114X(97)00118-3.

[20] Y. Cheng, G. Ren, S. Dai, The multi-body system modelling of the


Gough-Stewart platform for vibration control, Journal of Sound and Vi-
bration 271 (3-5) (2004) 599614. doi:10.1016/S0022-460X(03)00283-9.

[21] C. Gosselin, Stiffness mapping for parallel manipulators, Robotics and


Automation, IEEE Transactions on 6 (3) (1990) 377382.

[22] B. El-Khasawneh, P. Ferreira, Computation of stiffness and stiffness


bounds for parallel link manipulators, International Journal of Machine
Tools & Manufacture 39 (1999) 321342.

[23] S. Muruganandam, S. Pugazhenthi, Stiffness-based workspace atlas of


hexapod machine tool for optimal work piece location, The International
Journal of Advanced Manufacturing Technology 42 (1-2) (2008) 202210.
doi:10.1007/s00170-008-1573-7.

[24] C. Huang, W.-h. Hung, I. Kao, New conservative stiffness mapping for
the Stewart-Gough platform, in: Proceedings 2002 IEEE International
Conference on Robotics and Automation (Cat. No.02CH37292), Vol. 1,
Ieee, 2002, pp. 823828. doi:10.1109/ROBOT.2002.1013459.

[25] J. Chen, F. Lan, Instantaneous stiffness analysis and simulation for hexa-
pod machines, Simulation Modelling Practice and Theory 16 (4) (2008)
419428. doi:10.1016/j.simpat.2008.01.003.

[26] V. Portman, V. Chapsky, Y. Shneor, Workspace of parallel kinemat-


ics machines with minimum stiffness limits: Collinear stiffness value
based approach, Mechanism and Machine Theory 49 (2012) 6786.
doi:10.1016/j.mechmachtheory.2011.11.002.

[27] J. M. Prajapati, L. N. Patel, Dynamic (Forward and Inverse Force Trans-


mission Capability) Analysis of the Stewart Platform as Robot Manip-
ulator, 2007 International Conference on Mechatronics and Automation
(2007) 28482853doi:10.1109/ICMA.2007.4304011.

[28] H.-Z. Jiang, J.-F. He, Z.-Z. Tong, Characteristics analysis of joint
space inverse mass matrix for the optimal design of a 6-DOF parallel
manipulator, Mechanism and Machine Theory 45 (5) (2010) 722739.
doi:10.1016/j.mechmachtheory.2009.12.003.
42 BIBLIOGRAPHY

[29] J.-F. He, H.-Z. Jiang, Z.-Z. Tong, B.-P. Li, J.-W. Han, Study on dy-
namic isotropy of a class of symmetric spatial parallel mechanisms with
actuation redundancy, Journal of Vibration and Control 18 (8) (2011)
11561164. doi:10.1177/1077546311409608.

[30] a. Preumont, M. Horodinca, I. Romanescu, B. de Marneffe, M. Avraam,


A. Deraemaeker, F. Bossens, A. Abu Hanieh, A six-axis single-stage ac-
tive vibration isolator based on Stewart platform, Journal of Sound and
Vibration 300 (3-5) (2007) 644661. doi:10.1016/j.jsv.2006.07.050.

[31] S. Pedrammehr, M. Mahboubkhah, N. Khani, A study on vibration


of Stewart platform-based machine tool table, The International Jour-
nal of Advanced Manufacturing Technology 65 (5-8) (2012) 9911007.
doi:10.1007/s00170-012-4234-9.

[32] Z. Tong, J. He, H. Jiang, G. Duan, Optimal design of a class of gen-


eralized symmetric Gough-Stewart parallel manipulators with dynamic
isotropy and singularity-free workspace, Robotica 30 (02) (2011) 305
314. doi:10.1017/S0263574711000531.

[33] H. Abdellatif, B. Heimann, Computational efficient inverse dynam-


ics of 6-DOF fully parallel manipulators by using the Lagrangian
formalism, Mechanism and Machine Theory 44 (1) (2009) 192207.
doi:10.1016/j.mechmachtheory.2008.02.003.

[34] K. Harib, K. Srinivasan, Kinematic and dynamic analysis of Stewart


platform-based machine tool structures, Robotica 21 (5) (2003) 541554.
doi:10.1017/S0263574703005046.

[35] T. Yang, J. Ma, Z.-g. Hou, M. Tan, Robust backstepping control of


active vibration isolation using a stewart platform, in: 2009 IEEE Inter-
national Conference on Robotics and Automation, Ieee, 2009, pp. 1788
1793. doi:10.1109/ROBOT.2009.5152486.

[36] H. Lin, J. E. McInroy, Disturbance attenuation in precise hexapod point-


ing using positive force feedback, Control Engineering Practice 14 (11)
(2006) 13771386. doi:10.1016/j.conengprac.2005.10.002.

[37] Y. Chen, J. McInroy, Decoupled control of flexure-jointed hexapods using


estimated joint-space mass-inertia matrix, Control Systems Technology,
IEEE 12 (3) (2004) 413421.
BIBLIOGRAPHY 43

[38] J. McInroy, J. Hamann, Design and control of flexure jointed hexapods,


IEEE Transactions on Robotics and Automation 16 (4) (2000) 372381.
doi:10.1109/70.864229.

[39] L. Liu, G. Zheng, W. Huang, Octo-strut vibration isolation platform and


its application to whole spacecraft vibration isolation, Journal of Sound
and Vibration 289 (4-5) (2006) 726744. doi:10.1016/j.jsv.2005.02.040.

[40] H. Z. Jiang, J. F. He, Z. Z. Tong, W. Wang, Dynamic isotropic de-


sign for modified Gough-Stewart platforms lying on a pair of circular
hyperboloids, Mechanism and Machine Theory 46 (2011) 13011315.
doi:10.1016/j.mechmachtheory.2011.04.003.

[41] M. Olivier, G. Duchini, Hexapod Mechanical Assembly modal survey


test, in: 5th International Symposium on Environmental Testing for
Space Programmes, Noordwijk, The Netherlands, 2004, pp. 165172.

[42] H. Jiang, J. He, Z. Tong, Modal Space Control for a Hydraulically


Driven Stewart Platform, Journal of Control Engineering and Technology
2 (July) (2012) 106115.

[43] M. Mahboubkhah, M. J. Nategh, S. Esmaeilzadeh Khadem, A compre-


hensive study on the free vibration of machine tools hexapod table, The
International Journal of Advanced Manufacturing Technology 40 (11-12)
(2008) 12391251. doi:10.1007/s00170-008-1433-5.

[44] P. Mukherjee, B. Dasgupta, A. Mallik, Dynamic stability index and vi-


bration analysis of a flexible Stewart platform, Journal of Sound and
Vibration 307 (3-5) (2007) 495512. doi:10.1016/j.jsv.2007.05.036.

[45] J. Selig, X. Ding, Theory of vibrations in Stewart platforms, in: Interna-


tional Conference on Intelligent Robots and Systems. Proceedings. IEEE,
2001, pp. 21902195.

[46] M. Mahboubkhah, M. J. Nategh, S. Esmaeilzade Khadem, Vibration


analysis of machine tools hexapod table, The International Journal
of Advanced Manufacturing Technology 38 (11-12) (2007) 12361243.
doi:10.1007/s00170-007-1183-9.

[47] W. Wang, H.-y. Yang, J. Zou, X.-d. Ruan, X. Fu, Optimal design of
Stewart platforms based on expanding the control bandwidth while con-
sidering the hydraulic system design, Journal of Zhejiang University SCI-
ENCE A 10 (1) (2009) 2230. doi:10.1631/jzus.A0820329.
44 BIBLIOGRAPHY

[48] T. Andersen, A. Enmark, Integrated Modeling of Telescopes, Springer


New York, 2011.

[49] P. Y. Bely, The Design and Construction of Large Optical Telescopes,


2003.

[50] F. Roddier, M. Northcott, J. E. Graves, A simple low-order adaptive


optics system for near-infrared applications (1991). doi:10.1086/132802.

[51] P. Schipani, L. Marty, Stewart platform kinematics and secondary mirror


aberration control, Optomechanical Technologies for Astronomy.

[52] J. P. Merlet, Parallel Robots, Parallel Robots, Springer, 2006.


Paper I Paper I

Behrouz Afzali-Far, Per Lidstrm, Kristina Nilsson and Arne Ardeberg


Analytical Stiffness Optimization of High-Precision Hexapods for
Large Optical Telescope Applications
Proceedings of the 25th Nordic Seminar on Computational Mechanics, 2012

45
Proceedings of the 25th Nordic Seminar on Computational Mechanics
K. Persson, J.Revstedt, G. Sandberg, M. Wallin (Eds.)
Lund University, 2012

Analytical Stiffness Optimization of High-Precision


Hexapods for Large Optical Telescope Applications

Behrouz Afzali Far1,2, Per Lidstrm1, Kristina Nilsson1, and Arne Ardeberg2

(1) Division of Mechanics, Lund University, Lund, Sweden,


Behrouz.Afzali_Far@mek.lth.se, Per.Lidstrom@mek.lth.se,
Kristina.Nilsson@mek.lth.se
(2) Lund Telescope Group AB, Trollebergsvgen 5, Lund, Sweden,
Arne.Ardeberg@ltgab.se

Summary. An analytical stiffness and eigenfrequency model of symmetric parallel 6-6 Stewart
platforms (hexapods) is developed based on geometrical design variables to optimize the
dynamical performance. The model is based upon Lagrangean dynamics in which the Bryant
angles are used for the kinematics formulation. With the analytical eigenfrequency model,
optimum stiffness characteristics can be obtained for any industrial application with limited
workspace such as optical collimation systems. The actuator length-flexibility dependency is
also considered in the analytical model. It is proposed that to increase the actuation bandwidth
in six degrees of freedom, an eigenfrequency cost function can be defined and optimized.

Key words: Gough Stewart platform, hexapod, analytical, stiffness, eigenfrequency, telescope

Introduction

Cassegrain-type [1] large optical telescopes for astronomy have two mirrors; the secondary
mirror (M2) is the optical element which reflects the light collected by the primary mirror (M1)
to the scientific instrument (figure 1). Active collimation of M2 corrects image-quality defects
due to structure deformations caused by gravity and thermal variations. Due to the tracking
velocity of optical telescopes, the bandwidth required is approximately 0.01 Hz [2]. The image-
quality of a stable telescope with a well-figured M1 surface and well-collimated M2 is limited
by the atmospheric disturbances. Further improvement of the image-quality requires correction
of at least first-order defects due to the atmospheric disturbances as well as errors caused by
telescope structural vibrations. Such first-order defects of atmospheric disturbances can be
compensated by tilting M2 with a sufficient high bandwidth. In practice, tilting is responsible
for a major part of the image degradation caused by the atmosphere. Given the importance of
tilting, major improvements of observed image-quality can be achieved through two-axes
rotations (tip-tilt), first-order compensatory movements of M2 [3]. Successful operation requires
a very stiff yet light-weight system including the active mechanism, M2 and its support
structure. The tip-tilt movements are controlled with a bandwidth of at least 10 Hz, in closed-
loop mode via an optical sensor. The corresponding angular range should be several arcsec. The
system will, in addition to effects of atmospheric disturbances, correct for image degradation
due to structural vibrations caused by wind buffeting. Mainly due to precision and stiffness
requirements for the collimation system of M2, high-precision hexapods can be used in large
telescopes [4]. For further improvements of image-quality, a stiff high-bandwidth hexapod can
provide three additional useful degrees of freedom (decentering in two axes and focusing)
compared with only using tip-tilt.

SSecondary mirror

Primary mirror

Scientific Instrument

Figure 1. Model of a Cassegrain-type large optical telescope

Hexapods have attracted the attention of researchers and designers since the time it was first
introduced by Stewart in 1965. A hexapod is the minimum arrangement of a parallel robot to
provide six degrees of freedom. Hexapods are also known as Stewart platforms or Gough-
Stewart platforms which have a wide range of precision engineering applications such as in
machining technology, astronomical telescopes and surgical robots [5]. Parallel robots in
general and hexapods in particular have some advantages over serial robots. Among these
advantages, for the large optical telescope application which is the interest of this paper,
precision and stiffness are of great importance. Much research has been carried out on the
kinematics of hexapods, but few on stiffness [6,7]. The stiffness matrix can be obtained using
Jacobian matrix methods or as a product of force and deformation related matrices [8]. The
stiffness matrix can also be assembled from the linearization of the equations of motion. There
are some studies about the workspace evaluation based on stiffness [9]. H. Jiang [10,11]
calculates the eigenfrequencies and studies geometric conditions leading to equal
eigenfrequencies, however, the length-flexibility dependency and optimum geometries using an
eigenfrequency cost function remain to be evaluated.
This abstract introduces an analytical eigenfrequency model for hexapods, based on the
geometrical design variables to optimize the geometry restricted by eigenmodal requirements.
Geometrical variables of hexapods are important, especially in the telescope applications where
the workspace is quite limited. Therefore, to increase the actuation bandwidth and precision for
all six degrees of freedom, which are limited by the eigenfrequencies of any system,
optimization of the hexapod geometry is a key objective.
Stiffness and eigenfrequency Modeling

In the present study, formulation of the stiffness model of hexapods starts with solving the
reverse kinematics of the system. Reverse kinematics means obtaining the length of each
actuator with a given configuration of the platform (figure 2). The kinematics is formulated
using a general geometry of symmetric hexapods. There are totally 12 nodes on the planar base
(figure 2) and the planar platform where the actuator ends are attached with spherical joints. All
the joint nodes of each platform are located on a circle with an arrangement leading to three
pairs of actuators with 120 degrees axial symmetry in the neutral configuration. The base and
the platform are considered rigid but the actuators are flexible elements without mass. The
geometrical design variables are; the ratio of the base and platform radii, the ratio of the height
of the hexapod over the base radius, and the angles between the nearest two nodes on each
platform.

Figure 2. A general symmetric 6-6 Stewart platform

By assuming small angles, which is the case in vibrations, the choice of Bryant angles for
the rotation matrix leads to a very simple formulation for the angular velocity and it is shown
that these angles are decoupled for all six eigenmodes of symmetric hexapods. In addition to the
simplicity of equations, the nonsingularity for small angles is an advantage of the Bryant angles.
The equations of motion of the system are solved with the Lagrangean dynamics approach.
Stiffness modeling is based on two generalized coordinate systems, Cartesian and length
coordinates. Analytical Stiffness and mass matrices for both coordinate systems are developed.
Furthermore, the analytical eigenfrequencies of the system as well as the eigenmodes are
formulated based on the geometric design variables.
To avoid optimum results on the boundary of the design variables domain and to consider
stiffness practicalities, the flexibility of each actuator is assumed to be proportional to its length
in the neutral configuration. This realistic approach leads to analytical optimum values for each
design variable based on the requirements of the system dynamical performance.
To reduce the system of six equations for each eigenfrequency, an eigenfrequency cost
function is defined which is driven mainly by the minimum eigenfrequency. The method
presented is an analytical approach to design an optimal hexapod with the maximum dynamical
performance.
Conclusion

In the present work, an analytical eigenfrequency model of a general symmetric hexapod is


formulated based on geometrical design variables for optimizing stiffness of limited workspace
applications such as in large optical telescopes. High-precision hexapods can be used as an
active mechanism for large telescopes to collimate mirrors against gravity and temperature in
low bandwidth, and to improve the image-quality by compensating for structural vibrations and
atmospheric disturbances in high bandwidth. The stiffness model is based on the equations of
motion, using Lagrangean dynamics. To avoid singularity and to have simpler eigenfrequency
equations, Bryant angles have been chosen for the generalized kinematic modeling. It is
suggested that to increase the actuation bandwidth of such a system, an eigenfrequency cost
function can be defined and optimized to avoid resonances in six degrees of freedom. It is also
concluded that, considering the actuators length-flexibility dependency, more realistic optimum
results can be obtained.

References

[1] Torben Andersen, Anita Enmark, Integrated Modeling of Telescopes, Springer


New York, 2011.
[2] Pierre Y. Bely, The Design and Construction of Large Optical Telescopes,
Springer, 2003
[3] Roddier, Francois; Northcott, Malcolm; Graves, J. Elon, A simple low-order
adaptive optics system for near-infrared applications, Astronomical Society of the
Pacific, 103:131-149, 1991.
[4] P. Schipani, L. Marty, Stewart platform kinematics and secondary mirror
aberration control, Optomechanical Technologies for Astronomy, 62733B, 2006.
[5] Y. D. Patel, P. M. George, Parallel Manipulators Applications-A Survey, Modern
Mechanical Engineering, 1(1): 57-64, 2012.
[6] Bhaskar Dasguptaa, T.S. Mruthyunjaya, The Stewart platform manipulator: a
review, Mechanism and Machine Theory, 35:15-40, 2000.
[7] J. Chen, F. Lan, Instantaneous stiffness analysis and simulation for hexapod
machines, Simulation Modelling Practice and Theory, 16:419-428, 2008.
[8] Jokin Aginaga, Isidro Zabalza, Oscar Altuzarra, Jasiel Njera, Improving static
stiffness of the parallel manipulator using inverse singularities, Robotics and
Computer-Integrated Manufacturing, 28(4):458-471, 2012.
[9] V.T. Portman, V.S. Chapsky, Y. Shneor, Workspace of parallel kinematics
machines with minimum stiffness limits: Collinear stiffness value based approach,
Mechanism and Machine Theory, 49:67-86, 2012.
[10] Hong-Zhou Jiang, Jing-Feng He, Zhi-Zhong Tong, Characteristics analysis of joint
space inverse mass matrix for the optimal design of a 6-DOF parallel manipulator,
Mechanism and Machine Theory, 45(5): 722-739, 2010.
[11] Hong-zhou Jiang, Jing-feng He, Zhi-zhong Tong, Wei Wang, Dynamic isotropic
design for modified GoughStewart platforms lying on a pair of circular
hyperboloids, Mechanism and Machine Theory, 46(9):1301-1315, 2011
Paper II Paper II

Behrouz Afzali-Far, Per Lidstrm, Kristina Nilsson


Parametric Damped vibrations of Gough-Stewart Platforms for
Symmetric Configurations
Technical Rep. ISRN LUTFD2/TFME-13/1014-SE
Division of Mechanics, Lund University
Submitted to Journal of Mechanism and Machine Theory, 2013

51
Parametric Damped Vibrations of Gough-Stewart Platforms for
Symmetric Configurations
Behrouz Afzali-Fara, , Per Lidstrmb and Kristina Nilssonb
a
Division of Machine Elements, Lund University, P.O. Box 118, Lund 22100, Sweden
b
Division of Mechanics, Lund University, P.O. Box 118, Lund 22100, Sweden

Abstract
Modal behavior of a Gough-Stewart Platform (GSP) is sensitive to several variables related to
its inertia, damping and stiffness as well as its complex 3-D geometry. To optimize its
dynamical performance, due to the complications of this system, it is crucial to have the
equations parametrically at the neutral configuration. However, in the literature, no complete
parametric solution to this problem is presented. In this paper, we establish a fully-parametric
and closed-form formulation for the vibrations of GSPs where we have additionally
introduced strut damping in the system. Parametric expressions of the damped
eigenfrequencies and the corresponding eigenvectors as well as the Jacobian, stiffness and
damping matrices are developed, all parametrically in terms of the design variables.
Interestingly, despite the complexity of the system, it is shown how well-structured algebraic
expressions are obtained using the Cartesian-space approach. Having analytically studied the
eigenvectors, the necessary and sufficient conditions for fully decoupled vibrations are
formulated. Finally, using a reference GSP, the sensitivity of the damped eigenfrequencies to
stiffness and damping variations are investigated accompanied by a cross-check with an FEM
simulation.
Keywords
Gough-Stewart platforms; parametric modal analysis; decoupled vibrations; Cartesian space;
stiffness matrix; damped eigenfrequency

1 Introduction
Interestingly, the study of manipulators dates back to ancient times, see [1]. Generally,
manipulators are categorized by their kinematic architecture, where parallel kinematic
manipulators are those with a closed kinematic chain. Gough-Stewart Platforms (GSPs),
which are also known as hexapods have, among the six Degrees Of Freedom (DOFs) parallel
kinematic manipulators, attracted the most attention of researchers and engineers. This has
been the case ever since the concept of GSPs was introduced by Gough and Stewart in 1947
and 1965, respectively [2-4]. GSPs have a wide range of applications, e.g. in vibration
isolation, machining technology, astronomical telescopes and surgical robots, see [5, 6]. From
the mid-1990s, when the first parallel machine tools came into practice, there has been an
increasing interest in studying stiffness and vibrational properties of GSPs, see [7, 8].
The bandwidth and precision of a GSP is mechanically restricted by its eigenfrequencies,
while the eigenfrequencies are sensitive to the geometrical variables. By definition there are
six independent geometrical variables for a GSP at its initially symmetrical configuration

Corresponding author. Tel.: +46-46-222-9094; e-mail: behrouz.afzali_far@mek.lth.se

53
(neutral configuration). If a GSP is in a gravity field, the neutral configuration is defined
where the structure is in its static equilibrium. Considering this geometrical complexity, and
in order to optimize the dynamical performance, it is crucial to have access to a parametric
vibrational model of the system. To obtain such a model, firstly parametric kinematics and
dynamic equations in terms of the design variables are required. See the detailed definition of
design variables at the end of section 2.2. Next, the eigenvalue problem based on inertia,
stiffness and damping matrices has to be obtained in terms of design variables. To obtain
more realistic eigenfrequencies of GSPs, it is also essential to consider damping in the system.
In particular, the axial damping of the struts should be added into the model. This is due to for
instance the viscous damping in the actuators as well as additional damping elements in some
applications (e.g. shock absorbers) which could lead to a dramatic change in the vibrational
behavior.
Performing the parametric modal analysis is a challenging process which requires major
simplifications of the complex equations that are obtained. Due to these complexities, a
straightforward and useful parametric formulation of the vibrational problem can only be
done for the neutral configuration. Introducing asymmetric configurations means adding 5
additional parameters. As a result even the expressions related to the Jacobian matrix
increases in size dramatically and accordingly makes it impossible to obtain useful equations
for dynamics and modal analysis. Furthermore, on the one hand, many GSP applications that
require dynamical performance optimization have quite limited workspace e.g. optical
collimation systems and hexapod nano-positioners. On the other hand, even for the
applications with large workspace the neutral configuration is the most efficient one to be
optimized. Hence, in both cases having a fully-parametric model of the vibrations at the
neutral configurations is of great importance. However, asymmetric configurations of the
system can be studied using numerical methods. The importance of the neutral configuration
in optimization is further motivated by Jiang et al. [9].
Generally in the literature, the Lagrangian method, see [10-13], and the Newton-Euler
method, see [14-17], are most frequently used for GSP dynamic analysis. See also [18].
However, due to the lack of strut flexibility in these models or the lack of parameterized
equations of motion in terms of design variables, these references could not be directly used
to carry out the modal analysis. Regarding the stiffness modeling, there are a few papers
dealing with the stiffness modeling and the modal analysis of GSPs. Those references can be
divided into two categories; static-stiffness analyses and dynamic based analyses. In the first
group, see [19-23], the stiffness matrix is mainly studied using Jacobian-based approaches
and this is done without modeling the dynamics of the system. In the second group, see [24-
30], the dynamical behavior is taken into account by dealing with the eigenfrequencies of the
system. Mukherjee et al. [24] numerically solve the eigenvalue problem of a 6-UPS Stewart
platform. Numerical calculations of eigenfrequencies are also performed by Mahboubkhah et
al. [25, 26, 29] followed by FEM simulations. Jiang et al. [27] analytically solve the
eigenmodes of the inverse mass matrix M 1 in the joint-space as a method to study the
undamped vibrations. Tong et al. [28], calculate the singular values of the product matrix of
the inverse mass and the stiffness matrices M 1K in the Cartesian-space for a class of
decoupled modified GSPs. Tian et al. [30] consider damping to analyze its influence on
decoupling of GSPs but do not present an parametric solution of the damped eigenvalue
problem. In addition, the authors of this paper have also carried out an analytical GSP
optimization, based on the eigenfrequencies in an advanced optical telescope application [31].
However, in the literature, a complete parametric solution to the damped eigenvalue problem

54
for the neutral configuration of GSPs is not available and in particular a parametric
formulation of damped eigenfrequencies in terms of design variables is not addressed.
The aim of this paper is to fill in these gaps and to establish a fully parametric model of the
GSPs damped vibrations in terms of design variables for the neutral configuration. The
vibrational model presented in this paper is a parametric study covering the kinematics,
equations of motion, damped vibrations and decoupled vibrations of GSPs, which can be
directly used to design, analyze and optimize every application of GSPs. This model primarily
includes the analytical formulations of inertia, stiffness and damping matrices as well as
results such as analytical damped eigenfrequencies and eigenvectors in the Cartesian-space.
The parametric model developed in this paper can be potentially employed in the optimization
of GSPs using different techniques [32-33]. In particular, this model gives fundamental
information to a forthcoming paper on the GSP geometrical optimization. In the developed
model, 12 independent variables are introduced which represent inertia, stiffness, damping
and geometry of the system, see Eq. (23). Having considered these 12 variables including
strut damping, the parametric model presented in this paper serves as a comprehensive fully-
parametric reference for the study of GSP vibrations.
In this work, the equations of motion are linearized about a parametric neutral
configuration. This parametric neutral configuration is defined as any symmetric
configuration along the z-axis. The obtained vibrational model is based on a general inertia
property definition of the platform and the axial flexibility and damping of the struts. Prior to
establishing the vibrational model, the parameterization of the geometrical dimensions is done
by introducing variables that are practical from the design point of view. Bryant angles are
used for the rigid body kinematics and the dynamic equations are obtained by the Lagrangian
method. Finally, for a reference GSP the analytical method is cross-checked with a FEM
simulation developed by using ABAQUS and the behavior of six eigenfrequencies with
respect to stiffness and damping is investigated.
2 Parametric Damped vibrations of GSPs
To establish the parametric model of the GSP damped vibrations, analytical solutions to
kinematics, Jacobian matrix, equations of motion and the damped eigenvalue problem of the
system are required. This parametric approach is presented in this section in order to finally
formulating the damped vibrational model of GSPs. This vibrational model is parameterized
based on design variables of the system including the geometrical design variables, the inertia
properties of the platform, the axial flexibility of the struts and the axial Rayleigh damping of
the struts.
2.1 Kinematics
A general schematic drawing of a GSP, with six joints on the planar base and six joints on the
planar platform with a tri-axially symmetric arrangement, is shown in Fig. 1. The geometry is
parameterized using a set of geometrical design variables which are practical from a design
point of view. These design variables consist of the radius of the base Rb , the radius of the
platform Rp , the initial distance from the base to the platform H, the distance from the
platform-plane to the platform center of mass h, half of the angel between the two
neighboring joints on the base plane Gb and on the platform plane G p , see Fig. 2. The initial
length of the struts is denoted l0 . To simplify the equations, a set of non-dimensional
parameters is introduced as

55
Platform: m, I xx , I yy , I zz

z y
ep ep
x
ep
cp , Op
h p1
P4 P3 P2
P1
Rp

P5 P6

l3 t l2
H l4 l1 l0

l5 l6
B3 B2

z y
eb eb
Rb
b1
Ob x
eb
B4 Base B1
B5 B6

Fig. 1 Schematic drawings of a general GSP

Rp H h l0
PR , PH , Ph , Pl (1)
Rb Rb Rb Rb
Geometrical design space is defined by

Gs ^ R , P , P
b R H , Ph , Gb , G p \6 Rb ! 0, PR ! 0, 0  PH  Pl , Gb  G p  S 3, Gb , Gb t 0 ` (2)

y
eb
B2

B3

P2 G
p
Gp
P1

P3 B1
60
0 Gb
x
60
0 Gb eb
P4 B6
Rp

P6

P5 Rb
B4

B5

Fig. 2 Top view of a general GSP

56
The base-frame eb (e px e py e pz ) are the two right-
(ebx eby ebz ) and the platform-frame e p
handed orthonormal bases used for the kinematics formulation. The base-frame is fixed to the
base with Ob as the origin, whereas the platform-frame is located at O p on the platform
plane. B1, B2, , B6 are points representing the joints which are fixed to the base. P1, P2, ,
P6 are points representing the joints which are fixed to the platform. In the neutral
configuration, the z-axes of the base-frame and the platform-frame are collinear which means
that the Platforms center of mass cp is located on this common axis. Throughout the text the
indices b and p denote quantities associated with base and platform, respectively. The
notation used in this paper is given in the appendix.
The platform is considered to be a rigid body with six DOFs. From the geometry of GSPs,
according to Fig. 1, we may formulate the equation
li t  pi  bi , i 1,..., 6 (3)

where, t pObc p , pi pBi Pi and bi pOb Bi are position vectors shown in Fig. 1. The bases eb
and ep are related by

ep Reb eb > R@e (4)


b

where R SO(V ) and > R @e is its matrix representation in the basis eb . Eq. (3) may now be
b

written
eb > li @e eb > t @e  ep > pi @e  eb > bi @e (5)
b b p b

Combining Eqs. (4) and (5), the following is obtained


eb > li @e eb > t @e  eb > R@e > pi @e  eb > bi @e (6)
b b b p b

Thus, the matrix representation of the geometric Eq. (3) is obtained as

> li @e > t @e  > R@e > pi @e  >bi @e


b b b p b
(7)

The angular velocity for the platform can be written


 T)
ax( RR e p > @ e (8)
p


 T V
where ax RR is the axial vector corresponding to the skew-symmetric tensor
 T Skew(V ) . Hence, the matrix representation of the angular velocity can be written
RR
< T
> @e > R @ e > @e > R @e ax > R @e > R @e Z Z p, y Z p,z
T T T
p,x (9)
p b b b
b b

Rotation tensors can be represented in many ways as a combination of rotations around
three independent axes. Mainly there are two possibilities which are referred to as Euler and
Bryant angle representations. In the literature of the GSP kinematic formulation, it seems that
the application of Euler angles is more frequent; however, the application of Bryant angles
can also be found e.g. in [34, 35]. In this paper Bryant angles are chosen to formulate the
kinematics. This is mainly due to two reasons. First, for the evaluation of the rotational
eigenmodes, the Bryant angles have a clear physical meaning such that for decoupled
57
vibrations, they can correspond to three distinct rotational eigenmodes. This property is
studied in section 2.4. Second, Bryant angles do not have singularity problems at the neutral
configuration of GSPs as Euler angles would. The rotation matrix based on Bryant angles is
cos E cos J  cos E sin J sin E

> R @e cos D sin J  sinD sin E cos J cos D cos J  sin D sin E sin J  sinD cos E (10)
sin D sin J  cosD sin E cos J cosD cos E
b

sin D cos J  cosD sin E sin J

where D , E and J are consecutive rotations around x , y and z axes, respectively starting
from the base-frame. Finally, combining Eqs. (9) and (10), the components of the angular
velocity are obtained as
Z p,x E sinJ D cos E cosJ , Z p , y E cos J  D cos E sin J , Z p , z J  D sin E (11)

2.2 Equations of motion


Throughout the text C and J denote Cartesian-space and Joint-space, respectively. The
Cartesian-space is parameterized by the generalized coordinates qC consisting of three
Cartesian coordinates representing the central mass position of the platform and three Bryant
angles representing the platform orientation

t E J \3u1 u )
T
qC x ty tz D (12)

where

t tz > t @e
T
x ty \3u1 ,
b

S
D E J ) D E J 0 d D  2S , 0 d E  , 0 d J  2S
T T
(13)
2
In the joint-space, the generalized coordinates correspond to the strut lengths which can be
expressed as

l1 ... l6 ( \  ) 6u1
T
qJ (14)

where l i li li li . Consider the coordinate transformation

qJ q J ( qC ), qC \ 3u1 u ) (15)
According to Eq. (7) it is, in component form, given by the continuously differentiable
mappings ( i 1,..., 6 )

li li (qC )

> t @e > t @e  > pi @e > pi @e  >bi @e >bi @e  2 > t @e > R @e > pi @e  2 > t @e > bi @e  2 > bi @e > R @e > pi @e
T T T T T T
b b p p b b b b p b b b b p

(16)
Consequently, the initial lengths of all struts in the neutral configuration are given by

S
12

l0 Rb 1  P R2  P H2  2P R cos(  G ) (17)
3

58
where G G b  G p , see Fig. 2.
The kinetic energy for the system can be written
1 1
T I c p  t t m (18)
2 2
where I c p is the inertia tensor of the platform at its center of mass and t eb t e ,
b

t e (tx ty tz ) is the velocity vector of the platforms center of mass and the angular
T
b

velocity is given by Eqs. (8) and (11).


The elastic potential energy V is given by the stiffness of the struts ( ki \  , i 1,..., 6 ,
see Fig. 3)
1 6
V ki 'l i2
2i1
(19)

where 'li 'qJ li  l0 , i 1,..., 6 . The Rayleigh dissipation function is given by the
damping coefficients of the struts ( ci \  , i 1,..., 6 , see Fig. 3)

1 6 2
D ci li
2i1
(20)

ki

Bi Pi

ci
Fig. 3 Stiffness and damping components of each strut (i 1,..., 6)

Thus, the Lagrangian (L T V ) is written


1 1 1 6
L I c p  t t m  ki 'l i2 (21)
2 2 2i1
and the equations of motion
d wL wL wD
  0, i 1,..., 6 (22)
dt wqi wqi wqi
If the struts have identical stiffness and damping properties ( ki k and ci c ), the number
of design variables reduce from 22 to 12. Hence, the Geometrical design space is defined by
12 parameters as follows

Gs ^ d , I xx , I yy , I zz , m, k , c \12 d Gs , m, k , c ! 0 ` (23)

59
where I xx , I yy , I zz , m represent inertia properties of the platform, see section 2.3, and the
geometrical design space Gs is defined in Eq. (2). It should be noted that GSPs are designed
to have relatively light struts as well as the fact that the struts have less transplacements in
comparison with the platform (due to a stationary point Bi on each strut). As a result, in
nearly all the existing GSPs the kinetic energy of the platform is much larger than the kinetic
energy of the struts. Hence, the strut inertia is neglected in this paper. In a forthcoming paper
we have dealt with the strut inertia separately which, due to the mathematics involved,
required a full-length paper.

2.3 Damped vibrations


A linearization of Eq. (22) at the neutral configuration q0 , q0 06u1 , results in the equations
Mq  Cq  Kq 06u1 (24)

where q 'qC or 'qJ and M , C and K are the linearized inertia, damping and stiffness
matrices which are defined by
w 2T w2 D w 2V
M ,C , K (25)
wq 2 q0 , q0
wq 2 q ,q wq 2 q0
0 0

respectively. The matrices M , C, K Sym (\6u6 ) are parametric constant matrices.


The platform is defined as a general rigid body with the consideration of its inertia
properties including its mass m , its moments of inertia I xx , I yy , I zz and its center of mass
distance h from the platform-plane. It is clear that the non-diagonal elements of the inertia
tensor only appear if one places the platform in a non-principal direction with respect to the
platform-frame. It is a practical and efficient choice to keep the platform attached to the
platform-frame in its principal directions in order to avoid further complications. Moreover,
due to the complexity of the stiffness matrix, having a diagonal inertia matrix is of great
interest in order to obtain well-structured solutions to the eigenvalue problem. Hence, we
assume that ep represents principal axis directions at the mass center of the platform with the
corresponding moments of inertia I xx , I yy , I zz . It should be indicated that, at the neutral
configuration, the platform-frame and base-frame have the same directions. Consequently, the
angular velocity of the platform is the same as the platform-frames. Accordingly, the kinetic
energy is written
1 2 2 2
T (t x , t y , t z , D , E , J , tx , ty , tz , D , E , J ) 1

t x  t y  t z m  I xx E sin J D cos E cos J
2

2 2
I yy E cos J  D cos E sin J  I zz J  D sin E
1 2 1 2
(26)
2 2
and the linearized inertia matrix in the Cartesian-space then reads
w 2T
MC diag ( m m m I xx I yy I zz ) \ 6 u6 (27)
wqC2 qC ,0 , qC ,0

60
Whereas, to formulate the stiffness matrix, one needs to write the elastic potential energy
k 6
V 'l i2 , l i l i (t x , t y , t z , D , E , J )
2i1
(28)

Then, to calculate the stiffness matrix, first the coordinate transformation mapping from C
to J may be written
wq J
qJ q J (qC ) q J (qC ,0 )  (qC  qC ,0 )  o( qC  qC ,0 ) (29)
wqC qC ,0

0 0 H  h 0 0 0 , qJ ,0 l0 ... l0 represent the neutral


T T
where qC ,0 q J (qC ,0 )
configuration and
o(qC  qC ,0 )
lim 0 (30)
qC  qC ,0 o0 qC  qC ,0

Thus the linearized transformation, at the neutral configuration, is given by


'qJ J 'qC (31)
where 'qC qC  qC ,0 and 'qJ qJ  qJ ,0 and

wq J
J \ 6u6 (32)
wqC qC ,0

is the Jacobian matrix of the coordinate transformation. The matrix elements jik of J are
obtained according to

wli 1 w l i li 1 w l i li
jik (33)
wqC ,k qC ,0
2li wqC ,k q 2l0 wqC ,k
C ,0 q C ,0

where qC ,k , k 1,..., 6 are the elements of qC . One may, according to Eq. (16), write

w li li
w > t @e > t @e  2 > t @e > R@e > pi @e  2 > t @e > bi @e  2 > bi @e > R @e > pi @e
T
b b
T
b b p
T
b b
T
b b p
(34)
wqC ,k wqC ,k
After the calculations, the Jacobian matrix can be structured using 14 variables as follows

j11 j12 j13 j14 j15 j16



j21 j22 j13 j24 j25  j16
1 j31 j32 j13 j34 j35 j16
J (35)
2Pl j31  j32 j13  j34 j35  j16
j21  j22 j13  j24 j25 j16

j11  j12 j13  j14 j15  j16

61
The structure of the Jacobian matrix has a special symmetry property which is in accordance
with [12], see also [36, 37]. The elements of the Jacobian matrix are parametrically obtained
as
j11 P R cos G p  2 cos G b  3P R sin G p

j P R cos G p  cos G b  3P R sin G p  3 sin G b
21

j31 cos G b  2P R cos G p  3 sin G b

j12 3P R cos G p  P R sin G p  2sin G b

j22 P R sin G p  3 cos G b  sin G b  3P R cos G p

j sin G b  3P R cos G b  2P R sin G p
32
j 2P H
13

j14 Rb P R P H  Ph

3 cos G p  sin G p  2Ph sin G b


 P  sin G  P

j24 Rb P R P H 3 cos G p 3 cos G b  sin G b
h p h


j34 Rb 2P R P H  Ph sin G p  Ph sin G b  3Ph cos G b


j15 Rb 2Ph cos G b  P R P H  Ph cos G p  3P R P H  Ph sin G p

j Rb 3P R P H  Ph sin G p  3Ph sin G b  Ph cos G b  P R P H  Ph cos G p
25

j35 Rb 2P R P H  Ph cos G p  Ph cos G b  3Ph sin G b

j Rb P R 3 cos(G b  G p )  sin(G b  G p )
16
(36)
Determinant of J

0 /3 
= +p
b

Fig. 4 Determinant of the Jacobian matrix as a function of G ; the hatched region shows the
geometrical design space

62
where G is that defined in Eq. (17). The determinant of the Jacobian matrix is given by
27 Rb3 P R3 P H3
det( J )
4Pl6 3 3 cos G  3sin G  2sin(3G ) (37)

In Fig. 4, det( J ) is shown as a function of G . Thus

R ,Pb R , P H , Ph , G b , G p Gs det J ! 0 (38)

It then follows, by the inverse function theorem, that there exist open sets :C \ 3u1 u ) and
: J (\  )6u1 with qC ,0 :C and qJ ,0 q J ( qC ,0 ) : J such that the mapping
q J :C
: :C o : J , is bijective and ( q J :C
1
) : : J o : C is differentiable. We thus have a
regular coordinate transformation qC l qJ in the vicinity of the neutral position.

Remark: Note that the mapping qJ q J (qC ) , defined in Eq. (15), is not bijective [38].
Using the obtained Jacobian matrix, the corresponding linearized stiffness matrix in the
Cartesian-space is.
6
w 2l 6
wl wl i
T T
w 2V wl wl i
KC k i  'l i 2i k i
wqC2 i 1 wqC wqC wqC i 1 wqC wqC
qC ,0 q C ,0
q
C ,0

k xx 0 0 0 k xE 0

k yy 0 k yD 0 0
k zz 0 0 0
kJ T J (39)
kDD 0 0
sym. kEE 0

kJJ

where the five stiffness elements are expressed in terms of design variables as
3k S
k xx k yy 2
1  P R 2  2 P R cos(  G )
P l 3

k 6k P H2
zz
Pl2

3k P R Rb2 Ph2 S
kDD kEE P P 2
  P R Ph2  2P R P H P h  2Ph P H  Ph cos(  G ) (40)
Pl
2 R H
PR 3

k 3k P R2 Rb2 S
1  cos(  2G )
JJ Pl2 3

3k P R Rb S Ph
k xE k yD P H  2 Ph cos(  G )  P R P H  P R P h 
P l
2
3 P R

63
Assuming identical damping for each strut, according to Eq. (20), the Rayleigh dissipation
function is
c 6 2
D li
2i1
(41)

where
wl i wl wl i
li qC i (42)
wqC wqC wqC
The corresponding linearized damping matrix is

w2 D 6 wl T wl w 2l 6 wl T wl


CC c i
i
 li 2i c i
i

wqC2 q i 1 wq
C wqC wqC i 1 wqC wqC
C ,0
C ,0 , q qC ,0 ,qC ,0 qC ,0 ,qC ,0

cxx 0 0 0 cxE 0

c yy 0 c yD 0 0
czz 0 0 0
cJ T J O kJ J
T
(43)
cDD 0 0
sym. cEE 0

cJJ

Hence, it is proven that the damping matrix is proportional to the stiffness matrix. This can be
expressed as
c
O , CC cJ T J O kJ T J O KC (44)
k
By obtaining the structure of the inertia, stiffness and damping matrices in the Cartesian-
space, the damped eigenvectors and eigenfrequencies can be calculated by the state space
method. However, damping consideration in the system makes the parametric solution of the
damped eigenvalue problem very complicated. Therefore, at first the undamped eigenvalue
problem is solved. Then by employment of a diagonalization technique, the damped
eigenvalue problem is solved. Having solved the undamped eigenvalue problem, the modal
matrix in the Cartesian-space is obtained as

0 0 F C ,1 0 F C ,3 0

0 0 0 F C ,2 0 F C ,4
0 1 0 0 0 0
XC (45)
0 0 0 1 0 1
0 0 1 0 1 0

1 0 0 0 0 0

64
where
I yy k xx  mk EE  <1 I xx k yy  mkDD  < 2
F C ,1 , F C ,2
2mk xE 2mk yD
(46)
I yy k xx  mk EE  <1 I xx k yy  mkDD  < 2
F C ,3 2mk xE
, F C ,4
2mk yD

and <1 , < 2 are defined by

I k  mk EE  4mI yy k x2E , < 2 I k yy  mkDD  4mI xx k y2D


2 2
<1 yy xx xx (47)

It should be noted that the modal matrices for the damped and undamped vibrations are,
since we are considering Rayleigh damping, the same. Having obtained the modal matrix
from the undamped vibrations, modal mass and modal stiffness are obtained to formulate the
damped eigenfrequencies. First, by defining
q XK (48)
where K denotes normal coordinates and X is the modal matrix, the vibrational equation
given in Eq. (24) can be rewritten as
PK  ONK  NK = 0 (49)
where the modal mass and stiffness matrices are given by
P X T MX diag ( P1 ! P6 ), N X T KX diag (N1 ! N 6 ) (50)
One may rewrite Eq. (49) as
PK
i i  ON iKi  NiKi = 0, i 1,...,6
  (51)
and obtains the undamped eigenfrequencies and damping ratios according to
Ni OZi
Zi , ]i , i 1,..., 6 (52)
Pi 2

Note that the critical damping ] i 1 corresponds to


2
Ocr ,i , i 1,..., 6 (53)
Zi
Then, eigenvalues (si , i 1,..., 6 ) including real parts V i and damped eigenfrequencies Zd ,i
are given by
2
ON ON N
si V i r iZd ,i  ir i  i (54)
2 Pi 2 Pi Pi
To show that the damped eigenfrequencies corresponding to critical/over damped modes are
zero, the following representation is used

65
O2
Zd ,i Im i Zi2  Zi4 (55)
4
where Im gives the imaginary part of a complex number. Eventually, the six damped
eigenfrequencies of a general GSP are formulated as
k O2 k
2
O 2 kJJ
2
Z kJJ
Im i  , Zd ,2 Im i zz  zz
d ,1 I zz 4 I zz m 4 m


2
k xx kEE <1 O2 k k <1
Zd ,3 Im i    xx  EE 
2m 2 I yy 2mI yy 4 2m 2 I yy 2mI yy


k 2
k <2 O 2 k yy kDD <2
Zd ,4 Im i  DD    
yy

2m 2 I xx 2mI xx 4 2m 2 I xx 2mI xx


2
Z k xx kEE <1 O2 k k <1
Im i    xx  EE 
d ,5
2m 2 I yy 2mI yy 4 2m 2 I yy 2mI yy


2
k yy kDD <2 O 2 k yy kDD <2
Zd ,6 Im i     
2m 2 I xx 2mI xx 4 2m 2 I xx 2mI xx

(56)
Using Eqs. (40) and (47), these eigenfrequencies may be represented directly in terms of
the design variables. It should be noted that the sequence of the eigenfrequencies varies
according to the input variables. In spite of the complications regarding the solution and the
simplifications of the damped eigenfrequencies, interestingly the obtained results are quite
straightforward such that the contribution of each element of the system on the
eigenfrequencies is clearly visible.
In this paper, twelve design-variables are taken into account and well-structured solutions
to the damped eigenvalue problem are parametrically formulated. The twelve design-variables
are specified in Eq. (23). However, additional detailed considerations (see [39-43]), such as
errors in the joints, may be studied using numerically solved eigenvalue problems.
2.4 Decoupled vibrations
The decoupled vibrations occur when the modal matrix is completely decoupled. This means
six pure eigenmodes corresponding to each Cartesian-space coordinates. Decoupled
eigenvectors can be accomplished in two ways. First, if the eigenvector components FC in
Eq. (45) go to zero or infinity. Second, if there exist multiple eigenfrequencies ( Z3 Z5 or
Z4 Z6 ). The latter can also be regarded as decoupled vibrations due to the fact that the
eigenmodes of multiple eigenfrequencies can be any linear combination of their
corresponding eigenvectors. In both cases, or in the combination of these cases, the necessary
condition to accomplish a fully decoupled modal matrix is k xE k yD 0 which means a
diagonal stiffness matrix.

66
Considering k xE k yD 0 , depending on the relations between I yy k xx and mk EE as well as
between I xx k yy and mkDD , there are nine different conditions for obtaining the modal matrix.
On the one hand, if I yy k xx z mk EE and I xx k yy z mkDD , according to Eqs. (46) and (47), the
elements of these matrices are obtained as

klim F C ,1 lim F C ,2 lim F C ,3 lim F C ,4 0
xE o 0 k yD o 0 k xE o 0 k yD o 0
I yy k xx ! mk EE I xx k yy ! mkDD I yy k xx  mk EE I xx k yy  mkDD

(57)

klim F C ,1 lim F C ,2 lim F C ,3 lim F C ,4 f
o0 k yD o 0 k xE o 0 k yD o 0
xE I yy k xx  mk EE I xx k yy  mkDD I yy k xx ! mk EE I xx k yy ! mkDD

and on the other hand, if I yy k xx mk EE or I xx k yy mkDD the elements of the modal matrix are
obtained as
I yy
FC ,3 I k  FC ,1 I r

mkEE yy k xx mkEE
yy xx m
(58)
I xx
FC ,4 I xx k yy  FC ,2 r
mkDD I xx k yy mkDD
m
Using Eqs. (57) and (58), all the nine possible decoupled modal matrices can be obtained.
Among them, for the case in which I yy k xx ! mk EE and I xx k yy ! mkDD , the normalized modal
matrix for the decoupled vibrations are given by
0 0 0 0 1 0

0 0 0 0 0 1
0 1 0 0 0 0
X Cdecoupled I yy k xx ! mkEE , (59)
I xx k yy ! mkDD 0 0 0 1 0 0
0 0 1 0 0 0

1 0 0 0 0 0
Whereas, assuming I yy k xx mk EE and I xx k yy mkDD simultaneously, the decoupled modal
matrix is given by

I yy I yy
0 0 B 0 r 0
m m
I xx I xx
0 0 0 B 0 r
X decoupled m m
(60)
C I yy k xx mkEE ,
I xx k yy mkDD
0 1 0 0 0 0

0 0 0 1 0 1
0 0 1 0 1 0

1 0
0 0 0 0

67
where the conditions implies I xx I yy and the existence of four multiple eigenfrequencies.

Accordingly, it can be shown that, considering k xE k yD 0 , all the possible modal


matrices become completely decoupled. Therefore, a diagonal stiffness matrix ( k xE k yD 0)
is not only a necessary condition but also a sufficient condition for the decoupled vibrations.
Eventually, the damped eigenfrequencies corresponding to the decoupled vibrations are
k 2
k
i JJ  O JJ , Z O 2 k zz
2 2
Z k
Im Im i zz

d ,1 I zz 4 I zz
d ,2
m 4 m


2
2
k EE O 2 k EE kDD O 2 kDD
Zd ,3 Im i  , Zd ,4 Im i  (61)
I yy 4 I yy I xx 4 I xx


k k 2
O 2 k xx O 2 k yy
2
Z Z xx
 yy

d ,5
Im i Im i
m
m 4 m 4 m
d ,6


where according to Eq. (40), it is known that k xx k yy . Therefore, it is proven that if the
decoupling condition ( k xE k yD 0 ) is satisfied, there exist at least two multiple
eigenfrequencies (Zd ,5 Zd ,6 ) . The eigenfrequencies given in Eq. (61) are sorted with respect
to Eq. (59). Generally, comparing the results in the Cartesian-space presented in this paper
with those given in [27] based on the joint-space coordinates, shows that although we have
additionally introduced damping to the system, employment of this Cartesian based approach
leads to simpler calculations and results. This is due to the larger number of zero elements in
the matrices belonging to the Cartesian-space.

3 Numerical study
In this section, by introducing a reference GSP, the parametric model developed in this paper
is numerically illustrated. It should be mentioned that we have not introduced any kind of
approximation in the parametric model developed in this paper (except the small-angle
assumption). The analytical method we have employed in this work is the Lagrangian
dynamics with clear analytical limitations concerning the number of variables introduced to
the model (12 variables in this model). Here only to cross-check the analytically obtained
equations, the numerical values for a reference GSP is compared to a FEM simulation using
ABAQUS. Finally, in this section the modal behavior of GSPs with respect to damping and
stiffness is studied.
The reference GSP has the following properties. Selected data are related to a GSP for a
collimation system of a large optical telescope where the inertia properties of the platform are
m 100 (kg), I xx I yy 5 (kgm2 ), I zz 10 (kgm2 )
the geometrical characteristics are
Rb 0.500 (m), P R 0.8, P H 1.8, Ph 0.2, G b 0.1745 (rad), G p 0.2618 (rad)
and the strut characteristics are

68
k 106 N/m , c 103 Ns/m
The inertia matrix in the Cartesian-space can be simply obtained using Eq. (27) as
MC diag (10 2 10 2 10 2 5 5 10 ) \6 u6
Whereas, using Eqs. (39) and (40) the stiffness matrix in the Cartesian-space is obtained as
276820.2 0 0 0 16092.1 0

0 276820.2 0 16092.1 0 0
0 0 5446359.5 0 0 0
KC
0 16092.1 0 436158.9 0 0
16092.1 0 0 0 436158.9 0

0 0 0 0 0 88483.8
The damping matrix is, according to Eq. (44), proportional to the stiffness matrix
c
O 10 3 (s) CC 1 0 3 K C Ns/m
k
The aim here is to compare and cross-check the analytically obtained eigenfrequencies and
eigenvectors with a totally different method (in this case a FEM simulation). The reference
GSP modeled in the FEM simulation has exactly the same characteristics as it does in the
analytical model. The model in ABAQUS is defined by two discrete rigid platforms, springs
and dashpots in the complex frequency procedure. Table 1 shows a comparison between the
six eigenfrequencies obtained by the analytical equations formulated in this paper and the
FEM simulation. The comparison shows an excellent agreement between these methods in
terms of modal properties.
Table 1 Damped and undamped eigenfrequencies obtained by the analytical and the finite
element methods

Z1 Z2 Z3 Z4 Z5 Z6
-1 -1 -1 -1 -1 -1
(rads ) (rads ) (rads ) (rads ) (rads ) (rads )
Method
Analy. (undamped) 52.555 52.555 94.066 233.374 295.361 295.361
FEM (undamped) 52.556 52,556 94.066 233.375 295.361 295.361
Analy. (damped) 52.537 52.537 93.962 231.780 292.122 292.122
FEM (damped) 52.537 52.537 93.962 231.780 292.120 292.120
The eigenvectors in the Cartesian-space directly represents the position and orientation of
the platform corresponding to each eigenfrequency according to Eq. (12). The modal matrix is
obtained using Eq. (45) and rearranged according to the eigenfrequencies sequence as follows
-26.246 0 0 0 -0.0019 0

0 26.246 0 0 0 -0.0019
0 0 0 1 0 0
XC
0 -1 0 0 0 1
-1 0 0 0 1 0

0 0 1 0 0 0
69
Having normalized the modal matrix, the mode shapes obtained by the analytical and FEM
results are exactly in correspondence, see Fig. 5. The given mode shapes for the damped
vibrations of the reference GSP could be useful in order to understand the modal behavior of
any GSP application.

Fig. 5 The six mode shapes of the reference GSP: (a), ..., (f) are the first to sixth modes
including xE ,1-mode , yD ,1-mode , J -mode , z -mode , xE , 2-mode and yD , 2-mode ,
respectively
70
Fig. 6 Sensitivity of the damped eigenfrequencies to the strut damping where k 10 6
(left), sensitivity of the damped eigenfrequencies to the strut stiffness where c 10 3 (right)

The sensitivity of the six eigenfrequencies to the axial strut damping is shown in Fig. 6.
These curves are plotted using Eq. (56). Combining Eqs. (44) and (53), the critical damping
with respect to each eigenfrequency is obtained as
2
ccr ,i k , i 1,..., 6 (62)
Zi
Figure 6 (left) shows that, according to Eq. (62), the higher the undamped eigenfrequencies,
the less axial damping lead to critical damping for the corresponding mode. Figure 6 (left)
also shows that the sequence of the eigenmodes changes at every intersection point of the
curves. Hence, it is important to note that the sequence of the eigenfrequencies in damped and
undamped vibrations might be different.
Figure 6 (right) illustrates the damped eigenfrequencies with respect to stiffness. These
curves are also plotted using Eq. (56). Note that since the scale of stiffness is relatively large
in Fig. 6 (right), the region which corresponds to critical damping is not visible, in this region
a similar type of sequence change as observed in Fig. 6 (left) is present. Therefore, for
relatively large stiffness values (when O is approaching zero) the sequence of the
eigenfrequencies remains unchanged. This may also be expressed as

Oc constant

1 according to Eq. (55)
k
   o lim Z d ,i
k of
c constant
 k (63)

However, if instead of c , O is kept constant, a different behavior is observed. This is shown


in Fig. 7 where damped eigenfrequencies are plotted versus stiffness with O 10 3 (s). It is
observed that all the damped eigenfrequencies have a peak at a unique value equal to Zd*
=1000 (rad/s). Note that Fig. 7 is plotted in a semi-logarithmic scale in order to show the
curves clearly.
According to Eq. (55), considering weak damping and for a constant O , Zd* is written

71
Fig. 7 Sensitivity of the damped eigenfrequencies to the strut stiffness where O 10 3


O 2 O2
Zd* max ^Zd ,i ` max Zi2  Zi4 Zd*2 max Zi2  Zi4 (64)
Zi Zi Zi
4  4

Expression A can be maximized as follows

wA O2 2
1 Zi2 Zi Z* (65)
w Zi
2
2 O

where, since O is a constant, there exists a unique Zi Z * in order to maximize A . Finally,


substituting Eq. (65) in Eq. (64), Zd* is obtained as

1 k
Zd* (66)
O c

Therefore, it is shown that by considering a constant O , all the damped eigenfrequencies


are bound to an upper limit equal to O 1 . Comparing Fig. 7 with Fig. 6 (right) shows how
considerably the results are changed. Damping consideration is especially important for high
stiffness applications in which, even with a small O , if stiffness is increased critical damping
can occur. Furthermore, for GSP applications with axial dampers installed on struts (e.g.
shock absorbers), damped and undamped eigenfrequencies might differ quite substantially in
such a way that there can be sequence-changes in the damped eigenfrequencies. The
analytically drawn conclusions in this section are applicable for GSPs with any number of
viscous dampers along struts.

72
4 Conclusions and future work
Since in the literature a fully parametric model of the GSP damped vibrations for the neutral
configuration has not been addressed, the aim of this paper is to establish this fully parametric
model in terms of the design variables that can be used directly to analyze and optimize GSPs
in any application. Parametric formulations of eigenvectors and damped eigenfrequencies as
well as stiffness, damping and Jacobian matrices are presented in terms of design variables. It
is shown that the use of Lagrangean dynamics and Bryant angles in the kinematic formulation
are suitable choices for a well-structured vibrational formulation of GSPs. Due to the larger
number of zero elements in the matrices belonging to the Cartesian-space, it is concluded that
using the Cartesian-space coordinates leads to less complicated calculations of the eigenvalue
problem. It is also shown that decoupled vibrations occur if and only if the stiffness matrix is
diagonal which leads to at least two multiple eigenfrequencies. The conditions for the
decoupled vibrations are analytically studied. For a reference GSP an FEM simulation is
carried out to support the analytical results which show an excellent agreement in the results.
Finally, the parametric model developed in this paper is numerically demonstrated for a
reference GSP and the sensitivity of its six eigenfrequencies to stiffness and damping is
investigated. It is concluded that damping consideration is crucial in order to obtain accurate
results and it is shown that for a constant O , there exists a maximum possible value for the
damped eigenfrequencies which is equal to O 1 .
In our future works, we intend to study the damped vibrations in the joint-space and further
refine our model by investigating the influence of the strut inertia on the modal behavior of
GSPs. We also aim to develop an analytical optimization method in order to achieve optimal
dynamical performance.

73
Appendix Mathematical notation
In this paper \ denotes the set of real numbers and \  the set of positive real numbers. The
set of n-dimensional, real column vectors is denoted by \ nu1 and the null vector in \ nu1 is
written 0nu1 . \ nu n denotes the set of real matrices of order n u n with the null matrix written
0nun . If A \ nun then AT \ nun denotes the transpose of A . Sym(\nun ) ^ A \ nun
A AT `
and Sym (\
 nun
) ^ A Sym(\ nun
`
) A is positive definite . The determinant of A \ nu n
is
1
denoted by det ( A) and if det ( A) z 0 then A denotes its inverse. 1nun is the identity matrix
nu n
in \ . Let E denote a 3-dimensional Euclidean point space with the corresponding
translation vector space V , i.e. dim (V ) 3 . Points in E are denoted by A, B..., X , Y ,... and
vectors in V by a , b,..., u, v ,... . The scalar and vector products of two vectors a and b are
denoted a b , a u b , respectively. The norm a of a vector a is defined by a a a . If
e (e1 e2 e3 ), ei V , i 1, 2, 3 is a basis in V then u = e > u@e e1u1  e2u2  e3u3 where

> u@e u1 u3 \ 3u1 is the matrix representation of the vector u V in the basis e .
T
u2
The relative position vector from point AE to point B E is denoted pAB . The space of all
second order tensors A on V , i.e. linear mappings V oV , is denoted End (V ) . If
A End (V ) then Ae = e > A@e where

A11 A12 A13



> A @e A21 A22 A23 \ 3u3 (67)
A A33
31 A32

is the matrix representation of A End (V ) in the basis e . The set of all rotations on V is
^
denoted SO(V ) = A End (V ) A A
T
AAT `
1, det A 1 . The space of all skew-symmetric

^
tensors by Skew(V ) = A End (V ) A
T
`
 A . If A Skew(V ) then ax ( A) V denotes its
axial vector, i.e. Au ax ( A) u u , u V . Given a V then the tensor a u 1 Skew(V ) is
defined by (a u 1) u a u u, u V .

References
[1] T.G. Chondros, Archimedes life works and machines, Mech. Mach. Theory, 45 (2010).
1766-1775.
[2] J. P. Merlet, Parallel Robots, second ed., Springer, 2006.
[3] V.E. Gough, S.G. Whitehall, Universal Tyre test Machine, Proceedings of 9th
International Congress FISITA, May 1962, pp. 117-137.
[4] D. Stewart, A platform with six degrees of freedom, Proceedings of The Institution of
Mechanical Engineers 1847-1982 (vols 1-196), 1965, pp. 371-386.

74
[5] A. Preumont, M. Horodinca, I. Romanescu, B. de Marneffe, M. Avraam, A. Deraemaeker,
F. Bossens, A. Abu Hanieh, A six-axis single-stage active vibration isolator based on Stewart
platform, J. Sound Vib. 300 (2007) 644-661.
[6] Y.D. Patel, Parallel Manipulators ApplicationsA Survey, Modern Mechanical
Engineering, 02 (2012) 57-64.
[7] M. Terrier, A. Dugas, J.Y. Hascoet, Qualification of parallel kinematics machines in high-
speed milling on free form surfaces, Int. J. Mach. Tool Manu. 44 (2004) 865-877.
[8] V. D. Sapio, Some approaches for modeling and analysis of a parallel mechanism with
stewart platform architecture, Proc. 1998 ASME Int., 1998.
[9] H.Z. Jiang, J.F. He, Z.Z. Tong, W. Wang, Dynamic isotropic design for modified Gough-
Stewart platforms lying on a pair of circular hyperboloids, Mech. Mach. Theory, 46 (2011)
1301-1315.
[10] Z. Geng, L.S. Haynes, J.D. Lee, R.L. Carroll, On the dynamic model and kinematic
analysis of a class of Stewart platforms, Robot. Auton. Syst, 9 (1992) 237-254.
[11] J.D. Lee, Z. Geng, A Dynamic-Model of a Flexible Stewart Platform, Comput. Struct. 48
(1993) 367-374.
[12] K. Liu, F. Lewis, G. Lebret, D. Taylor, The Singularities and Dynamics of a Stewart
Platform Manipulator, J. Intell. Robot. Syst. 8 (1993) 287-308.
[13] H. Abdellatif, B. Heimann, Computational efficient inverse dynamics of 6-DOF fully
parallel manipulators by using the Lagrangian formalism, Mech. Mach. Theory, 44 (2009)
192-207.
[14] B. Dasgupta, T.S. Mruthyunjaya, A Newton-Euler formulation for the inverse dynamics
of the stewart platform manipulator, Mech. Mach. Theory, 33 (1998) 1135-1152.
[15] B. Dasgupta, T.S. Mruthyunjaya, Closed-form dynamic equations of the general Stewart
platform through the Newton-Euler approach, Mech. Mach. Theory, 33 (1998) 993-1012.
[16] S. Pedrammehr, M. Mahboubkhah, N. Khani, Improved dynamic equations for the
generally configured Stewart platform manipulator, J. Mech. Sci. Technol. 26 (2012) 711-
721.
[17] Y.A. Cheng, G.X. Ren, S.L. Dai, The multi-body system modelling of the Gough-
Stewart platform for vibration control, J. Sound Vib. 271 (2004) 599-614.
[18] L.K. Liu, G.T. Zheng, W.H. Huang, Octo-strut vibration isolation platform and its
application to whole spacecraft vibration isolation, J. Sound Vib. 289 (2006) 726-744.
[19] V.T. Portman, V.S. Chapsky, Y. Shneor, Workspace of parallel kinematics machines
with minimum stiffness limits: Collinear stiffness value based approach, Mech. Mach.
Theory, 49 (2012) 67-86.
[20] J. Chen, F. Lan, Instantaneous stiffness analysis and simulation for hexapod machines,
Simul. Model. Pract. Th. 16 (2008) 419-428.
[21] V.T. Portman, B.Z. Sandler, E. Zahavi, Rigid 6-DOF parallel platform for precision 3-D
micromanipulation, Int. J. Mach. Tool Manu. 41 (2001) 1229-1250.
[22] B.S. El-Khasawneh, P.M. Ferreira, Computation of stiffness and stiffness bounds for
parallel link manipulators, Int. J. Mach. Tool Manu. 39 (1999) 321-342.

75
[23] C. Gosselin, Stiffness Mapping for Parallel Manipulators, Ieee T. Robotic Autom. 6
(1990) 377-382.
[24] P. Mukherjee, B. Dasgupta, A.K. Mallik, Dynamic stability index and vibration analysis
of a flexible Stewart platform, J. Sound Vib. 307 (2007) 495-512.
[25] M. Mahboubkhah, M.J. Nategh, S.E. Khadem, Vibration analysis of machine tool's
hexapod table, Int. J. Adv. Manuf. Tech. 38 (2008) 1236-1243.
[26] M. Mahboubkhah, M.J. Nategh, S.E. Khadem, A comprehensive study on the free
vibration of machine tools' hexapod table, Int. J. Adv. Manuf. Tech. 40 (2009) 1239-1251.
[27] H.Z. Jiang, J.F. He, Z.Z. Tong, Characteristics analysis of joint space inverse mass
matrix for the optimal design of a 6-DOF parallel manipulator, Mech. Mach. Theory, 45
(2010) 722-739.
[28] Z.Z. Tong, J.F. He, H.Z. Jiang, G.R. Duan, Optimal design of a class of generalized
symmetric Gough-Stewart parallel manipulators with dynamic isotropy and singularity-free
workspace, Robotica, 30 (2012) 305-314.
[29] S. Pedrammehr, M. Mahboubkhah, N. Khani, A study on vibration of Stewart platform-
based machine tool table, Int. J. Adv. Manuf. Tech. 65 (2013) 991-1007.
[30] T.X. Tian, H.Z. Jiang, J.F. He, Z.Z. Tong, Influence of passive joint damping on modal
space decoupling for a class of symmetric spatial parallel mechanisms, Applied Mechanics
and Materials, 2013, pp. 1152-1157.
[31] B. Afzali Far, P. Lidstrm, K. Nilsson, Analytical Stiffness Optimization of High-
Precision Hexapods for Large Optical Telescope Applications, Proceedings of the 25th
Nordic Seminar on Computational Mechanics, Lund, Sweden, October 2012, pp. 257-260.
[32] C. Menon, R. Vertechy, M. C. Markot, V. Parenti-Castelli, Geometrical optimization of
parallel mechanisms based on natural frequency evaluation: application to a spherical
mechanism for future space applications, Trans. Rob., 25 (2009) 12-24.
[33] B. Lim, F.C. Park, Minimum Vibration Mechanism Design Via Convex Programming, J
Mech Design, 131 (2009) 011009.
[34] P. Xu, D.H. Wang, Vibration damping modeling of Stewart platform through Newton-
Euler approach, Smart Structures and Materials 2005: Smart Structures and Integrated
Systems, 2005, pp. 650-661.
[35] Y.X. Su, B.Y. Duan, C.H. Zheng, Genetic design of kinematically optimal fine tuning
Stewart platform for large spherical radio telescope, Mechatronics, 11 (2001) 821-835.
[36] T. Huang, D.J. Whitehouse, J.S. Wang, The local dexterity, optimal architecture and
design criteria of parallel machine tools, Cirp Annals 1998 - Manufacturing Technology, Vol
47, No 1, 47 (1998) 347-351.
[37] T. Huang, J. Wang, C.M. Gosselin, D.J. Whitehouse, Kinematic Synthesis of Hexapods
with Specified Orientation Capability and Well-Conditioned Dexterity, Journal of
Manufacturing Processes, 2 (2000) 36-47.
[38] B. Dasgupta, T.S. Mruthyunjaya, A canonical formulation of the direct position
kinematics problem for a general 6-6 stewart platform, Mech Mach Theory, 29 (1994) 819-
827.
[39] A.D. Dimarogonas, S.A. Paipetis, T.G. Chondros, Analytical Methods in Rotor
Dynamics: Second Edition, Springer, 2013.
76
[40] T. Chondros, Identification of cracks in welded joints of complex structures, J Sound
Vib, 69 (1980) 531-538.
[41] H.L. Shi, H.J. Su, N. Dagalakis, J.A. Kramar, Kinematic modeling and calibration of a
flexure based hexapod nanopositioner, Precis Eng, 37 (2013) 117-128.
[42] T. Huang, D.G. Chetwynd, D.J. Whitehouse, J. Wang, A general and novel approach for
parameter identification of 6-DOF parallel kinematic machines, Mech Mach Theory, 40
(2005) 219-239.
[43] W. Tian, W. Gao, D. Zhang, T. Huang, A general approach for error modeling of
machine tools, International Journal of Machine Tools and Manufacture, 79 (2014) 17-23.

77
Paper III Paper III

Behrouz Afzali-Far, Anette Andersson, Kristina Nilsson and Per Lidstrm


Parametric Damped Vibrations of Gough-Stewart Platforms
Taking into Account Inertia of the Struts
Technical Rep. ISRN LUTFD2/TFME-14/1014-SE
Division of Machine Elements, Lund University
Submitted to Journal of Mechanism and Machine Theory, 2014

79
Parametric Damped Vibrations of Gough-Stewart
Platforms Taking into Account Inertia of the Struts

Behrouz Afzali-Fara,, Anette Anderssona , Kristina Nilssonb , Per Lidstrmb


a Division of Machine Elements, Lund University, P.O. Box 118, SE-221 00 Lund, Sweden
b Division of Mechanics, Lund University, P.O. Box 118, SE-221 00 Lund, Sweden

Abstract

Consideration of strut inertia can have significant influence on the modal behavior
of a Gough-Stewart Platform (GSP). However, in the literature, the strut inertia has
not yet been taken into account in its vibrations study with a parametric approach.
In this paper, we present a fully parametric approach to formulate the damped vibra-
tions of GSPs taking into account the strut masses and moments of inertia. For the
first time, a combined 6 6 equivalent inertia matrix (mass matrix), including both
the inertia properties of the platform and the struts, is formulated parametrically.
Consequently, the eigenvectors and the damped eigenfrequencies are also parametri-
cally developed in the Cartesian-space. Furthermore, the conditions for the decoupled
vibrations and the effect of strut inertia on these conditions are analytically investi-
gated. With the aid of a reference GSP, the influence of the strut inertia on its modal
behavior is systematically investigated. Accordingly, it is concluded that the para-
metric equations established in this paper can be directly employed in the analysis,
optimization and control of GSPs in any application.
Keywords: Gough-Stewart platform, Strut inertia, Analytical modal analysis,
Damped vibrations, Cartesian space, Parametric stiffness matrix, Eigenfrequency

1. Introduction

In the literature and the applications of parallel robots, the Gough-Stewart Plat-
form (GSP) is one of the most recognized mechanisms [1, 2]. In some applications
it is also known briefly as hexapod or Stewart platform. It has the simplest sym-
metric structure which can provide six Degrees-Of-Freedom (DOFs) by its six linear
actuators, see Figs. 1 and 2. Due to its parallel architecture, inherently it has some
advantages over serial robots such as higher stiffness, higher accuracy and elimination
of the cumulative error [2]. GSPs have been increasingly employed in several mod-

Coresponding author.
Email address: behrouz.afzali_far@mek.lth.se (Behrouz Afzali-Far)

Preprint submitted to Elsevier March 5, 2014


ern applications such as robotic surgery, astronomical telescopes and manufacturing
technology [35].
Despite this wide range of applications, in most of the cases the core design
requirement of GSPs is the dynamic stability which can be evaluated by modal
analysis. A GSP at its neutral configuration, can be defined by fourteen design
variables which represent its geometry, stiffness, damping and inertia and it is obvious
that the modal behavior of a GSP is also a function of the design variables. The
neutral configuration and the design variables are defined in the beginning of section
2.1 and in Eq. (1). Moreover, these design variables can be optimized based on
given requirements for a GSP modal behavior. Due to the multiple DOFs and the
complexity of the system, developing a fully parametric modal analysis is crucial to
accomplish a well-optimized GSP corresponding to the required properties of each
mode. Apart from analysis and optimization of GSPs, a parametric model can also
be directly used in the model-based control systems without any remodeling efforts.
Thus, in this respect the central problem is how to develop an accurate parametric
model for the vibrations of GSPs. In the following, we discuss the challenges and
review the potentially useful previous studies to establish a parametric model of the
GSP vibrations.
To obtain such a parametric model, the kinematics and the derivation of the
linearized equations of motion as well as the modal analysis should be carried out
parametrically. Generally, the kinematics of GSPs is well known in the literature
(see e.g. [68]) where in particular a parametric formulation of the Jacobian matrix
at the neutral configuration is developed by the present author [9]. In the following,
relevant studies in the literature on stiffness and inertia matrices as well as modal
analyses are indicated, respectively.
The linearized equations of motion are identified using stiffness, damping and
inertia (mass) matrices. The stiffness matrix is a function of the flexibility properties
of the struts as well as the GSP geometry. Due to the complexity in formulating the
stiffness matrix in the Cartesian-space, often a semi-parametric or a finite element
approach is used. The semi-parametric approach is referred to a formulation of
the stiffness matrix simply in terms of the Jacobian matrix without a parametric
representation in terms of the design variables [1013]. Formulations of the stiffness
matrix for some special conditions can also be found in [1416] and a parametric
formulation of the stiffness matrix in the Cartesian-space for the neutral configuration
is given in [9]. It should be noted that the consideration of axial Rayleigh damping
for the struts leads to damping matrices with the same structure as the stiffness
matrices [9, 17].
Due to the influence of the strut inertia, the derivation of the inertia matrix of the
system is quite challenging. To simplify the model, many researchers have ignored
the strut inertia and assumed the inertia matrix in the Cartesian-space simply as a
diagonal matrix containing only the inertia properties of the platform [1822]. The
inertia of rigid struts has been considered in some dynamic modelings (see e.g. [3, 23
25]) but due to the lack of flexibility in the system, these models cannot be employed
for a modal analysis. In some control models, the inertia of the struts is simply

82
assumed to be a diagonal matrix in the joint-space [2629]. Hence, generally in the
literature a parametric 6 6 equivalent inertia matrix of the system is not addressed.

Specifically regarding the modal analysis of GSPs, only a limited number of pa-
pers have been published which can be categorized into finite element [30], semi-
parametric [21, 3136] and parametric [9, 18, 32] approaches and none of them
parametrically takes the strut inertia into account. Among the semi-parametric ap-
proaches, Wang et al. [36] take the inertia of the struts into account for a hydraulic
GSP with a semi-analytical approach where the equivalent inertia matrix and the
eigenvalue problem are solved numerically. In an effort of constructing the equiva-
lent inertia matrix, Mahboubkhah et al. [32] expand the inertia matrix to a 24 24
matrix by considering three masses on each strut, while a parametric solution to the
eigenvalue problem is not addressed in this work. However, since the ultimate ob-
jective is to study the vibrational behavior of the platform, obtaining an equivalent
6 6 inertia matrix is much more efficient than introducing extra DOFs and adding
complications by a 24 24 inertia matrix.

Very few researchers have carried out parametric modal analyses of GSPs and
none of them has yet taken the inertia of the struts into consideration. Among them,
Jiang et al. [18] solved the eigenvalues of the inverse inertia matrix in the joint-space
and obtained the undamped vibrations parametrically. Tong et al. [22], formulated
the undamped eigenvalues for a class of decoupled modified GSPs. Previous work
by the present authors [9] developed a parametric modal analysis of the damped
vibrations. However, generally as indicated earlier neither a parametric modal anal-
ysis nor a parametric formulation of the inertia matrix is addressed in the literature
where strut inertia is taken into account.

To the best of our knowledge, this paper for the first time, presents a solution
to the above problems where we take the inertia of the struts into consideration
and present a fully parametric model of the GSP damped vibrations for the neutral
configuration. A parametric 6 6 equivalent inertia matrix in terms of the design
variables is also formulated where the inertia properties of the platform and the
struts are considered together. The equivalent inertia matrix is obtained through
a Lagrangian dynamics approach by introducing strut moment of inertia into the
system. Consequently, the modal analysis including the damped eigenfrequencies
and the eigenvectors are presented parametrically. Then, the decoupling conditions
are investigated analytically. In addition, a specific reference GSP is employed to
numerically demonstrate the represented analytical method and to systematically
evaluate the influence of the strut inertia on the system. This paper can also be
considered as the direct continuation of the previous work by the authors [9], in
which an improved vibrational model is developed. The present study serves as an
accurate parametric model of the vibrations that can be directly used in the analysis
and optimization of GSPs in any applications. In addition, it can also be employed
as an accurate mechanical model for model-based control systems.

83
2. Analytical damped vibrations
2.1. Parametric linearized equations of motion
In this section, the equations of motion are obtained and are linearized about
the neutral configuration. Neutral configuration is where the structure is in its ini-
tially symmetric configuration in a static equilibrium. It should be noted that the
neutral configuration can be any symmetric configuration with respect to the z-axis.
Considering identical properties for all the struts, the design space of a GSP neutral
configuration is
cp cp cp
, k, c R14
 
Ds = d, mst , Ist , mp , Ixx , Iyy , Izz

B cp cp cp

d Gs , mst , Ist , mp , Ixx , Iyy , Izz , k, c > 0 (1)
where the parameters are strut mass mst , strut moment of inertia Ist , platform
c c c
mass mp , platform moments of inertia Ixxp , Iyyp , Izzp , strut stiffness k and strut viscous
damping c as well as the geometrical design space Gs which is given by (see Figs. 1
and 2)

n o
(Rb , R , H , h , b , p ) R6 Rb > 0, R > 0, H > 0, b , p 0, <

Gs = ,
3

Rp H h
R = , H = , h = , = b + p (2)
Rb Rb Rb
Initially as a generic approach, the struts are considered as flexible bodies without
identical properties. To define strut moment of inertia, firstly, the material element
dmst,i is introduced (see Fig. 1) in which its position vector is given by
s
r i = r i (s, t) = li (t), 0 s li,0 , i = 1, ..., 6 (3)
li,0
where li = r Bi Pi , li,0 is the length of each strut at the neutral configuration. Equation
(3) is a realistic assumption for the transplacement of the struts, especially in high
stiffness applications where stiffness uniformity of the struts is a design requirement.
Consequently, the time derivative of the position vector may be written
s
r i = li (t), (4)
li,0
The mass element can also be written

dmst,i = dmst,i (s) = i (s)ds (5)


where i = i (s), 0 s li,0 is the mass density of the struts (mass per unit length).
The total mass of each strut is then given by
li,0
mst,i = i (s)ds (6)
0

84
Fig1 in paper 2

c c c
Platform: m p , I xxp , I yyp , I zzp
z y
ep ep
x
c p , Op ep
h p1
P4 P3 P2
P1
Op ' p'1
1

P5 P6

l3 t l2 l1,0
H l4 l1

l5 l6
B3 B2

z
dmst ,1
y
eb eb s
b1 r1
Ob x
eb
B4 Base B1
B5 B6

Figure 1: Parametrization of a general GSP

Consequently, the total kinetic energy of each strut is


li,0 li,0 2 B
Ist,i

1 1 s 2
Tst,i |T otal = r 2i dmst,i = li i (s)ds = 2 li (7)
2 0 2 0 li,0 2li,0
B
where Ist,i (the strut moment of inertia about the base joints Bi ) is defined by
li,0
B
Ist,i = s2 i (s)ds (8)
0

A combination of universal and spherical joints for the base and the platform
respectively, is the most common design of existing GSPs [37]. This combination is
recognized as a 6-Universal-Prismatic-Spherical (6-UPS) architecture which has the
minimum possible kinetic energy. Since the universal joints are fixed to the base,
they restrict the rotation of each strut about its axis. Hence, the moment of inertia
about the axis of each strut does not play a role.
There are two approaches for the calculation of the right hand side of Eq. (7)
depending on using the joint-space or the Cartesian-space coordinates. First, one
can write the kinetic energy of the struts in the joint-space by
2 2
li = li ei + li st,i ei li = li2 + li2 ( st,i ei ) (9)
where ei and st,i are

85
Fig2 in paper 1

eby
B2

B3
p
P2
P1

P3 B1
b
1200
P4
ebx
Rp B6
P6

P5
Rb
B4

B5

Figure 2: Top view of the general GSP

li
ei = ,
li

ei = st,i ei (10)
and the length of each strut li forms the joint-space coordinates as follows
T
qJ = (l1 ... l6 ) (11)
Hence, according to Eq. (9) the kinetic energy of each strut can be decomposed into
axial and rotational kinetic energies
B
Ist,i IB
Tst,i |T otal = Tst,i |Axi. + Tst,i |Rot. = 2 + st,i ( st,i li )2
l (12)
2 i
2li,0 2
2li,0
Second, by the employment of the following kinematic equation

li = t + pi bi , i = 1, ..., 6 (13)
where t = r Ob Op , pi = r Op Pi , bi = r Ob Bi , one can set up li in terms of the angular
velocity of the platform as follows
2 2 2
li = t + pi = t + pi li = t + 2t pi + ( pi ) (14)
The matrix representation of in the platform-frame is defined by

86
T
[]e = p,x p,y p,z (15)
p

where ep = exp , eyp , ezp with Op as the origin form the platform-frame (see Fig. 1).
[]e in terms of Bryant angles is given in Appendix 2.
p
Having compared the two above approaches, due to the complexity of setting up a
parametric st,i in the joint-space, the second approach seems to be more efficient for
the purpose of obtaining parametric equations of motion. Thus, the Cartesian-space
approach is used for the rest of the formulations in this paper.
The total kinetic energy of the system including the kinetic energy of the platform
and the struts, is given by
6 
X 
T = Tp + Tst,i |T otal (16)
i=1

Then, combining Eqs. (7) and (14) leads to

6  6
( )
B h
X  1X Ist,i 2 2
i
Tst = Tst,i |T otal = 2 t + 2t pi + ( pi ) (17)
i=1
2 i=1 li,0

where pi can be decomposed into two mutually orthogonal vectors according to


pi = p0i + h (see Fig. 1, 2) where p0i = r Op0 Pi and h = r Op Op0 that can also be
expressed as

h = hezp = hRezb = h (R13 exb + R23 eyb + R33 ezb ) (18)


where eb = (exb , eyb , ezb )
with Ob as the origin for the base-frame and the rotation
tensor R, based on the Bryant angles, is given by

R11 R12 R13
R = eb [R]e , [R]e = R21 R22 R23 (19)
b b
R31 R32 R33
The rotation matrix [R]e in terms of Bryant angles is presented in Appendix 2.
b
Now we return to the identical assumption for the struts where we assume that all
the struts have identical properties at the neutral configuration. Assuming li,0 = l0
B B
and Ist,i = Ist for the neutral configuration, see Eq. (53) in Appendix-2, then

( 6 6
! 6
)
IB 2
X X X 2
Tst = st2 6t + 2t p0i hRezb + [ (p0i hRezb )] (20)
2l0 i=1 i=1 i=1

P6
Note that i=1 p0i = 0 due to the tri-axially symmetric pattern of p0i . Furthermore,
by the small-angle assumption, Eq. (20) can be simplified in terms of the Cartesian-
space coordinates as follows

87
B 
Ist
6 t2x + t2y + t2z + 12h ty tx + C
 
Tst = (21)
2l02
T P6 2
where [t]e = (tx , ty , tz ) and C = i=1 [ (p0i hRezb )] . The Cartesian-space
b
coordinates, including the position of the platforms center of mass and three Bryant
angles, are defined as
T
qC = tx ty tz (22)
Now, to represent C in terms of the Cartesian-space coordinates, it can be written
6 h i
X 2 2
C= ( p0i ) + hezp 2h p0i ezp (23)
i=1
P6 P 
0 z
 6 0 z
where i=1 2h pi e p = 2h i=1 i ep = 0 so the double
p
product disappears and then
6
X 2
C= p,x exp p0i + p,y eyp p0i + p,z ezp p0i +
i=1
6
X 2
h2 p,x exp ezp + p,y eyp ezp (24)
i=1

Further simplifications due to the orthogonality properties lead to

6 h
X
2
  i
Rp2 sin2 i + p,y
2
Rp2 sin2 i + 2
Rp2 + 6h2 p,x
2 2

C= p,x + p,z + p,y =
i=1
2

6   6  
2
X 1 cos2i X 1 cos (2i + )
Rp2 2
Rp2 2
Rp2 +6h2 p,x
2 2

p,x +p,y +6p,z + p,y
i=1
2 i=1
2
(25)
where i is the angle between the exp and p0i on the platforms joint-plane (see Figs.
1, 2). Due to the rotational-symmetry property of the platform joints, one may prove
6
X 6
X
cos2i = 0, cos (2i + ) = 0 (26)
i=1 i=1

The proof of Eq. (26) is given in Appendix 1. Hence, C is obtained as

C = 6h2 + 3Rp2 p,x


 2
+ 6h2 + 3Rp2 p,y
 2
+ 6Rp2 p,z
2
(27)
Finally, the simplified total kinetic energy of the system, using the small-angle as-
sumption, in terms of Cartesian-space coordinates is obtained as

88
B
IB
   
1 Ist cp
t2x + t2y + t2z + Ixx + st 2 2
  2
T = mp + 6 6h + 3Rp +
2 l02 l02

B B B
    
cp Ist 2 2
 2 cp Ist 2 2 hIst 
Iyy + 2 6h + 3Rp + Izz + 6 2 Rp + 12 2 ty tx (28)
l0 l0 l0
Therefore, it is shown that it is possible to calculate the kinetic energy of the
struts directly from Eq. (7) and not necessarily by the decomposed format in Eq.
(12). In other words, in the Cartesian-space approach there is no need to formulate
the angular velocities and the elongation rates of each strut to obtain a parametric
formulation of the total kinetic energy of the struts. Therefore, the presented formu-
lation is perhaps the most efficient approach for an algebraic representation of this
problem.
The linearized equations of motion of the system is given by MC qC + CC qC +
KC qC = 061 in the Cartesian-space. According to Eq. (28), the equivalent inertia
matrix of the system in the Cartesian-space is formulated as


M 0 0 0 Mx 0
M 0 My 0 0
2

T M 0 0 0
MC = 2 = M C,p + M C,st =
qC qC,0 ,qC,0

M 0 0

sym. M 0
M
(29)
where the inertia matrix of the platform M C,p , in the Cartesian-space, is given by

2 Tp

c c c 
M C,p = 2 = diag mp mp mp Ixxp Iyyp Izzp (30)
qC qC,0 ,qC,0

and the inertia matrix of the struts, in terms of dimensionless variables defined in
Eq. (2), is obtained as

2 Tst

M C,st = 2 =
qC qC,0 ,qC,0

B B
6Ist 6Ist Rb h

l02
0 0 0 l02
0
B B
6Ist 6Ist Rb h


l02
0 l02
0 0

B
6Ist


l02
0 0 0
(31)

B 2
Ist Rb 

l02
62h + 32R 0 0

B 2
Ist Rb 

sym. l02
62h + 32R 0

B
6Ist
l02
Rb2 2R

89
These results interestingly show that despite the apparent symmetrical structure of
GSPs, when taking inertia of the struts into account, the inertia matrix turns out to
be non-diagonal. Furthermore, Eq. (31) can be directly used for lumped masses if
i (s) is defined as
n
X
i (s) = mji (s) (s sj ) (32)
j=1

where (s sj ) is the Dirac delta function. The position vector of the lumped mass
mji (the jth lumped mass on strut number i) is

r ji = r ji (j , t) = j li (t), 0 j 1, i = 1, ..., 6, j = 1, ..., n (33)


where
sj
j = j (sj ) = , 0 sj l0 (34)
l0
Therefore, assuming all struts with identical properties (mji = mj ) where each strut
with j number of lumped masses, the moment of inertia for the struts can be simply
calculated by
n
X l0 n
X
B 2
Ist = s mj (s sj ) ds = l02 j2 mj (35)
j=1 0 j=1

It is obvious that the stiffness and damping matrices are not influenced by adding
the strut inertia into the system. Thus, the stiffness matrix in the Cartesian-space is
that one developed in [9], as follows

kxx 0 0 0 kx 0

kyy 0 ky 0 0
kzz 0 0 0
KC = kJ T J =

(36)

k 0 0
sym. k 0
k
where kxx , kyy ... etc. are parametrically given in Appendix 2. Furthermore, the
damping matrix is [9]
c
CC = cJ T J = KC , = (37)
k

90
2.2. Parametric modal analysis

To parametrically obtain the damped eigenfrequencies of the system, first the


eigenvectors are formulated using the undamped eigenvalue equations

MC 1 KC i2 I66 Xi = 061


where I66 is the 6 6 identity matrix and i are the undamped eigenfrequencies.
Parametrically obtained eigenvectors Xi form the modal matrix


X = X1 X2 X3 X4 X5 X6

as follows


0 0 x,1 0 x,2 0
0
0 0 y,1 0 y,2

0 1 0 0 0 0
X=
0
(38)
0 0 1 0 1
0 0 1 0 1 0
1 0 0 0 0 0

where the components are

M kxx M k x


x,1 = 2M kx 2Mx kxx




M kyy M k y


y,1 =

2M ky 2My kyy

(39)
M kxx M k +x


x,2 = 2M kx 2Mx kxx






M kyy M k +y
y,2 = 2M ky 2My kyy

and

q
2

x
= (M kxx M k ) + 4 (M kx Mx k ) (M kx Mx kxx )

q
= (M k M k )2 + 4 (M k M k ) (M k M k )

y yy y y y y yy
(40)
The six undamped eigenfrequencies corresponding to the eigenvectors in Eq. (38)
are formulated as

91
q q
k kzz


= M , z = M





r
M k +M kxx 2Mx kx x



x,1 = 2
2M M 2Mx





q
M k +M kyy 2My ky y
y,1 = 2
2M M 2My
(41)





r
M k +M kxx 2Mx kx +x

=

x,2
2


2M M 2Mx




q
y,2 = M k +M kyy 2My2 ky +y


2M M 2M y

Since the inertia and stiffness matrices are positive definite, it follows that all the
expressions appearing under the root-sign in Eq. (41) have to be positive. Eventually
the six damped eigenfrequencies due to the axial Rayleigh damping of the struts
can be obtained according to [9]. Hence, considering weak damping, the damped
eigenfrequencies are given by
r
2 4
d = 2 (42)
4
where is given in Eq. (37).
According to Eq. (38), the six modes presented in Eq. (41) are the torsional, axial
and four tilting-lateral modes, respectively. The first two modes are straightforward
but the four latter modes are more difficult to recognize. A tilting-lateral mode
may be tilting-dominant or lateral-dominant based on the values of || in Eq. (38).
There exists tilting-dominant modes and lateral-dominant modes when ||  1 and
||  1, respectively.

2.3. Decoupled vibrations


By introducing the inertia of the struts (Mx = My 6= 0), the conditions for
the decoupled vibrations are also influenced. The decoupled vibrations are defined by
the existence of a fully decoupled modal matrix. This means six pure modes corre-
sponding to each one of the Cartesian-space coordinates. To obtain such a decoupled
modal matrix, eigenvector components given in Eq. (39) have to be either zero or
infinity. It means that when both x,1 and y,1 (which are always orthonormal)
are infinitely large, then both x,2 and y,2 are required to have a zero value or
vice versa, see Eqs. (44) and (45). It then follows that the denominators in Eq. (39)
have to be zero (2M kx 2Mx kxx = 0 and 2M ky 2My kyy = 0). Hence, the
necessary condition in order to achieve a fully decoupled modal matrix is

kxx kyy kx ky
= = = , Mx = My 6= 0 (43)
M M Mx My

92
By using the necessary decoupling condition specified in Eq. (43) in Eq. (38),
the components of the eigenvectors are obtained as follows

   

lim kx = lim ky y,1 =

Mx My

x,1

M kxx
M kyy

  M kxx <M k   M kyy <M k

lim Mx kx x,2 = lim My ky y,2 =


kxx

M M kyy
M kxx >M k M kyy >M k
(44)
While, the other two eigenvectors are given by

   

lim kx = lim kx =

Mx Mx

x,1 x,2
M kxx M kxx


M kxx >M k M kxx <M k

2M k 2M k

M kx Mx 6= 0

xx

   

lim My ky y,1 = lim My ky y,2 =

M kyy M kyy
M kyy >M k M kyy <M k

2M ky 2My k

6= 0

M kyy M k
(45)
The obtained results prove that only two eigenvectors can be decoupled and the
other two remain coupled. This results in either two pure lateral modes or two pure
tilting modes depending on the conditions specified in Eqs. (44) and (45). This
semi-decoupled configuration occurs if and only if Eq. (43) is satisfied, so it can be
B B
considered as a semi-decoupling condition. There exists Ist = Ist that satisfies the
semi-decoupling condition specified in Eq.(43)

B l2 k
0 x 0 y l2 k
Ist = 6(kx +k xx Rb h )
mp = 6(ky +kyy Rb h )
mp , h 6= 0 (46)

According to Eqs. (44) and (45), depending on M kxx < M k and M kxx >
M k the corresponding semi-decoupled modal matrices after normalization are ob-
tained as

0 0 1 0 x,2 0
0
0 0 1 0 y,2

0 1 0 0 0 0
X semidecoupled,1 =
0
(47)
0 0 0 0 1
0 0 0 0 1 0
1 0 0 0 0 0

and

93

0 0 x,1 0 1 0
0
0 0 y,1 0 1
semidecoupled,2
0 1 0 0 0 0
X =
0
(48)
0 0 0 0 0
0 0 0 0 0 0
1 0 0 0 0 0
B B
respectively, so in both cases when Ist = Ist , there exist two pure lateral modes.
Therefore, it can be concluded that it is impossible to achieve a pure decoupled
modal matrix when the inertia matrix is not diagonal, unless there exist multiple
eigenfrequencies corresponding to all four tilting-lateral modes. One can regard this
multiple-eigenfrequency condition as a special case of decoupled vibrations in which
the mode shapes can be any linear combination of the corresponding eigenvectors. To
obtain these four multiple eigenfrequencies the following condition must be satisfied

kxx kyy k k kx ky
= = = = = (49)
M M M M Mx My
c c
and it immediately implies Ixxp = Iyyp .
In addition, if the inertia matrix is diagonal (Mx = My = 0) the stiffness
matrix is also required to be diagonal (kx = ky = 0) in order to achieve decoupled
B
vibrations. Diagonality of the inertia matrix implies h = 0 and/or Ist = 0. Note
B
that for the first case (where h = 0 and kx = ky = 0), any arbitrary Ist leads to a
diagonal inertia matrix and consequently causes decoupled vibrations. Conditions for
decoupled vibrations in the case in which the inertia matrix is diagonal are studied
in [9].

3. Numerical evaluation

In this section, with the aid of a reference GSP in which inertia of the struts is
introduced, the developed analytical method is numerically demonstrated. The iner-
tia, stiffness and damping matrices as well as the eigenfrequencies and eigenvectors
of the system are numerically obtained based on the equations presented in section
2. In addition, the modal behavior of the GSP is investigated with respect to all
possible combinations of geometrical variables and strut inertia. Table 1 shows the
characteristics of this reference GSP.
The reference GSP is similar to the one studied in [9] but here inertia of the struts
is added in order to evaluate its influence on the vibrational behavior. Selected data
for the reference GSP are supposed to be relevant for a collimation system of a large
c c
optical telescope. It is also assumed that Ixxp 6= Iyyp to make the results more visible
by having distinct eigenfrequencies. According to Eqs. (36) and (37) the stiffness
matrix and the damping matrix are exactly those calculated in [9]

94
Table 1: Characteristics of the reference GSP
Vari. Description Value Unit

b Half angel between two joints on base 10 180 rad

p Half angel between two joints on platform 15 180 rad
k Axial stiffness of the struts 106 N/m
c Axial damping of the struts 103 Ns/m
mp Mass of the platform 102 kg
c
Ixxp Moment of inertia of the platform around x 7 kgm2
c
Iyyp Moment of inertia of the platform around y 5 kgm2
c
Izzp Moment of inertia of the platform around z 10 kgm2
Rb Radius of the base joints circle 0.5 m
R Radius ratio (R = Rp/Rb ) 0.8 -
H Height ratio (H = H/Rb ) 1.8 -
h Mass-center height-ratio (h = h/Rb ) 0.2 -
B
Ist Strut moment of inertia about base joints 10 kgm2


276820 0 0 0 16092 0

276820 0 16092 0 0

5446360 0 0 0
KC = 1 CC =



436159 0 0

sym. 436159 0
88484

where = 103 s. However, by introducing the strut inertia (Ist B


= 10 kgm2 ) the
inertia matrix of the system is dramatically influenced. The inertia matrix of the
system in the Cartesian-space is obtained by M C = M C,p + M C,st where according
to Eqs. (30) and (31) the inertia matrix of the platform and the struts are respectively
obtained as

M C,p = diag 102 102 102



7 5 10
and

67.239 0 0 0 6.724 0

67.239 6.724 0 0

67.239 0 0 0
M C,st =

6.052 0 0

sym. 6.052 0
10.759
B
Figure 3(a) shows a continuous decrease of all the eigenfrequencies when Ist is
increased. As indicated earlier it is clearly visible that the eigenfrequencies of the
c c
x, 2 and y, 2-modes are not identical due to the asymmetric platform (Ixxp 6= Iyyp ).

95
y,2-mode x,2-mode y,1-mode x,1-mode z-mode -mode
(a) (b)
300 200

250
100
200
d

150 0


100
-100
50

0 -200
0 10 20 30 40 50 0 10 20 30 40 50
IBst
2
(kgm ) IBst (kgm2)

Figure 3: (a) The six damped eigenfrequencies of the reference GSP with respect to the strut inertia
(b) Eigenvectors components of the four coupled modes with respect to the strut inertia;
B which lead to an infinite value of and o shows the value of I B which leads
shows the value of Ist st
to a zero value of .

B
If Ist is extremely larger than the platforms inertia, the inertia matrix is dominated
by the inertia of the struts. Hence, as the inertia matrix is dominated by the inertia
of the struts, these two distinct eigenfrequencies come close due to the symmetric
c c c
structure of the struts. It should also be noted that Ixxp , Iyyp and Izzp are principal
B
moments of inertia, while Ist is the moment of inertia about the base points Bi
which are not the mass center of the struts, so they cannot be directly compared
with respect to their magnitude.
B
The corresponding eigenvectors are plotted with respect to Ist in Fig. 3(b). Only
B
four coupled eigenvectors are influenced by Ist , so the eigenvectors are presented by
B
their eigenvector components given in Eq.(39). It can be observed that at Ist
according to Eq.(46), x,1 and y,1 go to infinity, while x,2 and y,2 do not go to
zero (as is analytically proven in section 2.3). Thus, as expected the reference GSP
B
has two pure lateral modes at Ist = 20.65.
To study how the vibrational behavior changes by introducing strut inertia in the
system, a systematic evaluation is presented in Figs. 4 and 5. In Fig. 4, all the six
damped eigenfrequencies according to Eq. (41) are plotted versus the geometrical
B
variables for three different values of Ist . The first column of the plots in Fig. 4
B
shows the vibrational behavior when there is no mass assigned to the struts Ist = 0,
B B
while the second and third columns are for Ist = 10/6 and Ist = 10, respectively.
The three plots in Fig. 4(a) demonstrate the influence of the radius ratio R on all
the six eigenfrequencies of the system. Similarly, the other geometrical variables are
studied in Figs. 4(b), (c), (d) and (e).
As it is expected all the eigenfrequencies in Fig. 4 decreases with the increase
c c
of strut inertia. Although Ixxp 6= Iyyp , nearly all the x, 1 and y, 1-modes are quite
close in all the plots in Fig. 4. This is due to the fact that, according to Fig. 5, x, 1
and y, 1-modes are mainly lateral dominant modes. In lateral-dominant modes
the eigenfrequencies are not much sensitive to the platforms moment of inertia.
However, tilting-dominant modes are quite sensitive to the platforms moments of

96
y,2-mode x,2-mode y,1-mode x,1-mode z-mode -mode

(a-1) (a-2) (a-3)


800 800 800

600 600 600


d

d
400 400 400

200 200 200

0 0 0
0 1 2 3 0 1 2 3 0 1 2 3
R=Rp/Rb R=Rp/Rb R=Rp/Rb
(b-1) (b-2) (b-3)
300 300 300

200 200 200


d

d
100 100 100

0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
H=H/Rb H=H/Rb H=H/Rb
(c-1) (c-2) (c-3)

300 300
800
d

400 150 150

0 0 0
-20 -10 0 10 20 -20 -10 0 10 20 -20 -10 0 10 20
h=h/Rb h=h/Rb h=h/Rb
(d-1) (d-2) (d-3)



=b+p (rad) =b+p (rad) =b+p (rad)
(e-1) (e-2) (e-3)
800 800 800

600 600 600


d

400 400 400

200 200 200

0 0 0
0 1 2 3 0 1 2 3 0 1 2 3
Rb (m) Rb (m) Rb (m)

Figure 4: The six damped eigenfrequencies of the reference GSP versus the geometrical variables
of the system; (a), (b), (c), (d) and (e) show the influence of the geometrical variables including
R , H , h , and Rb , respectively; all the variables are studied for three different strut moment of
inertia which are denoted by: (1) Ist B = 0, (2) I B = 10/6, (3) I B = 10.
st st

97
x,2 y,2 y,1 x,1

(a-1) (a-2) (a-3)


5 5 5

1 1 1


-1 -1 -1

-5 -5 -5
0 1 2 0 1 2 0 1 2
R=Rp/Rb R=Rp/Rb R=Rp/Rb
(b-1) (b-2) (b-3)
8 8 8

1 1 1


-1 -1 -1

-8 -8 -8
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
H=H/Rb H=H/Rb H=H/Rb
(c-1) (c-2) (c-3)
5 5 5

1 1 1

-1 -1 -1

-5 -5 -5
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
h=h/Rb h=h/Rb h=h/Rb
(d-1) (d-2) (d-3)



=b+p (rad) =b+p (rad) =b+p (rad)
(e-1) (e-2) (e-3)
5 5 5

1 1 1

-1 -1 -1

-5 -5 -5
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
Rb (m) Rb (m) Rb (m)

Figure 5: The corresponding eigenvectors components according to Fig. 4; all the variables are
B = 0, (2) I B = 10/6,
plotted for three different strut moment of inertia which are denoted by: (1) Ist st
B
(3) Ist = 10; shows the poles which leads to an infinity and o shows the zeros which leads
to a zero .

98
c c
inertia. Accordingly, when Ixxp 6= Iyyp as it is in this example, x, 2 and y, 2-modes
are clearly distinct in the plots.
The behavior of the x, 2 and y, 2-modes in Fig. 4(c) shows how the lack
of strut inertia can lead to unrealistic results. In this example, when there is no
mass assigned to the struts, the x, 2 and y, 2-modes have minimum eigenfrequency
values at h = 0.084 in Fig. 4(c-1). However, as strut inertia increases a totally
different behavior is observed in Fig. 4(c-2) and 4(c-3) in which there exists maximum
eigenfrequency values at h = 0.094 and h = 0.14, respectively. Note that the scales
of the d axis are different in the plots of Fig. 4(c). This dramatic change is due
to the fact that when strut inertia is introduced to the system, the inertia matrix
become dependent on h .
It should be noted that in Fig. 4(d) the variation of ( = b +p ) is studied. This
is due to the fact that only affects the results and not b and p individually, see Eqs.
(52) and (53). In Figs. 4(e-2) and 4(e-3), it is observed that all the eigenfrequencies
approach zero when Rb approaches zero. This is not the case in Fig. 4(e-1) where
there are three non-zero eigenfrequencies at Rb = 0. In fact, the decrease of Rb , scales
down the whole geometry without any change in the aspect ratios. When Rb = 0,
consequently all the dimensions are also zero including the initial lengths l0 . Hence,
this leads to an inertia matrix of the struts with infinitely large components which
implies zero eigenfrequencies when Rb = 0, see Figs. 4(e-2) and 4(e-3).
The corresponding eigenvectors of Fig. 4 are presented in Fig. 5 where only the
four coupled modes which can be affected by the geometrical variables are shown. In
the plots in Fig. 5, if the poles of the eigenvector components coincide with the zeros,
a complete decoupled vibration is accomplished (a pole corresponds to an infinity
which is marked by and a zero corresponds to a zero which is marked by o).
This is the case in the first column of the plots where the strut inertia is not taken
into account. According to [9], we also know that this case corresponds to a diagonal
stiffness matrix and is completely decoupled.
The analytically derived conclusion in section 2.3 can be observed in the second
and third column of Fig. 5, where it is shown that there is no coinciding zero and pole.
This result shows that there exists no configuration in which one can obtain complete
decoupled eigenvectors. As a clear example, Figs. 5(a-2) and 5(a-3) show how the
zeros depart from the poles while the coinciding zeros and poles can be seen in Fig.
5(a-1). Poles of the system are obtained by Eq. (43) which lead to two decoupled
modes but zeros of the system may only lead to one decoupled mode. Generally, as
it is observed in Figs. 4 and 5, not taking the strut inertia into consideration may
lead to unrealistic results. Therefore, it is crucial to consider the struts in order to
obtain a more accurate and realistic evaluation of a GSP and one may directly use
the equations developed in this paper for this purpose.

4. Conclusion

In this paper, for the fist time, a parametric model of damped vibrations for
GSPs are developed in which strut inertia is taken in to account. Based on the

99
stiffness-uniformity assumption, the total kinetic energy of the struts is developed in
the Cartesian-space. A well-structured formulation of the kinetic energy is presented
parametrically in terms of the design variables. It is concluded that the kinetic energy
of the system can be formulated easier in terms of the Cartesian-space coordinates
as compared with the joint-space coordinates, due to the fact that in the Cartesian-
space approach the formulation of the angular velocity of the struts can be avoided.
Accordingly, based on Lagrangian dynamics, the linearized equations of motion are
given in the form of parametric inertia, stiffness and damping matrices.
When considering inertia for the struts, the inertia matrix of the system is found
to be a non-diagonal matrix in the Cartesian-space. Obviously the stiffness and the
damping matrix of the system are not affected by this consideration but due to the
structural change of the inertia matrix, the solution of the eigenvalue problem is
entirely changed. Hence, the modal matrix and accordingly the damped eigenfre-
quencies are analytically solved and presented. In addition, a representation of the
obtained solution when the struts are considered as lumped masses is given.
The possibility of obtaining decoupled vibrations is analytically investigated. It is
concluded that it is impossible to achieve complete decoupled vibrations when taking
the inertia of the struts into account, unless the necessary condition for multiple
eigenfrequencies is satisfied. Only semi-decoupled vibrations are feasible and the
condition for this is formulated. It is proven that there exists a unique value of
B
Ist where the semi-decoupled vibrations are feasible with two pure lateral modes.
However, if the struts are considered mass-less, the complete decoupled vibrations
are accomplished if and only if the stiffness matrix is diagonal.
Finally, the analytically formulated results are numerically demonstrated. The
vibrational behavior of the system corresponding to all possible combinations of geo-
metrical variables and strut inertia is illustrated and discussed. The results indicate
that not considering the inertia of the struts can lead to dramatic errors in some
eigenfrequencies and eigenvectors. Therefore, the parametric model developed in this
work can serve as the first analytical reference to this problem, which can be directly
employed by engineers and researchers in order to obtain more realistic evaluations
of GSPs for any application.

References

[1] B. Dasgupta, T. Mruthyunjaya, The Stewart platform manipulator: a review,


Mech. Mach. Theory 35 (1) (2000) 1540. doi:10.1016/S0094-114X(99)00006-3.
[2] Y. D. Patel, Parallel Manipulators Applications-A Survey, Mod. Mech. Eng.
02 (03) (2012) 5764. doi:10.4236/mme.2012.23008.
[3] Y. Ting, Y.-S. Chen, H.-C. Jar, Modeling and control for a Gough-
Stewart platform CNC machine, J. Robot. Syst. 21 (11) (2004) 609623.
doi:10.1002/rob.20039.
[4] J. J. Zierer, J. H. Beno, D. a. Weeks, I. M. Soukup, J. M. Good, J. a. Booth,
G. J. Hill, M. D. Rafal, Design, testing, and installation of a high-precision

100
hexapod for the Hobby-Eberly Telescope dark energy experiment (HETDEX),
in: L. M. Stepp, R. Gilmozzi, H. J. Hall (Eds.), Proc. SPIE, Vol. 8444, 2012,
pp. 84444O84444O12. doi:10.1117/12.926394.
[5] M. Wapler, V. Urban, T. Weisener, J. Stallkamp, M. Drr, A. Hiller, A Stewart
platform for precision surgery, Trans. Inst. Meas. Control 25 (4) (2003) 329334.
doi:10.1191/0142331203tm092oa.
[6] K. Liu, F. Lewis, G. Lebret, D. Taylor, The singularities and dynamics of a
Stewart platform manipulator, J. Intell. Robot. Syst. 8 (3) (1993) 287308.
doi:10.1007/BF01257946.
[7] S. Bandyopadhyay, A. Ghosal, An algebraic formulation of kinematic isotropy
and design of isotropic 6-6 Stewart platform manipulators, Mech. Mach. Theory
43 (5) (2008) 591616. doi:10.1016/j.mechmachtheory.2007.05.003.
[8] B. M. St-Onge, Singularity Analysis and Representation of the Gen-
eral Gough-Stewart Platform, Int. J. Rob. Res. 19 (3) (2000) 271288.
doi:10.1177/02783640022066860.
[9] B. Afzali-Far, P. Lidstrm, K. Nilsson, Damped vibrations of Gough-Stewart
Platforms, (Technical Rep. ISRN LUTFD2/TFME-13/1014-SE (1-21), Div.
Mech. Lund Univ. 2013, Submitt. to Mech. Mach. Theory).
[10] C. Gosselin, Stiffness mapping for parallel manipulators, Robot. Autom. IEEE
Trans. 6 (3) (1990) 377382.
[11] B. El-Khasawneh, P. Ferreira, Computation of stiffness and stiffness bounds for
parallel link manipulators, Int. J. Mach. Tools . . . 39 (1999) 321342.
[12] S. Muruganandam, S. Pugazhenthi, Stiffness-based workspace atlas of hexapod
machine tool for optimal work piece location, Int. J. Adv. Manuf. Technol. 42 (1-
2) (2008) 202210. doi:10.1007/s00170-008-1573-7.
[13] C. Huang, W.-h. Hung, I. Kao, New conservative stiffness mapping
for the Stewart-Gough platform, in: Proc. 2002 IEEE Int. Conf.
Robot. Autom. (Cat. No.02CH37292), Vol. 1, Ieee, 2002, pp. 823828.
doi:10.1109/ROBOT.2002.1013459.
[14] J. Chen, F. Lan, Instantaneous stiffness analysis and simulation for
hexapod machines, Simul. Model. Pract. Theory 16 (4) (2008) 419428.
doi:10.1016/j.simpat.2008.01.003.
[15] V. Portman, V. Chapsky, Y. Shneor, Workspace of parallel kinematics machines
with minimum stiffness limits: Collinear stiffness value based approach, Mech.
Mach. Theory 49 (2012) 6786. doi:10.1016/j.mechmachtheory.2011.11.002.
[16] J. M. Prajapati, L. N. Patel, Dynamic (Forward and Inverse Force Transmission
Capability) Analysis of the Stewart Platform as Robot Manipulator, 2007 Int.
Conf. Mechatronics Autom. (2007) 28482853doi:10.1109/ICMA.2007.4304011.

101
[17] B. M. St-Onge, Singularity Analysis and Representation of the Gen-
eral Gough-Stewart Platform, Int. J. Rob. Res. 19 (3) (2000) 271288.
doi:10.1177/02783640022066860.

[18] H.-Z. Jiang, J.-F. He, Z.-Z. Tong, Characteristics analysis of joint
space inverse mass matrix for the optimal design of a 6-DOF par-
allel manipulator, Mech. Mach. Theory 45 (5) (2010) 722739.
doi:10.1016/j.mechmachtheory.2009.12.003.

[19] J.-F. He, H.-Z. Jiang, Z.-Z. Tong, B.-P. Li, J.-W. Han, Study on
dynamic isotropy of a class of symmetric spatial parallel mechanisms
with actuation redundancy, J. Vib. Control 18 (8) (2011) 11561164.
doi:10.1177/1077546311409608.

[20] a. Preumont, M. Horodinca, I. Romanescu, B. de Marneffe, M. Avraam, A. De-


raemaeker, F. Bossens, A. Abu Hanieh, A six-axis single-stage active vibration
isolator based on Stewart platform, J. Sound Vib. 300 (3-5) (2007) 644661.
doi:10.1016/j.jsv.2006.07.050.

[21] S. Pedrammehr, M. Mahboubkhah, N. Khani, A study on vibration of Stewart


platform-based machine tool table, Int. J. Adv. Manuf. Technol. 65 (5-8) (2012)
9911007. doi:10.1007/s00170-012-4234-9.

[22] Z. Tong, J. He, H. Jiang, G. Duan, Optimal design of a class of


generalized symmetric Gough-Stewart parallel manipulators with dynamic
isotropy and singularity-free workspace, Robotica 30 (02) (2011) 305314.
doi:10.1017/S0263574711000531.

[23] H. Abdellatif, B. Heimann, Computational efficient inverse dynamics of 6-DOF


fully parallel manipulators by using the Lagrangian formalism, Mech. Mach.
Theory 44 (1) (2009) 192207. doi:10.1016/j.mechmachtheory.2008.02.003.

[24] K. Harib, K. Srinivasan, Kinematic and dynamic analysis of Stewart


platform-based machine tool structures, Robotica 21 (5) (2003) 541554.
doi:10.1017/S0263574703005046.

[25] Z. Geng, L. S. Haynes, J. D. Lee, R. L. Carroll, On the dynamic model and


kinematic analysis of a class of Stewart platforms, Rob. Auton. Syst. 9 (4) (1992)
237254. doi:10.1016/0921-8890(92)90041-V.

[26] T. Yang, J. Ma, Z.-g. Hou, M. Tan, Robust backstepping control of active vibra-
tion isolation using a stewart platform, in: 2009 IEEE Int. Conf. Robot. Autom.,
Ieee, 2009, pp. 17881793. doi:10.1109/ROBOT.2009.5152486.

[27] H. Lin, J. E. McInroy, Disturbance attenuation in precise hexapod pointing


using positive force feedback, Control Eng. Pract. 14 (11) (2006) 13771386.
doi:10.1016/j.conengprac.2005.10.002.

102
[28] Y. Chen, J. McInroy, Decoupled control of flexure-jointed hexapods using esti-
mated joint-space mass-inertia matrix, Control Syst. Technol. IEEE . . . 12 (3)
(2004) 413421.
[29] J. McInroy, J. Hamann, Design and control of flexure jointed hexapods, IEEE
Trans. Robot. Autom. 16 (4) (2000) 372381. doi:10.1109/70.864229.

[30] M. Olivier, G. Duchini, Hexapod Mechanical Assembly modal survey test, in:
. . . Test. Sp. Program., 2004.
[31] H. Jiang, J. He, Z. Tong, Modal Space Control for a Hydraulically Driven Stew-
art Platform, J. Control Eng. Technol. 2 (July) (2012) 106115.
[32] M. Mahboubkhah, M. J. Nategh, S. Esmaeilzadeh Khadem, A comprehensive
study on the free vibration of machine tools hexapod table, Int. J. Adv. Manuf.
Technol. 40 (11-12) (2008) 12391251. doi:10.1007/s00170-008-1433-5.
[33] P. Mukherjee, B. Dasgupta, A. Mallik, Dynamic stability index and vibration
analysis of a flexible Stewart platform, J. Sound Vib. 307 (3-5) (2007) 495512.
doi:10.1016/j.jsv.2007.05.036.

[34] J. Selig, X. Ding, Theory of vibrations in Stewart platforms, in: . . . Syst. 2001.
Proceedings. 2001 IEEE . . . , 2001, pp. 21902195.
[35] M. Mahboubkhah, M. J. Nategh, S. Esmaeilzade Khadem, Vibration analysis
of machine tools hexapod table, Int. J. Adv. Manuf. Technol. 38 (11-12) (2007)
12361243. doi:10.1007/s00170-007-1183-9.

[36] W. Wang, H.-y. Yang, J. Zou, X.-d. Ruan, X. Fu, Optimal design of Stew-
art platforms based on expanding the control bandwidth while considering
the hydraulic system design, J. Zhejiang Univ. Sci. A 10 (1) (2009) 2230.
doi:10.1631/jzus.A0820329.

[37] J. P. Merlet, Parallel Robots, Parallel Robots, Springer, 2006.

103
Appendix 1

Eq. (26) can be proven based on the tri-axially symmetric distribution of the
joints on the platforms plane as follows, see Fig. (4)
6
X 4
cos2i = cos(2 2p ) + cos(2 + 2p ) + cos(2 + 2p )+
i=1
3

4 8 8
cos(2 + + 2p ) + cos(2 + 2p ) + cos(2 + + 2p ) =
3 3 3

4 8
2cos(2)cos(2p ) + 2cos(2 + )cos(2p ) + 2cos(2 + )cos(2p ) =
3 3


4 4 8
ig6 in paper 2
2cos(2 p ) cos(2) + cos(2)cos( ) sin(2)sin( ) + cos(2)cos( )
3 3 3

  
8 1 1
sin(2)sin( ) = 2cos(2p ) cos(2) cos(2) cos(2) = 0
3 2 2
P6
Hence, i=1 cos (2i + ) = 0 is also accordingly true.

y
ep
P2 p
p
P1

P3 0 1
120
x
0
ep
P4 120

P6

P5

Figure 6: Tri-axially symmetric distribution of the joints on the platforms plane where the platform
including the joints can rotate according to an arbitrary angle

104
Appendix 2

The rotation matrix based on Bryant angles: [9]


cos cos cos sin sin
[R]e = cos sin + sin sin cos cos cos sin sin sin sin cos
b
sin sin cos sin cos sin cos + cos sin sin cos cos
(50)

The angular velocity of the platform in the platform-frame: [9]

T
[]e = p,x p,y p,z ,
p

p,x = sin + coscos, p,y = cos cossin, p,z = + sin (51)

The components of the stiffness matrix given in Eq. (36): [9]

3kR2 

kxx = kyy = l2 b 1 + 2R 2R cos( 3 )



0




6k2H Rb2
kzz =




l02



3kR Rb4 2

n
k = k = R 2H + Rh + R 2h +


l2 0
(52)
2R H h 2h (H + h ) cos( 3 )








3k2R Rb4 

1 + cos( 3 + 2)

k =


l 2

0



k = k = 3kR Rb3 ( + 2 ) cos( )
h i
h
x y l 2 H h 3 R H R h R
0

where R , H and h are those given in Eq. (2) in which l0 is formulated as


h i 21
l0 = Rb 1 + 2R + 2H 2R cos( ) (53)
3

105
106

Vous aimerez peut-être aussi