ArticlePDF Available

Modeling and design optimization of a robot gripper mechanism

Authors:

Abstract and Figures

Structure modeling and optimizing are important topics for the design and control of robots. In this paper, we propose a process for modeling robots and optimizing their structure. This process is illustrated via a case study of a robot gripper mechanism that has a closed-loop and a single degree of freedom (DOF) structure. Our aim is to conduct a detailed study of the gripper in order to provide an in-depth step-by-step demonstration of the design process and to illustrate the interactions among its steps. First, geometric model is established to find the relationship between the operational coordinates giving the location of the end-effector and the joint coordinates. Then, equivalent Jacobian matrix is derived to find the kinematic model; and the dynamic model is obtained using Lagrange formulation. Based on these models, a structural multi-objective optimization (MOO) problem is formalised in the static configuration of the gripper. The objective is to determine the optimum force extracted by the robot gripper on the surface of a grasped rigid object under geometrical and functional constraints. The optimization problem of the gripper design is solved using a non-dominated sorting genetic algorithm version II (NSGA-II). The Pareto-optimal solutions are investigated to establish some meaningful relationships between the objective functions and variable values. Finally, design sensitivity analysis is carried out to compute the sensitivity of objective functions with respect to design variables.
Content may be subject to copyright.
1
Modeling and Design Optimization of a
Robot Gripper Mechanism
Alaa HASSAN1,, Mouhammad ABOMOHARAM2
1University of Lorraine, Equipe de Recherche sur les Processus Innovatifs (ERPI), Nancy, France
2Higher Institute for Applied Sciences and Technology (HIAST), Damascus, Syria
Abstract
Structure modeling and optimizing are important topics for the design and control of robots. In this paper, we
propose a process for modeling robots and optimizing their structure. This process is illustrated via a case study
of a robot gripper mechanism that has a closed-loop and a single degree of freedom (DOF) structure. Our aim is
to conduct a detailed study of the gripper in order to provide an in-depth step-by-step demonstration of the design
process and to illustrate the interactions among its steps. First, geometric model is established to find the
relationship between the operational coordinates giving the location of the end-effector and the joint
coordinates. Then, equivalent Jacobian matrix is derived to find the kinematic model; and the dynamic model is
obtained using Lagrange formulation. Based on these models, a structural multi-objective optimization (MOO)
problem is formalised in the static configuration of the gripper. The objective is to determine the optimum force
extracted by the robot gripper on the surface of a grasped rigid object under geometrical and functional constraints.
The optimization problem of the gripper design is solved using a non-dominated sorting genetic algorithm version
II (NSGA-II). The Pareto-optimal solutions are investigated to establish some meaningful relationships between
the objective functions and variable values. Finally, design sensitivity analysis is carried out to compute the
sensitivity of objective functions with respect to design variables.
Keywords
Robot modeling; Multicriteria design optimization; NSGA-II; Sensitivity analysis
1. Introduction
Robot design is a very complex process involving great modeling and simulation efforts. It has suffered an
important progress in the last decades and many approaches deal with this issue. Major steps in robot manipulator
design are; kinematics design, dynamics design, thermal design, and stiffness design [1]. In particular, robot
modeling and structural analysis are required in all industries. To address these requirements, a design process is
proposed in this paper that combines both; robot modeling and geometrical optimization. The proposed process is
a sub-process of the general robotics design process in which modeling and optimization activities play essential
and complementary roles in the design. As an illustrative case study, we carry out a modeling and an optimal
design of a planar single degree of freedom (DOF) mechanism that is used for robot hands or grippers. These kind
of mechanisms is amply used because of its simplicity and it only needs one actuator to move it, so many robots
use this kind of mechanisms as gripper. However, many researches deal with geometric, kinematic, and dynamic
modeling of the robots using different techniques. Some others work on optimization methods for multicriteria
robot design optimization. A survey of these research works is presented in the following paragraphs.
Corresponding author. Tel: +33372743500. E-mail address: alaa.hassan@univ-lorraine.fr. Postal address: 8 rue
Bastien Lepage, 54000 Nancy, France
2
Modeling is essential for design specifications, simulation, and advanced control of robots. Different techniques
of modeling are available for modeling robots, especially for parallel and closed-loop robots due to their
complexity [2-5]. Ibrahim and Khalil presented kinematic and dynamic modeling of three degrees of freedom 3-
RPS (revolute, prismatic, and spherical) parallel robot [6]. This robot is characterized by a coupling between the
6-DOF of the platform. After presenting a (6×3) kinematic Jacobian matrix, they developed a reduced (3×3)
Jacobian matrix relating the linear velocity of the platform with respect to the three actuated joints. In another
paper, Khalil and Guegan presented closed form solutions for the inverse and direct dynamic models of the Gough-
Stewart parallel robot. The models are obtained in terms of the Cartesian dynamic model elements of the legs and
of the Newton-Euler equation of the platform [7]. Andrzej et al. used forward and inverse kinematic problem as
well as working space and strength analysis issues for the construction of 3-DOF tripod electro-pneumatic parallel
manipulator [8]. Qin et al. proposed analytical modelling of a two-staged parallel mechanism composed by a rigid
platform in a serial connection with a compliant platform [9]. Hassan and Abomoharam performed a study of a
gripper that has two closed loop structure. After finding geometric and kinematic models, they determined the
geometrical solution space and verified it via a CAD model of the gripper [10]. Ha et al. employed Hamilton’s
principle, Lagrange multiplier, geometric constraints, and partitioning method to derive the dynamic equations of
a slider-crank mechanism. They showed that dynamic formulation could give a good interpretation of a slider-
crank mechanism by comparing the numerical simulations with experimental results [11]. Özgür and Mezouar
exploited screw theory expressed via unit dual quaternion representation and its algebra to formulate both the
forward (position and velocity) kinematics and pose control of an n-DOF robot arm [12].
Different researches of the optimum design of robot manipulators are available in the works of [1316]. Xie et al.
proposed a decoupled 3-DOF parallel tool head without parasitic motion. Using the atlases of the tool architecture
as bases, the optimal kinematic design of the tool head is carried out [17]. Jiang et al. presented a dynamic modeling
and redundant force optimization of a 2-DOF parallel kinematic machine with kinematic redundancy in order to
minimize the position errors of the manipulated platform [18]. Nevertheless, in real robot design problems, the
number of design parameters can be very large, and their influence on the value to be optimized (the objective
function) can be very complicated, having a strongly non-linear character. In these complex cases, stochastic
optimization techniques including evolutionary algorithms such as genetic algorithms (GA) may offer solutions to
the problem [19]. Coello et al. proposed GA-based multiobjective optimization hybrid technique to optimize the
counterweight balancing of a robot arm [20]. Jamwal et al. used a modified genetic algorithm to optimize the
kinematic design of a parallel ankle rehabilitation robot [21]. Osyczka and Krenich discussed some new methods
for multicriteria design optimization using evolutionary algorithms. The main aims of these methods is to reduce
the computing time and to facilitate the decision making process. Examples of a robot gripper mechanism and a
clutch break design are presented in this paper showing that these methods can be used to solve different design
optimization problems [22]. Gao et al. described the implementation of genetic algorithms and artificial neural
networks as an intelligent optimization tool for the dimensional synthesis of the spatial 6-DOF parallel
manipulator. The multi-objective optimization (MOO) problem was consisted of two functions: system stiffness
and dexterity, which are derived according to kinematic analysis of the parallel mechanism [23].
The rest of the paper is organized as follows. The proposed modeling and optimal design process of the robots and
its advantages are described in Section 2. In Section 3, our case study of a robot gripper mechanism is described
and its geometric modeling is recalled. Section 4 reviews the kinematic modeling of the gripper, then, the dynamic
model is derived in Section 5. After describing and modeling the gripper, the corresponding multi-objective
optimization problem is formalized in Section 6. Section 7 describes the solution algorithm of the optimization
problem, and the non-dominated sorting genetic algorithm version II (NSGA-II), it discusses the results. Section
8 presents the sensitivity analysis of the gripper mechanism design. Finally, Section 9 summarizes the contributions
and results made in this paper and gives some perspectives.
2. Modeling and Optimal Design Process
The design of robots is a complex engineering task, in which certain mathematical models are required. This task
can often be seen as an optimization problem in which the robot parameters or structure describing the best quality
design is sought. In this paper, an integrated modeling-optimizing robot design process is proposed where the
modeling steps are combined with the optimal structural design process, Fig. 1 illustrates this proposed process.
During this procedure, the geometric information is transferred from one step to the next step. The modeling stage
information is captured as input by the optimization stage, while the optimal design information feeds back the
modeling stage. These interactions give the designer the advantage to better define the design parameters and to
take into account both the modeling and the optimization issues in one integrated process. These two issues are
3
often handled separately as presented in the literature survey above, but in our presented process, they are
combined together in order to benefit of their complementarity.
The process starts by defining the problem that must be solved. Depending on the objectives of the study, the
applied steps may vary; it could be finding the geometric, kinetic, or dynamic models of the robot using different
techniques. These models are important to apply high performance control algorithms, to improve stiffness, to
increase payload, to improve force/torque capacity, etc. The objective could be also finding the optimal design that
aims at enhancing the performance indexes by adjusting the structural parameters, such as the geometrical lengths.
In the optimal design, several performance indices are involved, such as stiffness, transmission ratio, and accuracy.
The modeling stage starts by the geometric modeling, which represents the relations between the location vector
of the end-effector X and the joint coordinate vector q (Eq. (1)). Several methods and notations have been proposed
to find the geometric model; the most widely used one is that of Denavit-Hartenberg [24]. However, this method
is developed for simple serial-structured robot. Khalil and Kleinfinger have proposed a unified description of
parallel and tree-structured robots [25].
𝑋 = 𝑓(𝑞) (1)
Kinematic model is to find the relation between the end-effector velocity and the joint velocities. Kinematic model
could be written using the Jacobian matrix J. This matrix appears in calculating the derivation of the geometric
model. It gives the differential variations of the operational coordinates 𝑋̇ in terms of the differential variations of
the joint coordinates 𝑞̇ (Eq. (2)). For parallel manipulator, the key concept is to “break” the parallel manipulator
into simpleserial chains. Derived from the loop-closure or constraint equations, the equivalent Jacobian matrix
could be found in terms of active and passive joint variables.
𝑋̇= 𝐽(𝑞). 𝑞̇ (2)
The Jacobian matrix has multiple applications in robotics. It facilitates the calculation of singularities and of the
dimension of accessible operational space of the robot [26]. In static force model, we use the Jacobian matrix in
order to calculate the forces and torques of the actuators in terms of the forces and moments exerted on the
environment by the end-effector. The static model is essential in structural analysis as well as in formulating the
optimal design problem.
Fig. 1. Robot modeling and optimal design process.
Robot design problem
Geometric modeling
Kinematic modeling
Dynamic modeling
Modeling stage
Design parameterization
Optimization problem
formulation
(objective functions,
constraints)
Optimal design solutions
(optimization algorithm)
Design sensitivity
analysis
Optimized?
Stop
Yes
Design update
Optimization stage
No
4
The dynamic model is the relation between the torques (and/or forces) applied to the actuators and the joint
positions, velocities and accelerations. The relation in Eq. (3) represents the dynamic model.
= 𝑓(𝑞, 𝑞̇, 𝑞̈ , 𝜏𝑒𝑥𝑡) (3)
where
is the joint torques and forces. 𝑞, 𝑞̇, 𝑞̈ are the vectors of joint positions, velocities, and accelerations,
respectively. 𝜏𝑒𝑥𝑡 is the vector representing the external forces and moments that the robot exerts on its
environment. The dynamic model is typically used in actuator dimensioning and in robot simulation and control.
Several formalisms are used to obtain the dynamic model; the Lagrange multipliers, the principle of virtual work,
and the Newton-Euler formulation.
The optimization stage is a process consists of design parameterization, optimization problem formulation, optimal
design solutions, and design sensitivity analysis. The principal role of design parameterization is to define the
geometric parameters that characterize the structural model of the robot and to collect a subset of the geometric
parameters as design variables. Modeling stage supports design parameterization task by generating mathematical
models that describe geometric structure, kinetic, and dynamic behaviours of the robot. Hence, geometric
parameters and the structural design problem can be derived from the modeling stage. Only proper design
parameterization will yield a good optimum design, since the optimization algorithm will search within a design
space that is defined for the optimal design problem.
In optimization problem formulization step, stiffness, transmission ration, accuracy, cost, etc. can be defined by as
objective functions with appropriate constraint bounds. This involves the selection of objective functions,
expressed in terms of the design variables, which we seek to minimize or maximize. Beside the objective functions,
it involves the selection of a set of variables to describe the design alternatives. Constraint functions are the criteria
that the robot variables have to satisfy for each feasible design. In real world, it is common that a given structural
design problem has multiple and often conflicting objectives. This will result in a multi-objective optimization
problem from this step.
After defining the multi-objective design optimization problem, the next step is to find the optimal design
solutions. The presence of multiple objectives in a problem, in principle, gives rise to a set of optimal solutions
(largely known as Pareto-optimal solutions), instead of a single optimal solution. In the absence of any further
information, one of these Pareto-optimal solutions cannot be said to be better than the other. This demands a user
to find as many Pareto-optimal solutions as possible. Since evolutionary algorithms (EAs) work with a population
of solutions, a simple EA can be extended to maintain a diverse set of solutions. With an emphasis for moving
toward the true Pareto-optimal region, an EA can be used to find multiple Pareto-optimal solutions in one single
simulation run. The NSGA-II proposed by Deb et al. [27], is one of the EAs widely used to solve MOO problems.
Design sensitivity analysis is used to compute the sensitivity of objective functions with respect to design variables.
Based on the design sensitivity results, a design engineer can decide on the direction and amount of design change
needed to improve the objective functions. In addition, design sensitivity information can provide answers to what
if questions by predicting objective function perturbations when the perturbations of design variables are
provided. Regarding the geometric robot design, sensitivity analysis is of a great value to designers if a realistic
and economical allocation of tolerances on design variables is to be achieved. There are different approaches to
performing a sensitivity analysis like scatter plots, regression analysis, and partial derivative methods. Depending
on the design problem, the design engineer could test some of the optimal solutions to improve the robot design at
each iterative step. As a result, new designs could be obtained from optimization and sensitivity analysis steps.
Thus, the geometric model, in the modeling stage, has to be updated for the optimal set of design variables supplied
by the optimization stage.
In the rest of this paper, the proposed modeling and optimal design process is applied to a robot gripper mechanism.
The motivation of this case study is to model this closed-loop structure, and then to design it optimally.
3. Description of the gripper mechanism and geometric
modeling
The gripper is planner closed-loop mechanism with a single DOF. The gripping force F, applied on the object, is
generated by the actuating force P. Due to the symmetry; we can perform the study on a half of the mechanism
that is composed of three links and four joints (one prismatic and three revolute joints), as shown in Fig. 2..
5
The notations of Khalil and Kleinfinger [25], are used to describe the geometry of the closed-loop structure of the
gripper. The definition of the local link frames are given in Fig. 2, while the geometric parameters are given in
Table 1.
Fig. 2. Link frames of gripper mechanism.
Table 1
Geometric parameters of the gripper.
j
r
j
j
 
Pj
j
0
1
0
0
0
0
1
2
r
0
0
90
1
0
2
0
3
3
d
0
0
1
3
0
4
4
d
90
0
2
4
0
0
5
d
0
0
3
5
P(j) denotes the frame antecedent to frame j, and σj = 1 if joint j is prismatic and σj = 0 if it is revolute. The
homogeneous transformation matrix iTj, which defines the frame Rj relative to frame Ri, is obtained as a function
of four geometric parameters (αj, dj, θj, rj). Thus iTj, are obtained as:
11
11
01
cos sin 0 0
sin cos 0 0
0 0 1 0
0 0 0 1
T








,
3 3 3
33
13
cos sin 0
sin cos 0 0
0 0 1 0
0 0 0 1
d
T








,
5
35
1 0 0
0 1 0 0
0 0 1 0
0 0 0 1
d
T






(4)
3
1 0 0
0 0 1
0 0 1 0
0 0 0 1
tool
f
e
T






,
2
02
1 0 0 0
0 0 1
0 1 0 0
0 0 0 1
r
T






,
4 4 4
2444
cos sin 0
0 0 1 0
sin cos 0 0
0 0 0 1
d
T








The geometric closed-loop constraint can be expressed as
0 1 3 0 2
5
1 3 2 4
T T T T T
(5)
Thus
4 1 3
3 1 5 1 3 4
3 1 5 1 3 2
cos cos( )
sin sin( )
d d d
d d r
 
 
 

 
 
(6)
6
4. Kinematic modeling
Let
 
 
1,..., , 0
n
qq
  

be the constraint equations (Eq. (6)) previously found. Taking the time derivative
and rearranging these equations, Eq. 7 can be obtained as
   
*0K q K q


(7)
where
2
r
is the actuated variable (joint) and
 
1 3 4
,, T
  
are the passive variables (joints). n is the total
number of the gripper joints, and m be the number of the actuated joints. Columns of the
 
 
n m m
matrix [
 
Kq
] are the partial derivatives of
 
q
with respect to the actuated variables
i
, i = 1,…, m. Columns of the
 
 
n m n m  
matrix [
 
*
Kq
] are the partial derivatives of
 
q
with respect to the passive variables
i
, i
= 1,…, n-m.
If
 
 
*
det 0Kq
, we can determine the linear and angular velocities of the passive joints (
) in terms of only
( the linear and angular velocities of the actuated joints)
 
1
*
KK


(8)
Hence, one can write the linear and angular velocities of the end-effector, denoted by tool in Fig. 2, with respect
to the fixed coordinate frame R0 as
0*
0*
tool v v
tool
v J J
JJ


 


(9)
which leads to
 
 
 
 
,
,
1
0 * *
1
0 * *
v eq
eq
tool v v
J
tool
J
v J J K K
J J K K




(10)
Define the equivalent Jacobian matrix Jeq as
,
,
v eq
eq eq
J
JJ



(11)
Eq. (9) can be written as
0
0tool
tool eq
tool
v
VJ



(12)
Thus, kinematic model of the gripper can be found by the calculation of Jeq. The derivative of constraint equations
Eq. (6) with respect to time gives
 
   
 
   
3 1 5 1 3 5 1 3
*3 1 5 1 3 5 1 3
0 sin sin sin 0
1 , cos cos cos 0
0 1 1 1
d d d
K K d d d
   
   
 

 

 

 

 
(13)
The transformation matrix
0tool
T
=
0 1 3
13
..
tool
T T T
is combined of rotation matrix
0tool
R
and translation vector
0tool
P
with respect to R0
   
 
   
     
 
 
1 3 1 3 1 1 3 1 3
1 3 1 3 1 3 1
0
00
13
cos sin 0 3cos cos sin
sin cos 0 cos 3sin sin
0 0 1 0
0 0 0 1
0 0 0 1
tool
tool tool
d f e
ed
T
RP
f
       
       






 
 



(14)
The linear velocity of the tool
0tool
v
is calculated by the derivation of position vector
0tool
P
with respect to time
7
 
       
 
       
1
3 1 1 3 1 3 1 3 1 3
3 1 1 3 1 3 1 3
*
33
4
02 1
d sin cos sin cos sin 0
d cos cos sin cos s
0
0 in 0
0 0 00
tool
vv
e f e f
ev er
JJ
ff
       
       

   



    









 

(15)
While the angular velocity vector of the tool
0tool
is obtained from the angular velocity matrix
.
0 0 0
tool tool
0
. 0 ,
0
x
zy
T
z x tool y
yx z
RR

 







  





(16)
hence
4
*
1
032
0 0 0 0
0 0 0 0
0 1 1 0
tool
J
r
J




 


(17)
Substituting Eqs. (13), (15), and (17) into Eq. (10) gives
,v eq
J
and
,eq
J
, then, the equivalent Jacobian matrix Jeq
is obtained
 
   
 
 
 
 
 
   
 
 
 
 
 
1 1 3 5 1 3
53
1 1 3 5 1 1 1 3
53
1
53
sin cos (d )sin
d sin
sin cos d cos sin sin
d sin
0
0
0
sin
d sin
eq
ef
fe
J
   
   









  
 





(18)
Based on the obtained kinematic model, we can establish the relationship between the external applied
forces/torques and the joint forces/torques. To find this relationship, in our case study, the principle of virtual work
is applied
00
. X .
TT
tool tool
F
 
(19)
where
0tool
F
is the vector of the external forces/torques applied on the tool, assuming that there is no other external
forces/torques.
0Xtool
is the virtual displacement vector of the tool.
is the vector of forces/torques applied on
the actuated joints.

is the virtual displacement vector of the actuated joints.
Substituting Eq. (12) into Eq. (19) gives
0.
TT
tool eq
FJ
  
(20)
Therefore
0
.
T
eq tool
JF
(21)
Substituting
0
0
0
0
0
0
0
tool
tool
RF
F















(22)
/2P

, and
eq
J
(Eq. (18)) into Eq. (21), the relationship between the actuator force and the force applied on
the object can be found
8
 
 
 
53
5 1 5 1 3
d sin
(d 2 )sin d sin 2
P
Ff
 
 
(23)
Eq. (23) is resulted from kinematic modeling, and it will be transmitted to design parametrization. This relationship
is the basis of the optimization problem formulization, as shown in the next paragraphs.
5. Dynamic modeling
To obtain dynamic model of the gripper, Lagrangian formulation is used. As above, n is the total number of joints,
and m is the number of the actuated joints. The Lagrangian formulation for a closed-loop mechanism is
1, 1,....,
nm j
ij
j
i i i
d L L Q i n
dt q q q
(24)
where the scalar Lagrangian is defined from the total kinetic and potential energy
1
( , ) ( )
n
ii
i
L q q KE PE

(25)
i
KE
kinetic energy of the link i
i
PE
potential energy of the link i
i
Q
the externally generalised forces applied on link i, it includes the actuator force/torque (if i is actuated) and the
external forces/torques
 
1 3 2
T
qr

joint coordinate vector
j
()nm
loop-closure constraint equations
j
()nm
Lagrange multipliers
Eq. (24) can be written in matrix form
   
,T
M q q C q q q G q Q
 
(26)
()Mq
nn
mass matrix
 
,C q q
nn
Coriolis/centripetal matrix, where
1
1
2
nij kj
ik
ij k
kk j i
MM
M
Cq
q q q


 


 

 
Gq
1n
vector of gravity terms, where
()
ii
PE
Gq
()n m n
constraint matrix that is obtained from the partial derivatives of
()nm
constraint equations with
respect to
j
q
,
i
ij j
q

For the gripper mechanism, let
( , , , ), 1,..., ,
i i i i
m l r I i n
denote mass, length, center of gravity location, and
component of inertia matrix of link i, respectively. For planar case, only
ii
I Iz
is relevant, the gravity is along (-
x0) axis (Fig. 2). Thus, Eq. (26) matrices are written as
 
 
 
 
 
 
 
2 2 2
1 1 1 3 3 1 3 3 1 3 3 3 3 1 3 3
2
3 3 3 1 3 3 3 3 3
2
2cos cos 0
cos 0
00
m r Iz Iz m l r l r Iz r l m r
M q Iz r l m r m r Iz
m


   

 




(27)
9
 
   
 
 
132 2 2 1 2
21
3 1 3 3
1 3 3
sin sin 0
sin 0 0
00
,
0
l m r l
C q q
mr
l m r
 







(28)
 
 
 
 
 
1 3 3 3 1 1 3 1 1
1 3 3 3
sin sin
sin
0
g m r g l m m r
g m rGq
 





(29)
 
 
 
 
 
   
1 3 5 1 3 1 3 5
1 3 1 3 5 1 3 5
sin sin sin 0
cos cos cos 1
dd
dd
qd
d
   
   
 
   




(30)
Q
1n
vector of joint torque/force (
) and externally applied torques/forces (
ext
), this vector is written as
ext
Q


(31)
The actuating force vector is
 
0 / 2 0 T
P

with respect to the fixed coordinate frame R0.
The externally applied force is due to the gripping force F, and it is already expressed in Eq. (22). The externally
applied force term in Lagrangian equation is
T
ext F

(32)
Where
is the matrix of the relationship between the linear velocity of F application point
0tool
V
and the joint
velocity
1 3 2
T
qr



, i.e.
1
03
2
.
tool
V
r





(33)
To find
, as in Eq. (15) and Eq. (17), we can obtain that
 
       
 
     
3 1 1 3 1 3 1 3 1 3 1
03
2
3 1 1 3 1 3 1 3 1 3
d sin cos sin cos sin 0
d cos cos sin cos sin 0
0 0 0
v
tool
e f e
r
feV
f
fe
       
       




   





   
(34)
and
1
03
2
000
000
1 1 0
tool r











(35)
which leads to
10
 
 
     
 
       
1 3 3 1 1 3 1 3 1 3
3 1 1 3 1 3 1 3 1 3
cos d sin sin cos sin 0
d cos cos sin cos sin 0
0 0 0
0 0 0
0 0 0
1 1 0
v
e f e f
f e f e
       
       
   





  







   
(36)
Substituting the right side of Lagrangian equation Eq. (26) gives
 
   
 
   
 
   
 
3 3 1 1 1 2 5 1 3 1 1 3 2
5 1 3 1 1 3 2
2
cos sin cos sin cos
sin cos
2
T
fF d F d
fd
P
Q F
   
   


 

 



(37)
After finding the dynamic model of the gripper, the relationship between the actuator force and the force applied
on the object can be found. By setting the joint velocities, accelerations, and gravity terms to zero, we can find
 
   
 
   
 
   
 
3 3 1 1 1 2 5 1 3 1 1 3 2
5 1 3 1 1 3 2
2
cos sin cos sin cos
0
0 sin cos
0
2
fF d F d
fF d
P
   
   


 
 

 

 

 


(38)
Solving this system of equations gives the relationship of the gripping force
 
 
 
 
53
5 1 5 1 3
sin
sin n2 si 2
dP
Fd f d
 
 
(39)
Eq. (39), which is resulted from the dynamic model, is the same one we found in Eq. (23) that is resulted from
kinematic model. Depending on the study objectives, finding kinematic and/or dynamic model could be carried
out, and both lead to the same static equation. This equation is transmitted to optimization stage, particularly to
formulize the gripper optimization problem.
6. Optimization problem formulation
The goal of the optimization problem is to find the dimensions of the gripper elements and to optimize objective
functions simultaneously by satisfying the geometric and force constraints. The vector of six design variables is
 
3 4 5
, , , , ,x d d d l e f
, where
3 4 5
, , , , ,d d d l e f
are the gripper link variables used in geometrical modeling. The
structure of geometrical dependencies of the mechanism is described in Fig. 3. The angles
1
2


and
3
 

can be written in terms of the design variables as
C
y/2
Fig. 3. Force distribution and geometrical variables of the gripper mechanism.
11
2 2 2
53
5
2 2 2
35
3
arccos 2
arccos 2
g d d
gd
g d d
gd

 







  

(40)
where
 
2
2
4
4
arctan
g d l z
d
lz
 



6.1 Objective functions
Based on the relationship between the gripping force and the actuator force in Eq. (23) or Eq. (39), we can
rewrite this relationship in terms of the design variables,
and
 
5
5
sin( )
,cos( )cos( ) cos( ) 2
dP
F x z df

 

(41)
The objective functions of the gripper are borrowed from [22] to perform a bi-objective study to understand the
trade-off between chosen objectives. These two objective functions can be formulated as follows:
1. The first objective function (Eq. (42)) can be written as the difference between the maximum and
minimum gripping forces for the assumed range of gripper ends displacement. The minimization of this
objective ensures that there is not much variation in the gripping force during the entire range of operation
of the gripper. Thus, it ensures the minimization of stress variation in the gripper links.
 
1max ( , ) min ( , )
z
z
f x F x z F x z
(42)
2. The second objective function (Eq. (43)) is the force transmission ratio, the ratio between the applied
actuating force
P
and the resulting minimum gripping force at the tip of the gripper end. The
minimization of this objective will ensure that the gripping force experienced at the tip of link f has the
largest possible value. Thus, it ensure the maximization of force transmission ratio.
 
2min ( , )
z
P
fx F x z
(43)
In the previously mentioned multi-objective optimization problem, both objective functions depend on the vector
of decision variables and on the displacement
z
. The parameter
z
is the displacement parameter of the gripper
actuator, which takes its minimum value at 0 and its maximum value at
max
Z
. Checking for a number of different
solution vectors (
x
), it is observed that
F
 
1max ( , ) min ( , )
z
z
f x F x z F x z
is a decreasing function as shown in
Fig. 4. So that, the maximum value of
F
takes place at
0z
and the minimum value takes place at
max
zZ
.
Thus, the objective functions may be written as:
 
 
1 max
2max
( ,0) ( ,Z )
( , )
f x F x F x
P
fx F x Z

(44)
6.2 Constraints
Before presenting the gripper constraints, we need to conclude the relationship between the displacement of the
gripper end
y
and the displacement of the gripper actuator
z
(Fig. 3).
 
 
45
, 2 sin siny x z d d c
 
 
(45)
where
22
c f e
arctan e
f



12
Fig. 4. Variation of force
F
with the displacement
z
for a typical design vector
x
.
From the geometry of the gripper, a number of non-linear constraints can be derived:
1. The minimum displacement between the ends of gripper (corresponding to the maximum displacement
of actuator value at
max
Z
) should be less than the minimum dimension of the gripping object:
 
 
1 min max
: , 0g x Y y x Z
(46)
2. The distance between the gripper ends corresponding to
max
Z
should be greater than zero:
 
 
2 max
: , 0g x y x Z
(47)
3. The maximum distance between the gripping ends corresponding to no displacement of actuator (
0z
)
should be greater than the maximum dimension of gripping object:
 
3 max
: ,0 0g x y x Y
(48)
4. Maximal range of the gripper ends displacement (
G
Y
) should be greater than or equal to the distance
between the gripping ends corresponding to no displacement of actuator:
 
4: ,0 0
G
g x Y y x
(49)
5. Maximal displacement of actuator should be greater than
l
, the actuator stroke should not reach the point
O
(Fig. 3):
 
5 max
:0g x l Z
(50)
6. The gripper ends displacement (
y
) decreases when the actuator displacement (
z
) increases. To ensure
this condition and to not reverse the gripper ends displacement direction, the angle
should be less than
2
:
 
6:0
2
g x z

(51)
7. The angle
is the gripper declination refers to the horizontal axis, a stable gripping requires a limitation
of this angle:
 
7:5gx
(52)
8. Geometrical properties are preserved by two constraints on
to maintain the triangle
OAB
. From Eq.
(40), the arccos function input should be less than 1:
2
2 2 2 2 2 2
5 3 5 3
55
11
22
g d d g d d
gd gd

   
 


After simplification and substitution of
g
, we get:
 
 
 
 
 
 
2 2 2 2
22
3 5 4 3 5 4
12
0d d d l z d d d l z
aa
   
So that,
1
a
and
should be positives for all
 
max
0,zZ
.
5
10
15
20
25
z
20
40
60
80
100
F
z (mm)
F (N)
(mm)
13
 
 
22
2
8 4 max 3 4
:0g x d l Z d d 
(53)
 
 
222
9 3 4 4
:0g x d d d l  
(54)
9. The geometric bounds of link lengths, or design variables, (in mm), are:
3
4
5
10 50
10 50
10 60
10 50
5 15
50 100
d
d
d
l
e
f






10. The geometric and force parameters are assumed to be as:
min max max
30mm, 70mm, 100mm,Z 25mm
G
Y Y Y  
and
95NP
Thus, the gripper optimization problem can be formulated as follows:
Find
 
* * * * * * *
3 4 5
, , , , ,x d d d l e f
which will satisfy the 9 inequality constraints
 
1,...,9
k
g x k
(55)
and minimize the two objective functions
 
*12
min[ ( ), ( )]f x f x f x
(56)
7. Solution algorithm and result discussion
After formulating the optimization problem, the next step is to find an optimal design solution using an appropriate
algorithm. Our gripper case study is a MOO problem; and all of the above constraints must be taken into account.
In order to solve it and find an optimal solution, there are several methods which are proposed in the literature.
Due to the complexity, the size of problem and the importance of reducing the solving time, NSGA-II algorithm
is used. This algorithm results in Pareto front that consists of a set of solutions, which are not dominated by each
other. The Pareto-optimal solutions are thoroughly investigated to establish some meaningful relationships
between the objective functions and variable values.
The NSGA-II procedure is illustrated in Fig. 5. In this procedure, an initial combined population (Rt) is randomly
created and this population must be sort based on the non-domination strategy into each front. The fronts (F1, F2,
F3 ...) are compared with each other and a rank is assigned to each individual according to the fitness value. In
addition, a new parameter, named crowding distance is calculated in order to sort the solutions of F3. It is the
distance of an individual to its neighbors, and the large value of this parameter shows the better diversity in
population. The binary tournament selection is used to select the parents due to the rank and crowding distance.
The mutation and crossover operators are utilized for the creation of a new offspring (Qt+1) and so on for the next
generations.
NSGA-II parameters are as follows: population size = 200, number of generations = 2000, probability of SBX
(simulated binary crossover) recombination = 0.9, probability of polynomial mutation = 0.1, distribution index for
real-variable SBX crossover = 20, and distribution index for real-variable polynomial mutation = 100.
Pt
Qt
Rt
F1
Crowding
distance
sorting
F2
F3
Rejected
Pt+1
Nondominated
sorting Qt+1
Selection
+
Crossover
+
Mutation
14
Fig. 5. NSGA-II procedure [28].
Fig 6 shows the obtained Pareto front resulting from NSGA-II optimization procedure. From this set of solutions,
three are selected to investigate the changes in the gripper configuration. We have taken two extreme solutions (A
and C) and one intermediate solution (B) as shown in Fig. 7. These three solutions are presented in Table 2, the
design variable values are cited with their corresponding objective function values.
To investigate how one optimal solution varies from another, we can analyze the values of six design variables as
a function of one of the objectives (the force transmission ratio for example). It is clear that variables e and f are
fixed at their allowable lower limit value. We notice also that the force transmission ratio (f2) is proportional to
3 4 5
, , ,d d d l
design variables. On the other hand, the difference between the maximum and minimum gripping
forces (f1) is inversely proportional to these design variables. Therefore, the designer task is to select one of the
optimal solutions compromising the two objective functions in terms of the design requirements.
8. Design sensitivity analysis
After selecting an optimal design solution, the last step is to analyse the design sensitivity to design variable
changes. Design sensitivity analysis computes the rate of objective function change with respect to design variable
changes. With the robot structural analysis, the design sensitivity analysis generates a critical information, gradient,
for design optimization. Obviously, the objective function is presumed to be a differentiable function of the design,
at least in the neighbourhood of the optimal solution point, which is found, and selected, in the previous step.
Fig. 6. Set of optimal solutions obtained using NSGA-II.
Fig. 7. Three gripper configurations for solutions A, B, and C.
Table 2
Objective function and design variable values for solutions A, B, and C.
Solution
f1 (N)
f2
d3 (mm)
d4 (mm)
d5 (mm)
l (mm)
e (mm)
f (mm)
A
93.81
6.64
32.92
19.08
54.73
49.85
5.00
50.00
B
138.90
6.25
30.89
18.88
52.31
49.74
5.00
50.00
C
184.2
6.09
29.69
18.80
51.28
49.71
5.00
50.00
f2
100
120
140
160
180
6.1
6.2
6.3
6.4
6.5
6.6
f1 (N)
A
B
C
15
A sensitivity analysis of the mechanism is of a great value to designers if a realistic and economical allocation of
tolerances on link-lengths is to be achieved. Such analysis enables the designer to notice important trends, to
identify most critical link of a given mechanism, and to allocate the tolerances optimally [29]. In the gripper case
study, the design variables are all link-lengths, so the attention is drawn to the effect of practical manufacturing
tolerances on the objective functions. The local gripper design sensitivity to link-length tolerances is studied by
using deterministic approach, based on the worst-case analysis of the individual tolerances (maximum output
tolerance). Selecting an optimal solution (x*) on point B, illustrated in Table 3, the gripper objective functions
sensitivity can be conducted as follows.
Table 3
Objective functions and selected optimal solution for point B.
Solution
f1* (N)
f2*
d3* (mm)
d4* (mm)
d5* (mm)
l* (mm)
e* (mm)
f* (mm)
B
138.90
6.25
30.89
18.88
52.31
49.74
5.00
50.00
Considering tolerances on link-lengths, the actual lengths of the links deviate from the nominal:
iact i nom i
x x x  
. If the tolerances are much smaller than the link-lengths ti << xi, the variation in the objective function f1 may be
written in terms of Taylor series as:
   
**
1 1 1
1
. . .
2
TT
f x f x x H x x  
(57)
where
 
3 4 5
, , , , ,x d d d l e f  
1
1i
f
fx

is the first partial derivative of the objective function f1 with respect to the ith design variable (xi)
1
H
is the second partial derivative matrix called the Hessian matrix,
21
1,ij ij
f
Hxx

In our case, the terms in Taylor series expansion having order three and above could be neglected without
appreciable loss of accuracy. We also assume that the constraints are still satisfied when
i
x
occurs. The magnitude
of the sensitivity coefficients for a design variable indicates the relative importance and influence of that variable
on the variation of the objective function. Fig. 8 shows the values of the first order sensitivity coefficients of the
objective function f1 on the optimal solution point (x*). It shows that the objective function f1 is most sensitive to
the variation of design variables d5 and l, and less sensitive to the other variables.
The same formula in Eq. (57) is applied for the second objective function ( f2). Fig. 9 shows the values of the first
order sensitivity coefficients of the objective function f2 on the optimal solution point (x*). Like f1, the objective
function f2 is most sensitive to the variation of design variables d5 and l. It is less (approximately equal) sensitive
to the variables d3 and d4, and much less sensitive to e and f.
The variation of each objective function can now be evaluated when a rough manufacturing tolerances (variations)
of the design variables are assigned, as illustrated in Table 4. The nominal values of the design variables are equal
to the optimal solution values on point B, while the tolerances are derived from the International Tolerance Grades
table in ISO 286. Based on the worst-case analysis, the total variation of an objective function is therefore the sum
of the individual variations due to each of the design variables considered separately. From Eq. 57, the calculated
maximum variation for each objective function is
122.4Nf
, and
20.414f
Since d5 and l have the major influences on the gripper objective functions, the tolerance intervals of these two
variables can be restricted in order to get the objective function variations within acceptable limits. On the other
hand, it is desirable to give as much of tolerance as possible to keep the manufacturing costs low. The designer
has to select the optimal solution and the assigned tolerance intervals that compromise multiple conflicting design
16
objectives. Once the optimal solution is selected, the geometric model must be updated and the optimization stage
is completed.
Fig. 8. First order sensitivity coefficients of f1 to design variable variations on point B.
Fig. 9. First order sensitivity coefficients of f2 to design variable variations on point B.
Table 4
Manufacturing tolerances (in mm) assigned to the gripper design variables.
3
d
4
d
5
d
l
e
f
0.3
0.2
0.3
0.3
0.1
0.3
9. Conclusion
This paper has proposed a two-stage process for modeling and optimising robot structures. In the modeling stage,
an analytical step-by-step study is introduced to find geometric, kinematic and dynamic models. The optimization
stage shows the procedure to formulize and to solve a design optimisation problem and then to analyse the design
sensitivity. The data flow and interactions between these steps and stages are highlighted when a robot modeling
and optimal design are needed. To illustrate the proposed process, a case study of a robot gripper is carried out.
The process starts by the geometric modeling step to find the geometric closed-loop constraint equations of the
gripper. Based on these equations, the kinematic model is derived in terms of equivalent Jacobian matrix and the
velocity of the actuated joint, then, Lagrangian formulation is employed to find the dynamic model.
The modeling stage data is used in the optimization stage to formalize the optimization problem. The relationship
between the gripper force and the actuator is resulting from the kinematic and dynamic models whereas some of
geometrical constraints are derived from the geometrical model. Two objective functions are expressed: the
minimization of the difference between the maximum and minimum gripping force and the maximization of the
17
force transmission ratio. NSGA-II is applied to solve the problem resulting the Pareto-optimal solutions. Each one
of these solutions represents a configuration of the gripper with a set the link variable values. A local sensitivity
analysis of a trade-off solution has showed that the two objective functions are most sensitive to the variation of
two design variables d5 and l, and less sensitive to the other variables. The manufacturing tolerance intervals of
these two variables can be restricted in order to maintain the objective function variations within acceptable limits.
Further improvements in this design process will focus on other MOO algorithms that may be used to find the
optimal solutions. It is also vital to integrate multi-objective robust optimization principle in order to find multi-
objectively robust Pareto optimum solutions.
References
[1] Ölvander J, Tarkian M, Feng X. Multi-objective optimization of a family of industrial robots. In: Wang L,
Ng AHC, Deb K, editors. Multi-objective Evolutionary Optimisation for Product Design and Manufacturing.
Netherlands: Springer; 2011, p. 189-217.
[2] Bhattacharya S, Nenchev D.N, Uchiyama M. A recursive formula for the inverse of the inertia matrix of a
parallel manipulator. Mechanism and Machine Theory 1998;33(7):957-964.
[3] Miller K. Optimal design and modeling of spatial parallel manipulators. The International Journal of Robotics
Research 2004:23(2):127-140.
[4] Tsai L-W. Solving the inverse dynamics of a Stewart-Gough manipulator by the principle of virtual work.
Journal of Mechanical Design 2000:122(1):3-9.
[5] Wang M, Luo M, Li T, Ceccarelli M. A Unified dynamic control method for a redundant dual arm robot.
Journal of Bionic Engineering 2015:12:361-37.
[6] Ibrahim O, Khalil W. Kinematic and dynamic modeling of the 3-RPS parallel manipulator. Proceedings of
12th IFToMM World Congress, Besançon, France, 2007.
[7] Khalil W, Guegan S. Inverse and direct dynamic modeling of Gough-Stewart robots. IEEE Transaction on
Robotics and Automation (2004);20(4):754-762.
[8] Laski PA, Takosoglu JE, Blasiak S. Design of a 3-DOF tripod electro-pneumatic parallel manipulator.
Robotics and Autonomous Systems 2015:72:59-70.
[9] Qin Y, Zhang K, Li J, Dai JS. Modelling and analysis of a rigidcompliant parallel mechanism. Robotics and
Computer-Integrated Manufacturing 2013;29:33-40.
[10] Hassan A, Abomoharam. Design of a single DOF gripper based on four-bar and slider-crank mechanism for
educational purposes. Procedia CIRP (2014);21:379384.
[11] Ha J-L, Fung R-F, Chen K-Y, Hsien S-C. Dynamic modeling and identification of a slider-crank mechanism.
Journal of Sound and Vibration 2006;289:1019-1044.
[12] Özgür E, Mezouar Y. Kinematic modeling and control of a robot arm using unit dual quaternions. Robotics
and Autonomous Systems 2016:77:66-73.
[13] Zhang D, Wang L, Esmailzadeh E. PKM capabilities and applications exploration in a collaborative virtual
environment. Robotics and Computer-Integrated Manufacturing 2006;22(4):38495.
[14] Bergamaschi PR, Nogueira AC, Saramago FP. Design and optimization of 3R manipulators using the
workspace features. Applied Mathematics and Computation 2006;172(1):43963.
[15] Rout BK, Mittal RK. Parametric design optimization of 2-DOF RR planar manipulator: a design of
experiment approach. Robotics and Computer-Integrated Manufacturing 2008;24(2):23948.
[16] Yu A, Bonev IA, Paul ZM. Geometric approach to the accuracy analysis of a class of 3-DOF planar parallel
robots. Mechanism and Machine Theory 2008;43(3): 364-375.
[17] Xie F, Liu X-J, Wang J. A 3-DOF parallel manufacturing module and its kinematic optimization. Robotics
and Computer-Integrated Manufacturing 2012;28:334343.
[18] Jiang Y, Li T-M, Wang L-P. Dynamic modeling and redundant force optimization of a 2-DOF parallel
kinematic machine with kinematic redundancy. Robotics and Computer-Integrated Manufacturing
2015;32:110.
[19] Renner G, Ekart A. Genetic algorithms in computer aided design. Computer-Aided Design 2003;35:709
726.
[20] Coello C, Christiansen A, Aguirre A. Using a new GA-based multiobjective optimization technique for the
design of robot arms. Robotica 1998;16:401414.
18
[21] Jamwal P, Xie S, Aw K. Kinematic design optimization of a parallel ankle rehabilitation robot using modified
genetic algorithm. Robotics and Autonomous Systems 2009;57:10181027.
[22] Osyczka A, Krenich S. Some methods for multicriteria design optimization using evolutionary algorithms.
Journal of Theoretical and Applied Mechanics 2004;42(3):565584.
[23] Gao Z, Zhang D, Ge Y. Design optimization of a spatial six degree-of-freedom parallel manipulator based
on artificial intelligence approaches. Robotics and Computer-Integrated Manufacturing 2010;26:180189.
[24] Denavait J, Hartenberg R.S. A kinematic notation for lower pair mechanism based on matrices. Journal of
Applied Mechanics 1955;22: 215221.
[25] Khalil W, Kleinfinger J-F. A new geometric notation for open and closed-loop robots. IEEE conference on
robotics automation (1986):11741180.
[26] Paul P. Robot manipulators: mathematics, programming and control. MIT Press, Cambridge, 1981.
[27] Deb K, Agrawal S, Pratap A, Meyarivan T. A fast and elitist multi-objective genetic algorithm: NSGA-II.
IEEE Transactions on Evolutionary Computation (2002);6(2):182197.
[28] Deb K, Agarwal R. B. Simulated binary crossover for continuous search space. Complex systems
1995:9:115148.
[29] Pavlović N. Analysis of mechanical error in quick-return shaper mechanism. Proceedings of 12th IFToMM
World Congress, Besançon, France, 2007.
... Designing kinematic mechanisms is an important topic covered in a wide range of areas across the industry. Depending on the application, the mechanism is classified into a passive and actuating system [1][2][3][4][5][6][7][8][9][10][11][12]. For example, the passive system can be classified as automobile suspension, compensation mechanism, and the actuating system can be classified as a robot gripper, manipulator mechanism. ...
... The linkage mechanism is commonly used because it is inefficient to include the task region with only circular motion. The linkage mechanism has the potential to perform various tasks because it has an attractive feature that can transmit circular motion to different path motions [4,6,10,11,[13][14][15][16]. ...
... Moreover, since only the target path was considered as an object function, there is an obvious limitation in that it does not take into account the torque transmission required to actuate the mechanism. Hassan and Abomoharam also proposed a study that optimizes the structure of the gripper for holding force that has a linkage structure [4]. Dynamic models for maximizing grip were numerically calculated and optimized. ...
Preprint
Full-text available
This paper proposes a framework that optimizes the linkage mechanism of the quasi-serial manipulator for target tasks. This process is explained through a case study of 2-degree-of-freedom linkage mechanisms, which significantly affect the workspace of the quasi-serial manipulator. First, a vast quasi-serial mechanism is generated with a workspace satisfying a target task and it converts it into a 3D CAD model. Then, the workspace and required torque performance of each mechanism are evaluated through kinematic and dynamic analysis. A deep learning-based surrogate model is leveraged to efficiently predict mechanisms and performance during the optimization process. After model training, a multi-objective optimization problem is formulated under the mechanical and dynamic conditions of the manipulator. The design goal of the manipulator is to recommend quasi-serial mechanisms with optimized kinematic (workspace) and dynamic (joint torque) performance that satisfies the target task. To investigate the underlying physics from the obtained Pareto solutions, various data mining techniques are performed to extract design rules that can provide practical design guidance. Finally, the manipulator was designed in detail for realization with 3D printed parts, including topology optimization. Also, the task-based optimized manipulator is verified through a payload test. Based on these results, the proposed framework has the potential for other real applications as realized cases and provides a reasonable design plan through the design rule extraction.
... There are some research works that employ biomimetic grasping, which entails dragging a finger across a surface to create a protrusion in the fabric's body that the effector may grip [5][6][7][8]. There has been some numerical work, namely in simulation and optimization of robotic grippers, aiming to predict and/or improve their behavior in real engineering applications [9][10][11][12]. Robotic gripper design optimization is crucial for stable grasping. ...
... Robot design and control require structure modeling and optimization. In the work [10], the authors modelled and optimized a robot structure by means of a closed-loop, single-DOF robot gripper device. The authors' goal is to explore the gripper in detail, to explain the design process and its relationships. ...
... The authors examine Paretooptimal strategies to construct meaningful links between the objective functions and variable values. Design sensitivity analysis computes objective function sensitivity to design variables [10]. In [11], the research uses intelligent strategies to optimize a robot gripper's geometry. ...
Article
Full-text available
Simulation and optimization have become common tasks in engineering practice due to their advantages, namely cost reduction and unlimited testing prior to manufacturing. Over the last years, personal computers have become powerful enough to run complex simulations. On the other hand, industry has seen an increase in automation, where repetitive tasks done by humans, in the past, are gradually being replaced by robotic systems. Those robotic systems usually involve a robotic arm, a gripper and a control system. This article presents a methodology for the simulation and optimization of existing engineering parts i.e., based on reverse engineering. The models were subjected to static loadings and free vibration (modal) analysis, in the Finite Element Method (FEM) software ANSYS Workbench 2021 R2. The adaptive Multi-objective optimization (AMO) algorithm was applied, also in ANSYS Workbench 2021 R2. The effectiveness of the proposed methodology was evaluated, and the outcome was that significant improvement can be achieved in terms of both static and dynamic behavior of the analyzed part.
... and geometrical variables of the gripper mechanism[44]. ...
... Force distribution and geometrical variables of the gripper mechanism[44]. ...
... Force distribution and geometrical variables of the gripper mechanism [44]. Table A6. ...
Article
Full-text available
The teaching–learning-based optimization (TLBO) algorithm, which has gained popularity among scholars for addressing practical issues, suffers from several drawbacks including slow convergence speed, susceptibility to local optima, and suboptimal performance. To overcome these limitations, this paper presents a novel algorithm called the teaching–learning optimization algorithm, based on the cadre–mass relationship with the tutor mechanism (TLOCTO). Building upon the original teaching foundation, this algorithm incorporates the characteristics of class cadre settings and extracurricular learning institutions. It proposes a new learner strategy, cadre–mass relationship strategy, and tutor mechanism. The experimental results on 23 test functions and CEC-2020 benchmark functions demonstrate that the enhanced algorithm exhibits strong competitiveness in terms of convergence speed, solution accuracy, and robustness. Additionally, the superiority of the proposed algorithm over other popular optimizers is confirmed through the Wilcoxon signed rank-sum test. Furthermore, the algorithm’s practical applicability is demonstrated by successfully applying it to three complex engineering design problems.
... 15 The current stateof-the-art gripper state prediction has primarily focused on rigid systems, where standard robotic kinematics and dynamics can be applied. 16 However, soft grippers, such as bionic finger grippers, challenge these traditional models due to their inherent compliance and infinite degrees of freedom. 17,18 This challenge has necessitated the development of innovative approaches to predict and control their behavior. ...
Article
Full-text available
Soft grippers in automation, particularly those with variable joint stiffness, offer promising possibilities for precise manipulation tasks. However, accurately predicting finger joint bending angles in this field poses significant challenges due to the soft and complex nature of the grippers, making modeling and angle prediction difficult. This paper presents the development of a predictive model for precisely controlling bending angles in multi-material printed soft grippers with variable stiffness, which are pivotal for delicate manipulation tasks in automation. In particular, we explore a cable-driven gripper design made of thermoplastic polyurethane and conductive polylactic acid materials, featuring integrated resistive joints for stiffness modulation through controlled Joule heating. A data-driven modeling approach, combining numerical modeling of the gripper and machine learning techniques, was employed for the development of the predictive model. We performed static structural simulations using ANSYS Workbench to measure bending angles under various conditions for developing datasets for model training. In this work, we evaluated several machine learning models such as linear regression, decision tree, and K-nearest neighbor regression models to predict the correlation between temperature, pull distance, and bending angle. The K-nearest neighbor regression model demonstrated the highest accuracy, with a mean absolute error of approximately 11%. These findings underline the importance of precise angle prediction models in enhancing the functionality and reliability of soft grippers, paving the way for their broader application in automation and robotics.
... In addition to the deterministic nature of traditional optimization methods, their ability to use convergence problems and very few design variables make it difficult to be successful in such complex optimization problems . In such problems, as the complexity increases, the transition from traditional methods to optimization methods, including evolutionary algorithms, may offer significant opportunities to reach a solution (EL-Kribi et al., 2013;Hassan & Abomoharam, 2017). One of the crucial reasons for turning to evolutionary algorithms is their global search capabilities. ...
Article
Robot Gripper Design Optimization Problem is a well-known multi-objective optimization problem explored by various approaches in previous studies over the last few decades. The continuous parameter z plays an important role in determining a robot gripper's minimum and maximum gripping forces. However, previous studies have utilized a simple loop approach in which the value of z is iteratively increased by a constant value from 0 to 50 to determine the gripping forces of a solution. This approach limits the solution space by treating the continuous z parameter as a discrete variable. To overcome this limitation, we propose a nested optimization approach in which a single-objective optimizer is embedded within the fitness function of the main optimizer to determine the optimal value of z. The proposed approach is evaluated using a single-optimizer Simulated Annealing (SA) nested within multi-objective optimizers such as Multi-Objective Gray Wolf Optimizer (MOGWO), Multi-Objective Particle Swarm Optimization (MOPSO), Multi-Objective Artificial Algae Algorithm (MOAAA), and Non-Dominated Sorting Genetic Algorithm (NSGA-II). Experimental results demonstrate that our approach significantly outperforms previous studies and achieves the best-known performance. The findings of the study indicate that this method can present a novel framework for researchers in tackling design optimization challenges, encouraging a reevaluation and possible improvement of conventional techniques.
Article
Günümüzde endüstriyel sistemlerde nesnelerin kavranması, taşınması ve sabitlenmesi için kullanılan robot tutucular önemli araçlar olarak öne çıkmaktadır. Özellikle robotik sistemlerde, bir nesneyi en az manevrayla zarar vermeden tutabilme yeteneği büyük önem taşımaktadır. Bundan dolayı, son yıllarda robot tutucularının tasarım optimizasyonu ilgi çeken bir araştırma konusu haline gelmiştir. Bu çalışmada bu tasarım problemi için aritmetik optimizasyon algoritması (AOA) iyileştirilmiş ve çok stratejili aritmetik optimizasyon algoritması (ÇSAOA) adında yeni bir algoritma önerilmiştir. Bu algoritmada hem orijinal AOA’nın güncelleme mekanizmasını modifiye edilmiş, hem de farklı bir güncelleme mekanizması eklenilerek kendinden uyarlanabilen bir algoritma haline getirilmiştir. Bu yaklaşım, en iyi güncelleme stratejisine odaklanarak problemi daha verimli bir şekilde çözmeye olanak sağlamıştır. ÇSAOA, robot tutucu problemine uygulandığında, orijinal algoritmaya göre hem performans hem de hesaplama süresi açısından daha iyi sonuçlar ürettiği gözlemlenmiştir. Ayrıca, bu yeni algoritma literatürdeki diğer benzer algoritmalarla karşılaştırılmış ve önerilen ÇSAOA’nın daha performanslı algoritma olduğu görülmüştür.
Article
Full-text available
Supernumerary robotic limbs (SRLs) are a new type of wearable robot that add artificial limbs to the human body to perform collaborative tasks. In contrast with exoskeletons, SRLs are kinematically independent of human limbs, allowing the wearer to overcome the limitations of human physiological ability, such as realizing the expansion of space in the human body, rather than enhancing existing limbs. In this study, a lightweight and compact hexagonal reconfigurable lower‐limb SRL system is proposed to assist human locomotion in daily activities, including walking, crouching, and stair climbing. To adapt to multiple scenarios, the hexagonal mechanism can be adjusted to different configurations including convex hexagonal configuration, pentagonal configuration, and concave hexagonal configuration. To achieve more optimized performance of different configurations, the optimization for identifying the optimal dimensions of each link was carried out. Subsequently, the detailed design methodology and specifics are presented. Finally, the load and wearing performance experiments were evaluated. The experiment results demonstrated that the tested maximum load in different configurations exceeded 90% of the simulated value and the entire equipment has a good wearing adaptability. This study may inspire the design of other lower‐limb SRLs and provide efficient solutions for stable support assistance in various scenarios.
Article
While designing an end‐effector tool for a specific task using an already available robotic arm is a common strategy for developing robot‐based solutions, it is subject to the designer's subjective biases, especially in the medical robotics field. Consequently, this approach can result in suboptimal solutions that may limit the performance of the robotic arm and the overall system. This work introduces a novel framework for developing a virtual scenario that enables the robust optimization of design guidelines for robotics‐based medical solutions. The proposed framework aims to address the issue of subjectivity involved in the design process, improving the reliability and accuracy of the robotic system. Specifically, it aims to find the optimized 3D positioning of the Tool Center Point and the robot base that takes full advantage of the robotic arm kinematics. The feasibility of the proposed methodology is validated for designing a system for robotic‐guided vascular laser treatments. First, a simulated treatment environment is constructed to compute the optimal trajectories to treat a wide population of synthetic human models. Next, a design optimization methodology based on robot task performance metrics is developed. Overall, the proposed design guidelines allowed the performance of 94.1% of the treatment trajectories free of collisions with median precision of around 0.5 mm while guaranteeing high manipulability. Thus, proving the added value of the proposed strategy and showing the potential for promoting the development of optimized robotics‐based solutions.
Conference Paper
Full-text available
This paper presents the design and the realization of a novel robot gripper which can be used in various automation processes including "pick and place" operation. This project is mainly developed for educational and research purposes, especially for mechatronic and robotics students. The gripper consists of four fingers which move simultaneously allowing the gripper to open and close. It is a single degree of freedom (DOF) system. One finger is driven by an electrical motor via a four-bar mechanism. The motion is transmitted from the driver finger to the other three fingers by slider-crank mechanism. This gripper has two closed loop structure, a study is performed to determine its geometric and kinematic models. This leads to find a geometrical solution space, which is verified via a CAD model of the gripper. Considering some given dimensional constraints, the acceptable functional dimensions and the singular configurations are determined. The finger profile is designed to grasp diverse object shapes. In order to validate our design, a prototype is developed and tested. The student should analyze the system and propose other solutions in order to reply to pre-defined objectives. Thus, the gripper design will be modified or optimized, whether at the level of the mechanism embodiment design and/or at the level of the finger profile. The mechanical design and the electrical circuit of the gripper prototype are also described. © 2014 The Authors. Published by Elsevier B.V. Selection and peer-review under responsibility of the International Scientific Committee of "24th CIRP Design Conference" in the person of the Conference Chairs Giovanni Moroni and Tullio Tolio.
Chapter
Product family design is a well recognised method to address the demands of mass customisation. A potential drawback of product families is that the performance of individual members are reduced because of the constraints added by the common platform, i.e., parts and components need to be shared by other family members. This chapter presents a framework where the product family design problem is stated as a multi-objective optimisation problem and where multi-objective evolutionary algorithms are applied to solve the problem. The outcome is a Pareto-optimal front that visualises the trade-off between the degree of commonality (e.g., number of shared components) and performance of individual family members. The design application is a family of industrial robots. An industrial robot is a mechatronic system that comprises a mechanical structure (i.e., a series of mechanical links), drive-train components (including motors and gears), electrical power units and control software for motion planning and control.
Article
This paper exploits screw theory expressed via unit dual quaternion representation and its algebra to formulate both the forward (position+velocity) kinematics and pose control of an n-dof robot arm in an efficient way. Efficiency is in less computer memory usage, in fast computation of the equations, in singularity-free representation of task space, in robustness to numerical errors, and in compactness of the representations. The formulation is simple, intuitive and straightforward to implement. We validated this formulation experimentally on a 7 dof robot arm.
Article
Compared to single arm robot system, dual arm robot has the ability of performing human-like dexterity and cooperation. Dual arm cooperative operation has attracted more and more attention in industrial applications, such as in assembly of complex parts, manufacturing tasks and handling objects. A unified dynamic control method, which is divided into three modes, namely, independent mode, dependent mode, and half dependent mode, is proposed for a redundant dual arm robot with focus on the movement and force of the desired task being operated. Attention is devoted to develop a unified formulation of the above three modes. In addition, a closed form of inverse kinematic solution instead of numerical integration approach is proposed with the aim to guarantee position accuracy. Different from traditional dynamic controllers, where the independent redundancy resolution is obtained based on particular velocity or acceleration levels, here the two dynamic controllers are improved by combining a closed form of inverse kinematic solution with velocity and acceleration levels. Furthermore, the theoretical results of the proposed control method are validated by simulations and experiments.
Article
In this paper new multicriteria design optimization methods are discus-sed. These methods are evolutionary algorithm based methods, and their aim is to make the process of generating the Pareto front very effective. Firstly, the multistage evolutionary algorithm method is presented. In this method, in each stage only a bicriterion optimization problem is solved and then an objective function is transformed to the constrain function. The process is repeated till all the objective functions are considered. Secondly, the preference vector method is presented. In this method, an evolutionary algorithm finds the ideal vector. This vector provides the decision maker with the information about possible ranges of the objec-tive functions. On the basis of this information the decision maker can establish the preference vector within which he expects to find a preferred solution. For this vector, a set of Pareto solutions is generated using an evolutionary algorithm based method. Finally, the method for selecting a representative subset of Pareto solutions is discussed. The idea of this method consists in reducing the set of Pareto optimal solutions using the indiscernibility interval method after running a certain number of gene-rations. To show how the methods discussed work each of them in turn is applied to solve a design optimization problem. These examples show clearly that using the proposed methods the computation time can be re-duced significantly and that the generated solutions are still on the Pareto front.
Article
High precision is still one of the challenges when parallel kinematic machines are applied to advanced equipment. In this paper, a novel planar 2-DOF parallel kinematic machine with kinematic redundancy is proposed and a method for redundant force optimization is presented to improve the precision of the machine. The inverse kinematics is derived, and the dynamic model is modeled with the Newton–Euler method. The deformations of the kinematic chains are calculated and their relationship with kinematic error of the machine is established. Then the size and direction of the redundant force acting on the platform are optimized to minimize the position error of the machine. The dynamic performance of the kinematically redundant machine is simulated and compared with its two corresponding counterparts, one is redundantly actuated and the other is non-redundant. The proposed kinematically redundant machine possesses the highest position precision during the motion process and is applied to develop a precision planar mobile platform as an application example. The method is general and suitable for the dynamic modeling and redundant force optimization of other redundant parallel kinematic machines.
Article
Abstract—This paper ,presents ,the kinematic ,and ,dynamic modeling,of a ,three degrees of freedom ,3-RPS parallel robot. The orientation and ,position degrees of freedom ,of the ,plate- form of this robot are coupled, that leads to complicated kinematic,and ,dynamic ,models. We ,propose ,to exploit ,the architectural characteristics of the mechanism ,to give ,a closed form solution to these problems. Keywords: parallel robot, Jacobian matrix, kinematic modeling, dynamic modeling.