You are on page 1of 9

ARTICLE IN PRESS

Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525

Contents lists available at ScienceDirect

Robotics and Computer-Integrated Manufacturing


journal homepage: www.elsevier.com/locate/rcim

Improving feasibility of robotic milling through robot


placement optimisation
George-Christopher Vosniakos n, Elias Matsas
National Technical University of Athens, School of Mechanical Engineering, Manufacturing Technology Division, Heroon Polytehneiou 9, 15780 Athens, Greece

a r t i c l e in fo abstract

Article history: Milling performed with robots is quite demanding, even for low-strength materials, due to the high
Received 21 May 2008 accuracy requirements, the generally high and periodically varying milling forces and the low stiffness
Received in revised form of robots compared to CNC machine tools. In view of the generally improved recently robot stiffness, it
12 April 2010
is desirable to perform the milling operation in regions of the robot’s workspace where manipulability,
Accepted 12 April 2010
both kinematic and dynamic, is highest, thereby exhausting the robot’s potential to cope with the
process. In addition, by selecting the most suitable initial pose of the robot with respect to the
Keywords: workpiece, a reduction in the range of necessary joint torques may be reached, to the extent of
Robotic milling alleviating the heavy requirements on the robot. Two genetic algorithms (GAs) are employed to tackle
Genetic algorithms
these problems. The values of several robot variables, such as joint positions and torques, which are
Inverse dynamics
needed by the genetic algorithms, are calculated using inverse kinematics and inverse dynamics
Manipulability
models. In addition, initial positions and poses leading to singularities along the milling path are
penalized and, thus, avoided. The first GA deals solely with robot kinematics to maximize
manipulability. The second GA takes into account milling forces, which are computed numerically
according to the particular milling parameters, to minimise joint torque loads.
& 2010 Elsevier Ltd. All rights reserved.

1. Introduction for instance, an approximation of the minimum of the manip-


ulator velocity ratio along the end-effector path [7]. This is used as
The relative position of a robot with respect to the workpiece the objective function in a genetic algorithm searching for the
and the corresponding path have attracted researchers attention optimal location of a straight-line path of a PUMA-like robot.
starting in 1993 when optimum robot location was linked to Manipulability based optimisation of robot location is implemen-
manipulability/dexterity, avoidance of singular points and joint ted in [8] for a two-degree-freedom-robot. An elitist GA searches
motion constraints [1] and task and obstacle spaces were defined for the valid solution in the reachable area of the robot base.
using Monte Carlo simulation [2]. For planar 3R robots path A similar approach, with a numerical optimisation technique
placement in the robot workspace was found by systematically instead of the GA, and a new dexterity measure is advocated in
defining suitable positions separately for the first joint and then [9]. More recently, a hybrid heuristic method that combines a
for the other two [3], still without a clear optimisation flair. These genetic algorithm, a quasi-Newton algorithm and a constraints
works pointed at welding and pick-and-place applications. Later, handling method optimises base position and joint angles of a
spot-welding tasks were certainly a motivation for optimising spatial robot, in order to avoid the singular configurations when
robot placement. Chen and Tseng [4] plan a near-optimal path the end-effector poses are prescribed, i.e. in the presence of severe
and, then, use GAs to determine the location of the workpiece that constraints [10].
satisfies shortest path requirement between workpoints that Note that none of the above research papers deals with robotic
needed to be visited by the robot. Simulated annealing was also milling, as such. The use of an industrial robot arm with an end mill
used to minimise cycle time, as an aid to CAD-based robot attached as end-effector is relatively straightforward for rapid
programming [5]. prototyping [11] or milling soft materials [12], i.e. in the presence
Cycle time, however, recognised by industry was not to be the of very low cutting forces. However, scarcity of robotic milling
best criterion in optimising performance [6], robot location being literature for normal milling tasks, which are usually performed on
recognised as equally important. Various criteria were linked to it, CNC machine tools, reflects the restricted use of robots in such
tasks, until recently. This is primarily due to two issues: accuracy/
repeatability and stiffness of the milling robot. Accuracy required
n in milling is quite high compared to robot repeatability. In addition,
Corresponding author.
E-mail addresses: vosniak@central.ntua.gr (G.-C. Vosniakos), matsasme@hol.gr robot stiffness was, until recently, very low—lower than that of a
(E. Matsas). CNC machine tool by orders of magnitude, which caused large

0736-5845/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.rcim.2010.04.001
ARTICLE IN PRESS
518 G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525

deflections; hence, it caused large potential position errors, at the


end effector in the presence of even medium milling forces. For
instance, in robotic milling of aluminium alloys with articulated
milling robots of the previous generation robot stiffness values in
the range 20–85 mm/N were reported, depending on force direction
[13]. In view of accuracy/repeatability inadequacy, the resolution of
robot axes plays a secondary role, especially by current industrial
robot specifications.
As far back as 1990 it was tried to compensate for position errors
in robotic milling, based on process dynamics models and force–
motion control [14]. Similarly, calibration, sensor-based teaching
and post-process compensation were employed to minimise
position errors and enhance repeatability [15]. Accuracy compensa-
tion was advocated through regulation of the deformation of the
arm [13]. A custom serial–parallel structure robot was used for
milling and polishing tasks with in-program switching of kinematic
model parameters as obtained by suitable calibration [16].
Fig. 1. Milling force pattern.
Parallel kinematic chains offer advantages over serial robots
for milling tasks [17]. However, recent advances in robot
tangential, radial and axial directions as Ktc ¼1825, Krc ¼ 770,
construction achieve high stiffness serial configurations, with
Kac ¼735 N/mm2) and Ke ¼63.15 N/mm (analysed in the
local enhancement of stiffness in crucial axes, typical values
tangential, radial and axial directions as Kte ¼29.7, Kre ¼55.7,
reported in the order of 1 N/mm. Corresponding models are
Kae ¼1.8 N/mm) [20]. A softer material has lower cutting force
employed in industry for pre-machining of castings [18].
coefficients, e.g. an 6061-T6 aluminium alloy cut by a 2-flute
Several decisions may alleviate the inherent difficulty caused
25.4 mm end mill has Kc ¼693.8 N/mm2 and Ke ¼18.98 N/mm
by milling force being of relatively high magnitude and varying
[21], i.e. about 1/3 of those for the titanium case. Softer materials
both in magnitude and direction. One decision is to perform the
will have even lower specific cutting force coefficients. Therefore,
operation in regions of the robot workspace where manipul-
the cutting forces presented in Fig. 1 should be treated as a typical
ability, both static and dynamic, is highest. Another decision is to
pattern corresponding to various combinations of workpiece
select the most suitable initial position and pose of the robot with
material, axial depth of cut and feed, the product of the latter
respect to the workpiece, so as to reduce the range of necessary
being considered approximately constant for the same tool and
joint torques. This work deals with these two decisions and
entry–exit angles. However, for a typical static stiffness of 1 N/
suggests that genetic algorithms can be employed to make best
use of emerging new generation milling robots.
mm, a maximum cutting force of 145 N parallel to the milled
surface and 45 N in the axial direction, see Fig. 1, would cause a
The work is presented in the following manner. In Section 2, a
static deflection of 145 and 45 mm, respectively. Therefore, if
brief presentation is given on the calculation of milling forces. In
static deflection is an indication of dynamic deflection and
Section 3, the optimisation problem is described. In Section 4, the
machining error, then this level of milling forces would make
robot simulation model is explained. The optimisation system and
sense in a rough milling context with sizeable radial and axial
results obtained are presented in Section 5 making use of a robot
depth of cut, i.e. with softer materials and/or lower feed, still with
simulation model presented in Section 4. Conclusions are drawn
low accuracy requirements. This is exactly the area of application
in Section 6.
characterised as pre-machining [18].
The particular force values at specific points in time, as needed,
2. Milling forces are digitized off the extended pattern, see Fig. 1. The cutting path
measures 33.2 mm and it is covered in 10 s.
Milling forces are mostly calculated using semi-mechanistic
methods. These are based on the idea of discretising the cutting
edge and applying the classic cutting force equations that are 3. The optimisation problem
valid for flat endmills and constant radial and axial depth of cut as
well as constant entry and exit angle. In these approaches, milling 3.1. Manipulability concepts
force component coefficients have to be experimentally estab-
lished for each cutter design. These coefficients may then be used Singularities represent robot configurations at which mobility
in the mechanistic models for predicting the instantaneous force of the structure is reduced and in the neighbourhood of a
components [19]. singularity small velocities in the operational space may cause
The calculation process for the milling forces is presented in large velocities in joint space. In addition, infinite solutions to the
Appendix A. The input to this process comprises the following inverse kinematics problem may exist at a singularity [22].
variables, the specific values used in the test case being given in At singular points, a manipulator effectively loses one or more
parentheses: helix angle b, entry and exit angles jst and jex, the degrees of freedom, so actions of the manipulator are said to be
axial depth of cut a (mm), the number of teeth N, the feed poorly conditioned. The farther the distance of the manipulator
c (mm/tooth/rev), the tool rotational speed n (rpm), the tool from singularities, the more uniform its movement in all
diameter D (mm) and the cutting coefficients Kc, Ke (specific directions and the better its ability to apply forces [22]. Several
cutting forces) [19]. manipulability measures have been suggested to quantify this
A software program was created to calculate cutting forces as effect, effectively defining well-conditioned subspaces of the
above. Using this program the forces shown in Fig. 1 were workspace.
obtained for titanium alloy Ti6Al4V as workpiece material, Yoshikawa’s manipulability measure is based purely on
b ¼30o, jst ¼01 and jex ¼1801, a ¼0.5 mm, N¼4, c¼0.1 (mm/ kinematic data, and gives an indicator of how ‘‘far’’ the
tooth/rev), n ¼500, D¼19 mm, Kc ¼2113 N/mm2 (analysed in the manipulator is from singularities [22]. This is defined as
ARTICLE IN PRESS
G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525 519
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
! ! ! !
wð q Þ ¼ det½Jð q ÞUJ T ð q Þ where q is the joint variables and J the execution of the milling operation. For instance, for some poses
Jacobian matrix. In the case of a non-redundant manipulator (J is a the torque on the joints of the robot or its increments, which are
! !
square matrix), w reduces to wð q Þ ¼ 9detðJð q ÞÞ9. necessary to cope with the milling force, may be too high. In
Yoshikawa’s manipulability is non-negative; it is equal to zero contrast, for other poses torques or their increments may be at a
at a singular configuration. High values of w define a good tolerable level.
manipulator configuration. In addition, some poses of the manipulator may lead to
Asada’s manipulability measure is based on acceleration singularities and low manipulability measures, including poor
analysis or force-application capability. Asada suggested the force application capabilities, while others can give high kine-
examination of the eigenvalues of the Cartesian mass matrix matic and dynamic manipulability measures.
! ! ! !
Bx ð x Þ ¼ J T ð q ÞUBð q ÞUJ 1 ð q Þ as a measure of how well the Analytical optimisation methods are not suitable for optimis-
manipulator can accelerate in various Cartesian directions [22]. ing the initial end-effector position in the robot workspace,
A graphic representation of this measure is an inertia ellipsoid, because the search space is too large. Furthermore, computational
!T ! !
given by x UBx ð x ÞU x ¼ 1 which is the equation of an load would become intolerable, especially considering that, for
!
n-dimensional ellipse, where n is the dimension of x . The scalar each possible solution, robot manipulability and joint torques
manipulability measure that is used in this work is the ratio of the would need to be computed at a large number of points along the
smallest to the largest ellipsoid axes, symbolised by n. Well- allocated milling path. Therefore, genetic algorithm based
conditioned points in the manipulator workspace are characterised optimisation is advocated in this work.
by inertia ellipsoids that are spherical, giving an ideal ratio of 1. Robot’s trajectory, workspace and exact mathematical model
(including inverse kinematics and inverse dynamics) are impli-
3.2. The optimisation problem cated in the genetic algorithm, allowing to numerically simulate
the robot motion along a very large number of different milling’s
The end-effector (cutting tool) tip in robotic milling coincides trajectories. Each of these trajectories corresponds to a different
with the starting point of the milling path on the workpiece. relative position and orientation of the milling path with respect
However, the position and orientation of the workpiece and, to the robot, which is stochastically selected by the genetic
consequently, of the milling path with respect to the base algorithm.
(reference frame) of the robot may vary infinitely. The orientation The two optimisation criteria, i.e. manipulability and torque
of the milling path may be fixed in order to reduce computational adequacy, are considered via two separate genetic algorithms. The
load, if this does not preclude any good solutions. This, in the first deals with initial configuration and kinematics of the robot to
particular case examined, is justified by the symmetry of the maximize its kinematic and dynamic manipulability along
robot workspace about the vertical axis, see Fig. 2. the end-effector’s path. The second genetic algorithm searches
Finding the optimum location of the milling path with respect for the optimal initial position of the end-effector which will
to the robot is equivalent to finding the optimum initial position result in the minimal total torque variation range in the
of the robot end-effector, from which the milling trajectory manipulator joints along the milling path.
should start, with respect to the (global) reference frame. This Total torque range is used instead of maximal torque because
position also dictates the initial pose of the robot and the series of during the milling task joint torques vary intensely with time.
subsequent poses, which correspond to the robot trajectory. Thus, the resulting torque ‘‘wavelets’’ can take both positive and
In this light, the initial position of the robot end effector may negative values. In this case, torque variation can be better
be favourable or, at the other extreme, detrimental to the expressed by the total torque range than by maximum absolute
torque value.

4. Robot modelling

The robot used in the test-case examined is a fictitious one, see


Fig. 2, with six revolute axes. The respective D–H parameters are
shown in Table 1. The robot model was built on Robotics Toolbox
for MatlabTM [23]. It is assumed that milling is feasible with this
robot, i.e. its payload, accuracy, repeatability and resistance to
chatter are acceptable.
The model developed determines kinematic properties (posi-
tions, velocities and accelerations) of the robot in both joint and
Cartesian space and dynamic properties (joint torques), given the
end effector (cutting tool) trajectory in Cartesian space, which in
this case, is a straight line, and the milling force variations, as
calculated in Section 2.

Table 1
Denavit–Hartenberg parameters.

Joint i ai-1 ai-1 di hi

1 0 0 0 y1
2  90o 0 0 y2
3 0 a2 d3 y3
4  90o a3 d4 y4
5 90o 0 0 y5
Fig. 2. Robot configuration and local coordinate systems according to D–H 6  90o 0 0 y6
convention.
ARTICLE IN PRESS
520 G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525

The problem was simplified by assuming single point contact 5. Genetic optimisation
between the cutting tool and the workpiece. Thus, moments on
the end effector are neglected. A further simplification stems from GAs are a generic search and optimisation method mimicking
the fact that the cutting tool remains vertical throughout the biological evolution. The main functions of a GA are depicted in
cutting path, i.e. the last three joints of the robot keep the same Fig. 7 and technical detail can be found in [24]. GAs involve
angular position, see Fig. 2. populations of chromosomes, each chromosome representing a
The start and end of the cutting path, hence the corresponding possible solution to the problem being investigated. Genetic
positions of the end effector are given as vectors. First the operators are applied to a population in order to transform it to a
homogenous transformation corresponding to the position and better-fitter population of chromosomes-solutions. Fitness of a
orientation of the end effector at the start and at the end of the chromosome is the measure of its optimality. Classic genetic
cutting path are calculated. operators are selection, mutation and crossover operators. Each
The Cartesian trajectory between start and end points of population results from the previous one after application of the
the cutting path is created, involving definition of the time vector, genetic operators during a cycle termed generation. The final
which in this case is from 0 to 10 s, with a certain sampling solution is picked as the fittest chromosome of the population
period, which in this case is 0.028 s resulting in 358 time-point usually reached after tens or even hundreds of generations.
samples. A three-dimensional array, in this case [4  4  358], Two GAs were setup on MatlabTM Genetic Algorithm Toolbox.
defines the position and orientation of the end-effector for The Genetic Algorithm Toolbox is a collection of routines which
all sampled time-points. An example of the Cartesian trajectory implement the most important functions in genetic algorithms,
of the end-effector along X Cartesian axis is shown in Fig. 3. such as population generation, selection, crossover, mutation,
This as well as the calculation result examples presented in reinsertion, etc. The user begins by creating the initial population
Figs. 4–6 correspond to the Cartesian trajectory from point and the chromosome encoding (binary, integer, etc.). Once the
(0.65, 0.15, 0.0) to point (0.6832, 0.15, 0.0), the end effector objective and fitness functions are defined, selection method and
orientation, in terms of roll, pitch and yaw angles, being rate have to be chosen. Then the user can define crossover,
constant at (0,  p, 0). mutation and reinsertion parameters and rates. A common
Through inverse kinematics, the joint angles in the initial pose practice is to terminate the GA after a pre-specified number of
are readily calculated. Similarly, the joint angles corresponding to generations and then test the quality of the best members of the
the consecutive positions of the end effector (at the sampled time- population against the problem definition.
points) along its trajectory in Cartesian space, are also calculated, The GAs that were setup, are as follows:
see Fig. 4(a).
The Jacobian matrices corresponding to each configuration of (i) The first algorithm searches for the optimum initial position
the joints are calculated next. Following the inverse differential of the end-effector, which imparts maximum kinematic and
kinematics equation joint velocities are calculated based on the dynamic manipulability over the whole duration of the
desired end-effector velocity, in this case 0.00332 m/s, see milling task. This position is particular to the specific robot
Fig. 4(b). and it is related to the path followed and not to the milling
Singularity of each configuration is identified by checking the force.
rank and the determinant of the Jacobian. Therefore, Yoshikawa’s (ii) The second algorithm searches for the optimal initial position
and Asada’s manipulability measures can be computed at this of the end effector, which will result in the minimum total
stage. torque range in the robots joints 1, 2 and 3 over the whole
Joint accelerations are calculated numerically as time deriva- duration of the milling task. Milling forces do influence the
tives of joint velocities, see Fig. 5. Having calculated positions, choice of these positions, along with the robots characteristics
velocities and accelerations of the joints, and given the milling and the path followed.
forces and the gravity vector, the joint torques are calculated
without and with milling forces by employing the reverse
dynamics function, see Fig. 6. 5.1. Optimisation based on manipulability

The problem is defined as finding the initial position of the


end-effector (coinciding with the start of the milling path) that
will maximize the mean kinematic manipulability (Yoshikawa)
and the dynamic manipulability (Asada) in the set of discrete-
sampled points of the milling trajectory.
The chromosome consists of the three Cartesian coordinates of
the end-effector (cutting tool) tip, expressed in the Cartesian
coordinate system of the first joint (base coordinate system of the
robot), as shown in Fig. 2, i.e. three real values. Hence, real
encoding was preferred to binary, as it is more natural:
xEE ,yEE ,zEE
Note that the straight line trajectory of the end effector
(cutting tool tip) is along x-axis. It is not necessary to include in
the chromosome the end effector orientation angles, e.g. Euler
angles, because of the assumption that the end-effector will
always be vertical.
The range of Cartesian coordinates, i.e. maximum and mini-
mum, within which the end effector position may lie are
presented in Table 2. These are the limits of the values that the
Fig. 3. Cartesian path of end-effector in X-direction versus time. chromosome genes acquire in the course of the optimisation.
ARTICLE IN PRESS
G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525 521

Fig. 4. Variation with time of (a) joint angular position and (b) joint angular velocity for joints 2, 3 and 5. For the rest joints mean position is J1:  1.5  10  4 rad,
J4:  1.65  10  16 rad, J6:  1.5  10  4 rad and mean velocity is J1:7.5  10  7 rad/s, J4:9  10  19 rad/s, J6:7.5  10  7 rad/s.

The initial population was chosen to consist of 100 chromo- Note that many of the initial positions ‘suggested’ by the GA
somes. After several trials, the stopping criterion was set at 100 correspond to singular configurations, thereby blocking the
generations. Selection ratio was 60% (elitist re-insertion) cross- inverse kinematics function and, consequently, the GA to
over rate was 80% and mutation rate 5%. The mutation value range terminate. To overcome this problem the ikine function of Matlab
is the same as that set originally for the genes. Robotics Toolbox was slightly modified, in order to let the
The objective function expresses kinematic and dynamic evolution proceed by just penalising the respective solution with
manipulability in the whole of the trajectory, sampled at 50 a high objective function value.
points. The best objective function value, as well as the mean Fig. 8 depicts the evolution process. A relatively fast
value of the population, are recorded during the evolution. ‘convergence’ is noticed after about 30 generations. The
A medium-term tendency of the mean value to reach the best optimum initial position of the end-effector was computed as
value of the objective function is regarded as a sign of xEE ¼0.6368, yEE ¼ 0.2558, zEE ¼0.2329 in meters.
‘convergence’ of the evolution process to the optimum.
Due to the significance of milling forces and their variation,
5.2. Optimisation based on joint torques.
dynamic manipulability was given twice the weight of the
kinematic manipulability in the objective function. Maximising
The aim of this GA is to minimise the torque range of the joints
manipulability is transformed into minimising its negative. Thus
manifested over the whole trajectory.
Objective function ¼ w-2 n A significant simplification stems from the fact that due to
constant end-effector orientation along the trajectory, torque
The objective function executes the simulation model requirements for the last three joints are negligible. This was
presented in Section 4, except that the initial position of the observed in trial runs of the simulation model of Section 4, see
end-effector is not given by the user, but it is taken from the Fig. 9. By contrast, torque values for the first three joints were
chromosomes of the current population each time they need significant and exhibited intense variability. These observations
evaluation in order to compute fitness values. result in a reduced need for computation, i.e. just three torque
ARTICLE IN PRESS
522 G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525

Fig. 5. Variation with time of joint angular acceleration.

values are computed, instead of six, for each of the discretised assigning an immense torque range (10,000 N m) to essentially
trajectory points. exclude the respective solution from the evolution process.
The chromosome is the same as in the first GA presented. The result is depicted in Fig. 9(a). Slow improvement of the
The initial population was chosen to consist of 100 chromosomes. solution is noted even after 200 generations. Note the initial
After several trials, the stopping criterion was set at 200 generations. torque range of 120 N m and the finally reached mean torque
Selection ratio was 60%, i.e. elitist re-insertion was implemented, range of 32 N m, i.e. an improvement by 390%. In particular, after
crossover rate was 90% and mutation rate was 10%. The mutation generation 166 the best individual exhibited a torque range of
value range is the same as that set originally for the genes. 14.02 N m. The optimum position in this case was computed as
The value of the objective function is the sum of the absolute xEE ¼ 0.4806 m, yEE ¼0.5731 m, zEE ¼0.2794 m.
value of the torque ranges in all three significant joints:

Objective function ¼ 9Torque_range_J19 þ 9Torque_range_J29 6. Discussion and conclusions

þ9Torque_range_J39 Application of GAs results in considerable improvement in


For each potential solution (chromosome) the trajectory is manipulability and in significant reduction in joint torques.
generated as in the previous GA and, in addition, the torques of In practice, this means that a smaller robot may be proved to
joints 1, 2 and 3 are computed. Discretisation interval is now half suffice for the particular milling operation.
of that used previously, corresponding to a time step of 0.1 s. This Execution time for the first GA was about 230 min, whereas for
is deemed small enough to enable precise recording of the torque the second GA it was 3880 min, on a normal laptop computer
variation, at the same time avoiding lengthy computation times. with a 1.8 GHz CPU and 1024 MB RAM. This performance may
Singular points are treated by analogy to the previous GA by seem disappointing at first sight, but it has to be taken into
ARTICLE IN PRESS
G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525 523

Fig. 6. Joint torques required (a) without milling force consideration (b) with milling force interaction. Torque in joints 3, 4 and 5 are negligible.

account that in each generation all offspring created by the Both genetic algorithms make use of the robot model
genetic operators need to be evaluated, through the model presented above. The first GA can be used in a preliminary
constructed in the robotics toolbox, each evaluation taking a investigation when just the desired milling path is known,
few seconds of CPU time. without considering the particular external forces applied to the
The value of the main parameters used in the GAs was end effector. The second GA can be employed when interaction
determined in a trial and error fashion. For instance, in the second forces need to be examined, too.
GA, the mutation rate was changed by one order of magnitude, Several other criteria could be used instead of manipulability
the result being shown in Fig. 9(b). A mutation rate of 1% leads and torque range, or, indeed, in addition to them in a weighted
to faster convergence, albeit to a local minimum, since the sum fashion or, if desired, in a Pareto front optimisation exercise.
corresponding value of the objective function is much higher than Such criteria may be maximum torque for each joint, maximum
that achieved with mutation rate equal to 10%. acceleration/deceleration for each joint, etc.
This work deals with robot placement in a different way than In addition, constraints such as maximum velocity, accelera-
the papers reported in literature so far. Most previous research tion, etc. may be directly taken into account in the optimisation
work optimises manipulability, in particular kinematic, whereas function, in the form of penalties. In the case study presented
this work considers dynamic manipulability to be more signifi- above milling feed was small enough to render such considera-
cant. However, the main point made in this work is that tions unnecessary.
optimisation should tackle the essence of the robotic operation, The GA approach can also be applied to wider definitions of the
which in this case is the large (for the typical magnitude of robot optimisation problem. For instance, the straight line milling path
stiffness) and varying forces at the end effector, which create may not be constrained to lie along the x direction or on the
stringent requirements on torques at the joints. Minimisation of horizontal plane either. In this case, the chromosome should be
joint torque or, better still, torque variation, is a more concrete expanded to contain the end point of the milling path, too.
criterion than manipulability and, therefore, more useful. As far as milling strategy is concerned, note that the end-effector
orientation was considered constant – always vertical – therefore
ARTICLE IN PRESS
524 G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525

Fig. 7. Standard GA functional diagram.

Table 2
Cartesian coordinate limits for the end effector. Fig. 9. Evolution of torque minimisation solution with population ¼ 100, selection
rate¼ 0.6, crossover rate¼ 0.9 and (a) mutation rate¼ 0.1, (b) mutation rate¼ 0.01.
X Y Z

Max value (m) 0.864 0.864 0.864


larity of cutting tool axis and machined surface, can be easily
Min value (m) 0.15 0.15  0.679 included in the GA setup. However, in that case, a more general
algorithm would be necessary for computing milling force when
tool engagement changes constantly, especially when spherical
end mills are used, as in [19].
Although a fictitious robot was used as a demonstrator, it is
clear that in recent years robotic milling of metals and alloys are a
reality [18]; therefore, the concepts demonstrated in this work
can be applied in real-world problems, too.
Further considerations, such as the position error induced due
to robot compliance, may be pertinent and could also be included
in the objective function, provided a suitable model was available
for evaluating them. In a similar direction, chatter issues can also
be considered. Chatter presence in CNC milling is certainly a big
issue regarding restriction of the depth of cut, i.e. extension of
cutting time. In robotic milling chatter-related restrictions are
aggravated due to much lower stiffness and lower mass of the
robot compared to a machine tool, as already pointed out in
Section 1. According to established the chatter theory [25], the
minimum value of the real part of the frequency response
function (FRF) Gmin dictates the limiting depth of cut. The overall
Fig. 8. Evolution of the best and mean solution for manipulability. real part of the FRF is composed of the respective FRFs of the
individual modes of the structure Gi multiplied by direction
P
(scaling) factors: G¼ (Giui). Each direction factor ui expresses
the chromosome was shorter and the computational load was the influence of the angle b between the cutting force and the
lower. In spite of that, variable orientation of the end-effector, normal to the milled surface and the influence of the angles yi
such as in sculptured surface finishing maintaining perpendicu- between each individual mode direction and the normal to that
ARTICLE IN PRESS
G.-C. Vosniakos, E. Matsas / Robotics and Computer-Integrated Manufacturing 26 (2010) 517–525 525

surface. Changing the placement of the robot with respect to the dFy,j ðjj ðzÞÞ ¼ þ dFt,j sin jj ðzÞdFr,j cos jj ðzÞ,
o
workpiece causes a change in yi, which results in a different
dFz,j ðjj ðzÞÞ ¼ þ dFa,j ð5Þ
dynamic (chatter) behavior of the system in milling, hence in a
different limiting depth of cut. The methodology presented in this
paper can be used to improve chatter stability of milling robots, The total force along each main direction is the sum of the
provided either that the FRF of the robot can be computed via a corresponding elementary forces coming from all cutting edges.
suitable model or that the individual FRFs representing each Note that they refer to a specific time point or, equivalently, to a
mode are known in order to be multiplied by the respective specific rotation angle of the tool j:
direction factors for each position tried in the genetic algorithm. X
N 1 X
N 1 X
N 1
The objective function in this case should contain the computed Fx ðjÞ ¼ Fx,j , Fy ðjÞ ¼ Fy,j , Fz ðjÞ ¼ Fz,j ð6Þ
allowable depth of cut. This may well be another interesting j¼0 j¼0 j¼0
future step of this work.
References
Appendix A. End-mill force calculations
[1] Zeghloul S, Pamanes-Garcia JA. Multi-criteria optimal placement of robots in
constrained environments. Robotica 1993;11(2):105–10.
For a typical end mill with helix angle b, milling forces are [2] Tu Q, Rastegar J. Determination of allowable manipulator link shapes; and
calculated by adding up the individual force components task, installation, and obstacle spaces using the Monte Carlo method.
corresponding to each one of the discrete cutting edge parts. Transactions of the ASME Journal of Mechanical Design 1993;115(3):
457–61.
Axial depth of cut is assumed to be constant. The milling [3] Soman NA, Davidson JK. A two-dimensional formulation for path-placement
parameters for one slice are given in Fig. 10. Starting with one in the workcells of planar 3-R robots. Transactions of the ASME Journal of
of the cutting teeth tip as reference, assumed to be at angle j, the Mechanical Design 1995;117(3):479–84.
[4] Chen C-J, Tseng C-S. The path and location planning of workpieces by genetic
tooth angular pitch being jp, the tips of each of the rest of the algorithms. Journal of Intelligent Manufacturing 1996;7(1):69–76.
teeth j are located at angles given as: [5] Barral D, Perrin J-P, Dombre E, Liegeois A. Development of optimisation tools
in the context of an industrial robotic CAD software product. International
jj ð0Þ ¼ j þ jjp , j ¼ 0, 1, 2, :::, N1 ð1Þ Journal of Advanced Manufacturing Technology 1999;15(11):822–31.
[6] Mahr C. Beyond the standard cycle: increasing total work cell productivity by
At axial depth of cut z, hysterisis angle is c ¼ kbz, where optimising robot placement and other key performance influencers. Indus-
kb ¼(2 tan b)/z. Thus, a cutting point on tooth j and axial depth of trial Robot 2000;27(5):334–7.
cut z corresponds to an angle [7] Aspragathos NA, Foussias S. Optimal location of a robot path when
considering velocity performance. Robotica 2002;20(2):139–47.
jj ðzÞ ¼ j þ jjp kb z ð2Þ [8] Tian L, Collins C. Optimal placement of a two-link planar manipulator using a
genetic algorithm. Robotica 2005;23(2):169–76.
The tangential, radial and axial components of the milling [9] Abdel-Malek K, Yu W, Yang J. Placement of robot manipulators to maximize
force acting on a elementary part of the cutting edge with length dexterity. International Journal of Robotics and Automation 2004;19(1):
6–15.
dz are expressed as
[10] Mitsi S, Bouzakis K-D, Sagris D, Mansour G. Determination of optimum robot
dFt,j ðj,zÞ ¼ ½Ktc hj ðjj ðzÞÞ þ Kte dz, base location considering discrete end-effector positions by means of hybrid
genetic algorithm. Robotics and Computer-integrated Manufacturing
dFr,j ðj,zÞ ¼ ½Krc hj ðjj ðzÞÞ þ Kre dz, 2008;24(1):50–9.
o [11] Walstra WH, Bronsvoort WF. Interactive simulation of robot milling for rapid
dFa,j ðj,zÞ ¼ ½Kac hj ðjj ðzÞÞ þ Kae dz, ð3Þ shape prototyping. Computers & Graphics 1994;18(6):861–71.
[12] Chen YH, Hu YN. Implementation of a robot system for sculptured surface
cutting. Part 1: Rough Machining. International Journal of Advanced
Manufacturing Technology 1999;15(9):624–9.
where chip thickness is given as
[13] Matsuoka S, Shimizu K, Yamazaki N, Oki Y. High-speed end milling of an
hj ðj,zÞ ¼ c sin jj ðzÞ ð4Þ articulated robot and its characteristics. Journal of Materials Processing
Technology 1999;95:83–9.
These forces can be projected onto the three main directions [14] Cetinkunt S, Tsai R-L. Position error compensation of robotic contour end-
milling. International Journal of Machine Tools and Manufacture
x, y, z, as
1990;30(4):527–613.
dFx,j ðjj ðzÞÞ ¼ dFt,j cos jj ðzÞdFr,j sin jj ðzÞ, [15] Shirase K, Tanabe N, Hirao M, Yasui T. Articulated Robot Application in
End Milling of Sculptured Surface. JSME International Journal SERIES C,
Dynamics, Control, Robotics, Design and Manufacturing 1996;39(C2):
308–16.
[16] Zielinski C, Mianowski K, Nazarczuk K, Szynkiewicz W. A prototype robot for
polishing and milling large objects. Industrial Robot: An International Journal
2003;30(1):67–76.
[17] Terrier M, Dugas A, Hascoët J-Y. Qualification of parallel kinematics machines
in high-speed milling on free form surfaces. International Journal of Machine
Tools and Manufacture 2004;44:865–77.
[18] Pan Z, Zhang H, Zhu Z, Wang J. Chatter analysis of robotic machining process.
Journal of Materials Processing Technology 2006;173:301–9.
[19] Altintas Y. Manufacturing automation: metal cutting mechanics. Machine
Tool Vibrations and CNC Design. UK: Cambridge University Press; 2000.
[20] Budak E, Altintas Y, Armarego EJA. Prediction of milling force coefficients
from orthogonal cutting data. ASME Journal of Manufacturing Science and
Engineering 1996;118(2):216–23.
[21] Xu M, Jerard RB, Fussell BK. Energy based cutting force model calibration for
milling. Computer-aided Design and Applications 2007;4(1-4):341–51.
[22] Sciavicco L, Siciliano B. Modelling and control of robot manipulators, 2nd ed.
Springer-Verlag; 2000.
[23] Corke PIA. Robotics toolbox for matlab. IEEE Journal of Robotics and
Automation 1996;3(1):24–32.
[24] Gen M, Cheng R. Genetic algorithms and engineering optimization. New
York: Wiley; 2001.
[25] Altintas Y. Manufacturing automation: metal cutting mechanics, machine
tool vibrations and CNC design [chapter 3—static and dynamic deformations
Fig. 10. Geometry of slice j of an end-mill. in machining]. Cambridge University Press; 2000.

You might also like