21.03.2015 Views

Dynamic Performance of a SCARA Robot Manipulator With ...

Dynamic Performance of a SCARA Robot Manipulator With ...

Dynamic Performance of a SCARA Robot Manipulator With ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

206 IEEE TRANSACTIONS ON ROBOTICS, VOL. 25, NO. 1, FEBRUARY 2009<br />

<strong>Dynamic</strong> <strong>Performance</strong> <strong>of</strong> a <strong>SCARA</strong> <strong>Robot</strong> <strong>Manipulator</strong><br />

<strong>With</strong> Uncertainty Using Polynomial Chaos Theory<br />

Philip Voglewede, Anton H. C. Smith, and Antonello Monti<br />

Abstract—This short paper outlines how polynomial chaos theory (PCT)<br />

can be utilized for manipulator dynamic analysis and controller design in<br />

a 4-DOF selective compliance assembly robot-arm-type manipulator with<br />

variation in both the link masses and payload. It includes a simple linear<br />

control algorithm into the formulation to show the capability <strong>of</strong> the PCT<br />

framework.<br />

Index Terms—Monte Carlo methods, polynomials, robot dynamics,<br />

uncertainty.<br />

I. INTRODUCTION<br />

The design <strong>of</strong> any mass-produced system involves understanding<br />

how manufacturing variation will affect its performance. For robotic<br />

manipulators, this problem is exacerbated by changes in the payload.<br />

While these affects are minimal for highly geared manipulators, the<br />

affects are large for direct driven manipulators. <strong>Robot</strong>ic designers<br />

need to better understand how variation in both product manufacture<br />

and payload affects the dynamic performance <strong>of</strong> these types <strong>of</strong><br />

manipulators.<br />

Unfortunately, there is a lack <strong>of</strong> appropriate tools to generally assess<br />

the dynamic performance <strong>of</strong> mechanisms with uncertainty. There<br />

has been significant work on the study <strong>of</strong> how clearances statically<br />

affect the output <strong>of</strong> a mechanism [1]–[4]. There are also numerous<br />

computer-aided tolerancing tools that can perform static stack-ups via<br />

linearization or a statistical Monte Carlo (MC) analysis [5]–[8]. However,<br />

there is a lack <strong>of</strong> appropriate tools to simulate and predict the<br />

window <strong>of</strong> possible dynamic configurations resulting from the variation.<br />

As such, variational dynamic simulations are typically performed<br />

as an MC <strong>of</strong> the governing equations that, in turn, results in large run<br />

times.<br />

Polynomial chaos theory (PCT) can be used for uncertain dynamic<br />

analysis, including the controller design. PCT allows one to solve<br />

stochastic differential equations by using orthogonal polynomials in<br />

conjunction with a Galerkin projection. This allows the transformation<br />

<strong>of</strong> stochastic differential equations into an expanded set <strong>of</strong> standard<br />

ordinary differential equations [9]–[12]. PCT has been successfully applied<br />

to fluid mechanic [13], general oscillatory [14], heat transfer [15],<br />

and electrical [16] systems. More recently, it has been applied specifically<br />

to multibody dynamic systems [17]–[19], but specific examples<br />

outlining uses in this realm are lacking.<br />

This paper applies PCT to a selective compliance assembly robot<br />

arm (<strong>SCARA</strong>)-type manipulator to show how the methodology can<br />

Manuscript received May 25, 2007; revised September 2, 2008. First published<br />

January 13, 2009; current version published February 4, 2009. This paper<br />

was recommended for publication by Associate Editor I. Chen and Editor F.<br />

Park upon evaluation <strong>of</strong> the reviewers’ comments. This work was supported by<br />

the U.S. Office <strong>of</strong> Naval Research under Grant N00014-02-1-0623.<br />

P. Voglewede is with the Mechanical Engineering Department, Marquette<br />

University, Milwaukee, WI 53132 USA (e-mail: philip.voglewede@marquette.<br />

edu).<br />

A. H. C. Smith is with the Electrical Engineering Department, University <strong>of</strong><br />

South Carolina, Columbia, SC 29208 USA (e-mail: smith35@cec.sc.edu).<br />

A. Monti is with the Institute for Automation <strong>of</strong> Complex Power Systems at<br />

the EON Energy Research Center, Rhenish-Westphalian Technical University<br />

(RWTH) Aachen, Aachen, Germany (e-mail: amonti@eonerc.rwth-aachen.de).<br />

Color versions <strong>of</strong> one or more <strong>of</strong> the figures in this paper are available online<br />

at http://ieeexplore.ieee.org.<br />

Digital Object Identifier 10.1109/TRO.2008.2006871<br />

be utilized to better understand and predict dynamic performance <strong>of</strong><br />

manipulators under feedback control with uncertainty.<br />

II. PCT BASICS<br />

The basis <strong>of</strong> PCT is the expansion all the generic variables in the<br />

governing equations in terms <strong>of</strong> a orthogonal polynomial basis. Following<br />

a notation similar to [10], a generic variable u can be expressed<br />

as a infinite sum <strong>of</strong> orthogonal polynomials as<br />

u(β) =u 0 I 0 +<br />

∞∑<br />

u i1 I 1 (ξ i1 (β))<br />

i 1 =1<br />

+<br />

∞∑<br />

i 1 =1<br />

i 1 ∑<br />

i 2<br />

u i1 i 2<br />

I 2 (ξ i1 (β),ξ i2 (β)) + ··· (1)<br />

where β is the random event, ξ = {ξ 1 ,ξ 2 ,...,ξ n } is the random vector,<br />

I i is the orthogonal polynomial in the Askey scheme, and u i , u ij ,<br />

etc., are denoted as the PC coefficients. Since this notation is rather<br />

cumbersome to deal with, it is usually abbreviated as<br />

u(β) =<br />

∞∑<br />

u j Φ j (ξ(β)) (2)<br />

j =0<br />

where u j are the PC coefficients and Φ j (ξ) are the multidimensional<br />

orthogonal polynomials. For practical purposes the polynomial expansion<br />

is truncated to a finite number <strong>of</strong> polynomial terms and the<br />

dependence on the uncertain event β is assumed. How many terms are<br />

kept is dependent upon the number <strong>of</strong> uncertainties in the system n v<br />

and the order <strong>of</strong> the adopted polynomial n p . 1 The total number <strong>of</strong> terms<br />

<strong>of</strong> the expansion P is given by<br />

P =<br />

( (np + n v )!<br />

n p !n v !<br />

)<br />

− 1. (3)<br />

The expansion <strong>of</strong> the variables leaves the unknown vector ξ that<br />

must be dealt with. However, due to the special nature <strong>of</strong> the orthogonal<br />

polynomials, the Galerkin projection <strong>of</strong> the expanded equations will<br />

eliminate the dependence upon the unknown stochastic variable ξ.This<br />

is done by defining the inner product as<br />

∫<br />

〈f(ξ)g(ξ)〉 = f(ξ)g(ξ)W (ξ) dξ (4)<br />

where W (ξ) is the weighting function associated with the chosen basis<br />

in the Askey scheme (see [10] for detailed descriptions). Due to the<br />

special nature <strong>of</strong> the inner product, many <strong>of</strong> the inner products will be<br />

zero as<br />

〈Φ i Φ j 〉 = 〈Φ 2 i 〉δ ij (5)<br />

where δ ij is the Kronecker delta. The results are an expanded set<br />

<strong>of</strong> equations that can be solved using traditional techniques for the<br />

unknown PC coefficients. The PC coefficients can then be backsubstituted<br />

into (2) to yield the distribution <strong>of</strong> the original variables<br />

in time.<br />

A. PCT in <strong>Manipulator</strong> <strong>Dynamic</strong> Analysis<br />

Unfortunately, some nonlinearities cannot be solved using PCT, as<br />

the Galerkin projection does not eliminate the dependence upon the<br />

unknown variables ξ i . For multidimensional serial manipulators, the<br />

general form <strong>of</strong> the dynamic equations <strong>of</strong> motion follow [20]<br />

M ¨θ + C(θ, θ) ˙ ˙θ + G(θ) =τ (6)<br />

1 See [10] for a discussion on the convergence <strong>of</strong> this series.<br />

1552-3098/$25.00 © 2009 IEEE<br />

Authorized licensed use limited to: RWTH AACHEN. Downloaded on February 18, 2009 at 03:05 from IEEE Xplore. Restrictions apply.


IEEE TRANSACTIONS ON ROBOTICS, VOL. 25, NO. 1, FEBRUARY 2009 207<br />

TABLE I<br />

<strong>SCARA</strong> PARAMETERS FOR NUMERICAL STUDY<br />

Fig. 1. Schematic <strong>of</strong> the <strong>SCARA</strong> robot showing the geometric parameters and<br />

coordinate frames.<br />

where M is the mass matrix, C is a vector <strong>of</strong> Coriolis and centrifugal<br />

terms, G is the gravitational terms, and τ is a vector <strong>of</strong> generalized<br />

forces. These terms, in general, contain polynomial and trigonometric<br />

nonlinearities. Polynomial nonlinearities cause no problems to the formulation.<br />

Trigonometric functions can be solved easily using a Taylor<br />

series expansion with a small-angle approximation, as will be shown<br />

later in this paper.<br />

B. PCT in Control Analysis<br />

The addition <strong>of</strong> control laws to a serial manipulator simply changes<br />

the right-hand side <strong>of</strong> (6). For example, using independent potential<br />

difference (PD) control on the joints causes no issues when applying<br />

PCT. Nonlinear control laws can be more vexing if the nonlinearity<br />

is not polynomial or trigonometric based. Care must be exercised to<br />

ensure that additional nonlinearities are not introduced by division as<br />

well.<br />

III. PCT APPLIED TO <strong>SCARA</strong> ROBOT<br />

In this paper, we will assume a 4-DOF <strong>SCARA</strong>-type manipulator<br />

having variation in the mass (and subsequently, the inertia) <strong>of</strong> both the<br />

first two links as well as payload, as shown in Fig. 1 and Table I. Variation<br />

in the link lengths and centers <strong>of</strong> mass could also be incorporated<br />

into the model but are left out as their effects are assumed to be small<br />

compared to the mass variation. The other mechanical parameters are<br />

set to reasonable values, as given in Table I. The variation is assumed<br />

to be uniformly distributed across the range. 2<br />

A. <strong>Dynamic</strong>s <strong>of</strong> the <strong>SCARA</strong> <strong>Robot</strong><br />

An open kinematic chain’s dynamics can be derived utilizing either<br />

a Newton–Euler or Lagrangian formulation. Utilizing a Lagrangian<br />

formulation, the equations <strong>of</strong> motion for the <strong>SCARA</strong> robot as shown<br />

in Fig. 1 are found to be [21], [22]<br />

⎡<br />

⎤ ⎡ ⎤<br />

p 1 + p 2 c 2 p 3 +0.5p 2 c 2 0 −p 5<br />

¨θ 1<br />

p 3 +0.5p 2 c 2 p 3 0 −p 5<br />

¨θ 2<br />

⎢<br />

⎣ 0 0 p 4 0<br />

⎥ ⎢<br />

⎦ ⎣<br />

¨θ<br />

⎥<br />

3 ⎦<br />

−p 5 −p 5 0 p 5<br />

¨θ 4<br />

} {{ }<br />

M<br />

⎡ ⎤ ⎡<br />

⎤ ⎡ ⎤<br />

τ 1 −p 2 s 2 ˙θ2 −0.5p 2 s 2 ˙θ2 0 0 ˙θ<br />

⎡ ⎤<br />

1 0<br />

τ<br />

=<br />

2<br />

⎢<br />

⎣ τ<br />

⎥<br />

3 ⎦ − 0.5 ∗ p 2 s 2 ˙θ1 0 0 0<br />

˙θ 2<br />

⎢<br />

⎣ 0 0 0 0<br />

⎥ ⎢<br />

⎦ ⎣<br />

˙θ<br />

⎥<br />

3 ⎦ + 0<br />

⎢ ⎥<br />

⎣ p 4 g ⎦<br />

τ 4 0 0 0 0 ˙θ 4<br />

0<br />

} {{ }<br />

F<br />

(7)<br />

where c 2 and s 2 are cos(θ 2 ) and sin(θ 2 ), respectively, and<br />

p 1 =<br />

4∑<br />

I i + m 1 x 2 1 + m 2 (x 2 2 + a 2 1 )+(m 3 + m 4 )(a 2 1 + a 2 2 )<br />

i=1<br />

p 2 =2[a 1 x 2 m 2 + a 1 a 2 (m 3 + m 4 )]<br />

4∑<br />

p 3 = I i + m 2 x 2 2 + a 2 2 (m 3 + m 4 )<br />

i=2<br />

p 4 = m 3 + m 4<br />

p 5 = I 4 (8)<br />

where τ i is the input torque (or force), I i is the moments <strong>of</strong> inertia<br />

around the centroid, m i is the mass, x i is the mass center, and a i is the<br />

length for link i. In order to include the affect <strong>of</strong> mass on the inertia,<br />

the inertia terms are further defined as<br />

I i = m i κ 2 i , i =1,...,4 (9)<br />

where κ i is the radius <strong>of</strong> gyration <strong>of</strong> the links. Assuming a PD control<br />

law on each joint, the torques are<br />

τ i = K pi (θ ides − θ i )+K di ( ˙θ ides − ˙ θ i ), i =1,...,4 (10)<br />

2 It should be noted that virtually any type <strong>of</strong> distribution and even mixed<br />

distributions (one variable being normally distributed and one being uniform)<br />

can be incorporated into the framework.<br />

where des denotes the desired values.<br />

In order to use a general nonlinear ordinary differential equation<br />

solver, the equations are expressed in state space form [23]. For this<br />

particular problem, the governing equations can be put into eight<br />

Authorized licensed use limited to: RWTH AACHEN. Downloaded on February 18, 2009 at 03:05 from IEEE Xplore. Restrictions apply.


208 IEEE TRANSACTIONS ON ROBOTICS, VOL. 25, NO. 1, FEBRUARY 2009<br />

first-order equations with<br />

⎡ ⎤ ⎡ ⎤<br />

u 1 θ 1<br />

u 2<br />

θ 2<br />

u 3<br />

θ 3<br />

u 4<br />

θ 4<br />

u ≡<br />

=<br />

u 5<br />

˙θ 1<br />

u 6<br />

˙θ 2<br />

⎢ ⎥ ⎢ ⎥<br />

⎣ u 7 ⎦ ⎣ ˙θ 3 ⎦<br />

u 8<br />

˙θ 4<br />

where<br />

⎡<br />

˙u =<br />

⎢<br />

⎣<br />

and M and F are defined in (7).<br />

u 5<br />

u 6<br />

u 7<br />

u 8<br />

M −1 F<br />

⎤<br />

⎥<br />

⎦<br />

(11)<br />

(12)<br />

B. <strong>SCARA</strong> <strong>Dynamic</strong> Equations Incorporating Uncertainty<br />

The governing equations can then be expanded using the PCT framework.<br />

The benefit <strong>of</strong> using PCT is that it allows one to solve the stochastic<br />

differential equations with an expanded differential equation set. In<br />

other words, the random distributions need to be calculated only once<br />

and the ODE solver needs to run only once. This dramatically improves<br />

the efficiency <strong>of</strong> the solver. The trade<strong>of</strong>f is that the number <strong>of</strong> differential<br />

equations as well as the calculation <strong>of</strong> numerous inner products<br />

are much larger. Thus, as the number <strong>of</strong> unknown variables becomes<br />

large, the method becomes cumbersome, and a traditional MC would<br />

be better [24]. However, once the basis is chosen, these inner products<br />

need to be calculated only once.<br />

The first step is to express all our variables in the appropriate basis.<br />

For this particular example, a Legendre polynomial basis is chosen due<br />

to the uniform distribution on the inputs. 3 Each state space variable<br />

(u i ) is expanded in terms <strong>of</strong> the corresponding PC coefficients (u ij )<br />

u i = u i0 (t)+u i1 (t)ξ 1 + u i2 (t)ξ 2 + u i3 (t)ξ 3<br />

i =1,...,8. (13)<br />

To keep the notation simpler, the use <strong>of</strong> the hat is omitted and the<br />

dependence <strong>of</strong> the polynomial coefficients on time will be dropped.<br />

For simplicity, the polynomial basis is truncated at first order. As will<br />

be shown in this example, the results will concur with this assumption.<br />

We can now proceed to substitute the approximations into (12). The<br />

difficulty is that the equations contain trigonometric functions. These<br />

trigonometric functions will cause problems with the methodology and<br />

must be expressed as a linear combination <strong>of</strong> the stochastic variables<br />

ξ 1 , ξ 2 ,andξ 3 . Using a Taylor’s series approximation method as in [11],<br />

one can approximate the trigonometric functions as<br />

cos u 2 ≈ cos u 20 − sin u 20 (u 21 ξ 1 + u 22 ξ 2 + u 23 ξ 3 )<br />

sin u 2 ≈ sin u 20 +cosu 20 (u 21 ξ 1 + u 22 ξ 2 + u 23 ξ 3 ) (14)<br />

where a first-order polynomial approximation is utilized in this<br />

case. Higher order approximations will not cause problems with the<br />

formulation.<br />

3 It should be noted that the use <strong>of</strong> the Legendre may not be the most appropriate<br />

as the output variations may be better represented by a Gaussian distribution,<br />

and thus, a Hermite basis may be better [10].<br />

TABLE II<br />

PD CONTROLLER GAINS FOR NUMERICAL STUDY<br />

Now that all the nonlinearities have been eliminated, the equation<br />

can be projected onto the basis. These inner products are cumbersome<br />

to compute as they contain a large quantity <strong>of</strong> terms. In order to avoid<br />

mistakes, these were computed using an automated script in MAPLE<br />

[25]. Due to the projection, four equations are created for every row in<br />

(12) for a total <strong>of</strong> 32 equations. These 32 equations are now integrated<br />

using a standard ODE solver (in this case, MATLAB’s ode45 function<br />

was utilized) to solve for the 32 unknown PC coefficients (i.e., u ij )as<br />

functions <strong>of</strong> time. The time history <strong>of</strong> these can then be substituted back<br />

into (13) to determine the mean and variation. Due to the uncertain<br />

variables ξ i in (13), a “mini-MC” is performed on the output. An<br />

analytical solution to the problem has also been proposed in [26].<br />

IV. RESULTS<br />

In order to show the power <strong>of</strong> the PCT formulation, the <strong>SCARA</strong><br />

robot was put under closed-loop control with the parameters shown in<br />

Table II. The first two joints were commanded to go to 30 ◦ and 20 ◦ (i.e.,<br />

u 1des =0.5236 rad and u 2des =0.3491 rad). Controller gains were<br />

chosen to show differing types <strong>of</strong> output (e.g., overdamped versus<br />

underdamped, large variation versus small variation).<br />

The results <strong>of</strong> this analysis are shown in Fig. 2. An MC analysis<br />

was also performed to compare with the solution. Both the mean and<br />

plus/minus three standard deviations (±3σ), were plotted to show the<br />

resulting variation. A normal plot (not shown) was created at several<br />

time intervals to observe the normality <strong>of</strong> the output distribution. The<br />

normal plot showed that the distributions are normal with shortened<br />

tails. Thus, the ±3σ curves represent more <strong>of</strong> the distribution and<br />

slightly overestimate the variation that does not cause a problem for<br />

this study.<br />

Analysis <strong>of</strong> the plots show that the PCT analysis allows for the<br />

dynamic analysis <strong>of</strong> the stochastic differential equations <strong>of</strong> motion,<br />

and in this case, yields nearly identical results to the MC simulation.<br />

For this particular problem, the difference is shown in Table III. The<br />

most drastic difference is shown in joint 3 where some “noise” is seen<br />

in the PCT formulation. More accurate results can be garnered by<br />

adding more terms on the PCT expansion [27]. Also <strong>of</strong> interest is that<br />

the transition between underdamped and overdamped oscillations in<br />

θ 2 are handled easily. Also, the amount <strong>of</strong> variation if large or small is<br />

easily accommodated by the framework as shown in the graph for θ 3 .<br />

PCT has traded the number <strong>of</strong> ODE solver calculations by expanding<br />

the number <strong>of</strong> differential equations from 8 to 32 and requiring the<br />

calculation <strong>of</strong> 32 inner products, 16 <strong>of</strong> which were trivial. Once the<br />

inner products are determined, the analysis takes significantly less time<br />

as compared to the MC analysis. For a sample size <strong>of</strong> 1000, the MC<br />

simulation took approximately 159.70 s to compute while the PCT<br />

method took a mere 1.27 s. Thus, scenario analyses can be performed<br />

easily in real time without having to run numerous datasets.<br />

It is interesting to note the existence <strong>of</strong> times where the total variation<br />

decreases significantly. Further numerical analysis shows that at these<br />

points, the percentage error in the standard deviation estimate is the<br />

greatest. These situations (here called variation nodes) could possibly<br />

be useful for system testing and evaluation.<br />

The use <strong>of</strong> PCT in the controller design allows control analysis<br />

quicker than a traditional MC. Different controller gains can easily be<br />

tried and the resulting dynamics can be simulated. The PCT procedure<br />

Authorized licensed use limited to: RWTH AACHEN. Downloaded on February 18, 2009 at 03:05 from IEEE Xplore. Restrictions apply.


IEEE TRANSACTIONS ON ROBOTICS, VOL. 25, NO. 1, FEBRUARY 2009 209<br />

Fig. 2.<br />

Simulation results for all four joints <strong>of</strong> a <strong>SCARA</strong> robot with variation in the first two link masses and the payload.<br />

TABLE III<br />

MAXIMUM ABSOLUTE DIFFERENCE BETWEEN PCT<br />

AND MC FOR <strong>SCARA</strong> ROBOT<br />

6) PCT has no problem with both underdamped and overdamped<br />

type responses.<br />

Additionally, this example also corroborates several <strong>of</strong> the known<br />

advantages regarding the use <strong>of</strong> PCT. Specifically, PCT is more efficient<br />

from a simulation time standpoint, allows dynamic changing <strong>of</strong><br />

parameters during the simulation, and yields a richer result that can be<br />

utilized in other applications like controller design.<br />

also opens the door for easy optimization <strong>of</strong> the controller gains for<br />

the small amount <strong>of</strong> variation. Sensitivity studies can quickly and<br />

effectively be performed for different controller gains.<br />

V. CONCLUSION<br />

This paper has shown in detail how to use PCT to analyze the<br />

dynamic response <strong>of</strong> an open-loop mechanism by applying it to a<br />

<strong>SCARA</strong> robot manipulator. Through this particular example, several<br />

items were found.<br />

1) PCT on the <strong>SCARA</strong> robot is feasible as long as the number <strong>of</strong><br />

unknowns are small. Automating the process further aiding in<br />

speeding up the process. However, this even has its limits as<br />

even the automated process is still cumbersome.<br />

2) Using PCT on robotic applications requires a judicious choice <strong>of</strong><br />

states and formulation <strong>of</strong> the problem to make sure that nonlinearities<br />

are not introduced into the equations.<br />

3) PCT on a <strong>SCARA</strong> robot gives consistent results to a large MC<br />

for a simple one-term expansion. This was true even using a<br />

standard approximation <strong>of</strong> the trigonometric identities.<br />

4) PCT exhibits “antinodes” (positions in the output where the<br />

variation decreases significantly), even in this robotics example.<br />

However, this was not noted on all joints.<br />

5) PCT can be utilized with feedback control. However, integral<br />

and nonlinear control poses some interesting problems that have<br />

yet to be solved.<br />

REFERENCES<br />

[1] D. Hoeltzel and W.-H. Chieng, “A unified approach to the kinematic<br />

analysis <strong>of</strong> joint clearances and link length tolerances for determination<br />

<strong>of</strong> the rotational and positional accuracy <strong>of</strong> planar mechanisms,” in Proc.<br />

Adv. Des. Autom. Conf. (ASME Des. Eng. Div.), Montreal, QC, Canada,<br />

Sep. 1989, vol. 19-3, pp. 345–356.<br />

[2] J. Rhyu and B. Kwak, “Optimal stochastic design <strong>of</strong> four-bar mechanisms<br />

for tolerance and clearance,” J. Mechanisms, Transmiss., Autom. Des.,<br />

vol. 110, no. 3, pp. 255–262, Sep. 1988.<br />

[3] S. Lee and B. Gilmore, “The determination <strong>of</strong> the probabilistic properties<br />

<strong>of</strong> velocities and accelerations in kinematic chains with uncertainty,” J.<br />

Mech. Des., vol. 113, no. 1, pp. 84–90, Mar. 1991.<br />

[4] P. Voglewede and I. Ebert-Uph<strong>of</strong>f, “Application <strong>of</strong> workspace generation<br />

techniques to determine the unconstrained motion <strong>of</strong> parallel manipulators,”<br />

ASME J. Mech. Des., vol. 126, no. 2, pp. 283–290, Mar. 2003.<br />

[5] B. Fischer, Mechanical Tolerance Stackup and Analysis. New York:<br />

Marcel Dekker, 2004.<br />

[6] F. Chiesi and L. Governi, “Tolerance analysis with eM-TolMate,” J.<br />

Comput. Inf. Sci. Eng., vol. 3, pp. 100–105, Mar. 2003.<br />

[7] Z. Shen, “Tolerance analysis with EDS/VisVSA,” J. Comput. Inf. Sci.<br />

Eng., vol. 3, no. 1, pp. 95–99, 2003.<br />

[8] Z. Shen, G. Ameta, J. Shah, and J. Davidson, “A comparative study <strong>of</strong><br />

tolerance analysis methods,” J. Comput. Inf. Sci. Eng., vol. 5, no. 3,<br />

pp. 247–256, Sep. 2005.<br />

[9] N. Wiener, “The homogeneous chaos,” Amer. J. Math., vol. 60, pp. 897–<br />

936, 1938.<br />

[10] D. Xiu and G. Karniadakis, “The Wiener–Askey polynomial chaos for<br />

stochastic differential equations,” SIAM J. Sci. Comput., vol. 24, no. 2,<br />

pp. 619–644, 2003.<br />

[11] B.J.Debusschere,H.N.Najm,P.P.Pebayt,O.M.Knio,R.G.Ghanem,<br />

and O. P. Le Maitre, “Numerical challenges in the use <strong>of</strong> polynomial chaos<br />

Authorized licensed use limited to: RWTH AACHEN. Downloaded on February 18, 2009 at 03:05 from IEEE Xplore. Restrictions apply.


210 IEEE TRANSACTIONS ON ROBOTICS, VOL. 25, NO. 1, FEBRUARY 2009<br />

representations for stochastic processes,” SIAM J. Sci. Comput., vol. 26,<br />

no. 2, pp. 698–719, 2005.<br />

[12] X. Wan and G. E. Karniadakis, “Beyond Wiener–Askey expansions: Handling<br />

arbitrary PDFs,” J. Sci. Comput., vol. 27, no. 1–3, pp. 455–464, Jun.<br />

2006.<br />

[13] D. Xiu, D. Lucor, G. E. Su, and C.-H. Karniadakis, “Stochastic modeling<br />

<strong>of</strong> flow-stucture interactions using generalized polynomial chaos,” J.<br />

Fluids Eng., vol. 124, pp. 51–59, Mar. 2002.<br />

[14] D. Lucor, C.-H. Su, and G. Karniadakis, “Generalized polynomial chaos<br />

and random oscillators,” Int. J. Numer. Methods Eng., vol. 60, no. 3,<br />

pp. 571–596, May 2004.<br />

[15] D. Xiu and G. Karniadakis, “Modeling uncertainty in steady state diffusion<br />

problems via generalized polynomial chaos,” Comput. Methods Appl.<br />

Mech. Eng., vol. 191, pp. 4927–4948, 2002.<br />

[16] A. Monti, F. Ponci, and T. Lovett, “A polynomial chaos theory approach<br />

to uncertainty in electrical engineering,” in Proc. Int. Conf. Intell. Syst.<br />

Appl. Power Syst., Arlington, VA, Nov. 2005, pp. 534–539.<br />

[17] A. Sandu, C. Sandu, and M. Ahmadian, “Modeling multibody dynamic<br />

systems with uncertainties. Part I: Theoretical and computational aspects,”<br />

Multibody Syst. Dyn., vol. 15, no. 4, pp. 369–391, May 2006.<br />

[18] A. Sandu, C. Sandu, and M. Ahmadian, “Modeling multibody dynamic<br />

systems with uncertainties. Part II: Numerical applications,” Multibody<br />

Syst. Dyn., vol. 15, no. 3, pp. 241–262, Apr. 2006.<br />

[19] L. Li, C. Sandu, and A. Sandu, “Modeling and simulation <strong>of</strong> a full<br />

vehicle with parametric and external uncertainties,” in Proc. ASME<br />

Int. Mech. Eng. Congr. Expo., Nov. 2005, vol. 118A, no. 1, pp. 209–<br />

214.<br />

[20] J. Craig, Introduction to <strong>Robot</strong>ics: Mechanics and Control, 3rd ed.<br />

Reading, MA: Addison-Wesley, 2005.<br />

[21] L.-A. Dessaint, M. Saad, and B. Hébert, K. Al Haddad, “An adaptive<br />

controller for a direct-drive <strong>SCARA</strong> robot,” IEEE Trans. Ind. Electron.,<br />

vol. 39, no. 2, pp. 105–111, Apr. 1992.<br />

[22] L.-A. Dessaint, M. Saad, and B. Hébert, C. Gargour, “An adaptive controller<br />

for a direct-drive <strong>SCARA</strong> robot: Analysis and simulation,” in Proc.<br />

16th Annu. Conf. IEEE Ind. Electron. Soc., Pacific Grove, CA, Nov. 1990,<br />

vol. 1, pp. 414–420.<br />

[23] J. Ginsberg, Mechanical and Structural Vibrations: Theory and Applications.<br />

New York: Wiley, 2001.<br />

[24] P. Prempraneerach, “Uncertainty analysis in a shipboard integrated<br />

power system using multi-element polynomial chaos,” Ph.D. dissertation,<br />

Massachusetts Inst. Technol., Cambridge, MA, 2007.<br />

[25] A. Smith, A. Monti, and F. Ponci, “Robust stability and performance<br />

analysis using polynomial chaos theory,” in Proc. Soc. Comput. Simul.<br />

Summer Conf., Jul. 2007, pp. 45–52.<br />

[26] G. D’Antona, A. Monti, F. Ponci, and L. Rocca, “Maximum entropy analytical<br />

solution for stochastic differential equations based on the Wiener-<br />

Askey polynomial chaos,” IEEE Trans. Instrum. Meas., vol. 56, no. 3,<br />

pp. 689–695, Jun. 2007.<br />

[27] R. Field and M. Grigoriu, “On the accuracy <strong>of</strong> the polynomial chaos<br />

approximation,” Probablistic Eng. Mech., vol. 19, pp. 65–80, 2004.<br />

Authorized licensed use limited to: RWTH AACHEN. Downloaded on February 18, 2009 at 03:05 from IEEE Xplore. Restrictions apply.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!