25.01.2013 Views

A review of robust optimal design and its ... - Michael I Friswell

A review of robust optimal design and its ... - Michael I Friswell

A review of robust optimal design and its ... - Michael I Friswell

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Abstract<br />

A <strong>review</strong> <strong>of</strong> <strong>robust</strong> <strong>optimal</strong> <strong>design</strong><br />

<strong>and</strong> <strong>its</strong> application in dynamics<br />

C. Zang a , M.I. <strong>Friswell</strong> a, *, J.E. Mottershead b<br />

a Department <strong>of</strong> Aerospace Engineering, University <strong>of</strong> Bristol, Queen’s Building, Bristol BS8 1TR, United Kingdom<br />

b Department <strong>of</strong> Engineering, University <strong>of</strong> Liverpool, Brownlow Hill, Liverpool L69 3GH, United Kingdom<br />

Received 13 November 2003; accepted 9 October 2004<br />

The objective <strong>of</strong> <strong>robust</strong> <strong>design</strong> is to optimise the mean <strong>and</strong> minimize the variability that results from uncertainty<br />

represented by noise factors. The various objective functions <strong>and</strong> analysis techniques used for the Taguchi based<br />

approaches <strong>and</strong> optimisation methods are <strong>review</strong>ed. Most applications <strong>of</strong> <strong>robust</strong> <strong>design</strong> have been concerned with static<br />

performance in mechanical engineering <strong>and</strong> process systems, <strong>and</strong> applications in structural dynamics are rare. The<br />

<strong>robust</strong> <strong>design</strong> <strong>of</strong> a vibration absorber with mass <strong>and</strong> stiffness uncertainty in the main system is used to demonstrate<br />

the <strong>robust</strong> <strong>design</strong> approach in dynamics. The results show a significant improvement in performance compared with<br />

the conventional solution.<br />

Ó 2004 Elsevier Ltd. All rights reserved.<br />

Keywords: Robust <strong>design</strong>; Stochastic optimization; Taguchi; Vibration absorber<br />

1. Introduction<br />

Successful manufactured products rely on the best<br />

possible <strong>design</strong> <strong>and</strong> performance. They are usually produced<br />

using the tools <strong>of</strong> engineering <strong>design</strong> optimisation<br />

in order to meet <strong>design</strong> targets. However, conventional<br />

<strong>design</strong> optimisation may not always satisfy the desired<br />

targets due to the significant uncertainty that exists in<br />

material <strong>and</strong> geometrical parameters such as modulus,<br />

thickness, density <strong>and</strong> residual strain, as well as in joints<br />

<strong>and</strong> component assembly. Crucial problems in noise <strong>and</strong><br />

vibration, caused by the variance in the structural<br />

dynamics, are rarely considered in <strong>design</strong> optimisation.<br />

Therefore, ways to minimize the effect <strong>of</strong> uncertainty<br />

* Corresponding author. Fax: +44 117 927 2771.<br />

E-mail address: m.i.friswell@bristol.ac.uk (M.I. <strong>Friswell</strong>).<br />

Computers <strong>and</strong> Structures 83 (2005) 315–326<br />

0045-7949/$ - see front matter Ó 2004 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.compstruc.2004.10.007<br />

www.elsevier.com/locate/compstruc<br />

on product <strong>design</strong> <strong>and</strong> performance are <strong>of</strong> paramount<br />

concern to researchers <strong>and</strong> practitioners.<br />

The earliest approach to reducing the output variation<br />

was to use the Six Sigma Quality strategy [1,2] so<br />

that ±6 st<strong>and</strong>ard deviations lie between the mean <strong>and</strong><br />

the nearest specification limit. Six Sigma as a measurement<br />

st<strong>and</strong>ard in product variation can be traced back<br />

to the 1920s when Walter Shewhart showed that three<br />

sigma from the mean is the point where a process requires<br />

correction. In the early <strong>and</strong> mid-1980s, Motorola<br />

developed this new st<strong>and</strong>ard <strong>and</strong> documented more than<br />

$16 billion in savings as a result <strong>of</strong> the Six Sigma efforts.<br />

In the last twenty years, various non-deterministic<br />

methods have been developed to deal with <strong>design</strong> uncertainties.<br />

These methods can be classified into two approaches,<br />

namely reliability-based methods <strong>and</strong> <strong>robust</strong><br />

<strong>design</strong> based methods. The reliability methods estimate<br />

the probability distribution <strong>of</strong> the systemÕs response


316 C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326<br />

based on the known probability distributions <strong>of</strong> the r<strong>and</strong>om<br />

parameters, <strong>and</strong> is predominantly used for risk<br />

analysis by computing the probability <strong>of</strong> failure <strong>of</strong> a system.<br />

However, the variation is not minimized in the reliability<br />

approaches [3], which concentrate on the rare<br />

events at the tails <strong>of</strong> the probability distribution [4]. Robust<br />

<strong>design</strong> improves the quality <strong>of</strong> a product by minimizing<br />

the effect <strong>of</strong> the causes <strong>of</strong> variation without<br />

eliminating these causes. The objective is different from<br />

the reliability approach, <strong>and</strong> is to optimise the mean performance<br />

<strong>and</strong> minimise <strong>its</strong> variation, while maintaining<br />

feasibility with probabilistic constraints. This is achieved<br />

by optimising the product <strong>and</strong> process <strong>design</strong> to make<br />

the performance minimally sensitive to the various<br />

causes <strong>of</strong> variation. Hence <strong>robust</strong> <strong>design</strong> concentrates<br />

on the probability distribution near to the mean values.<br />

In this paper, <strong>robust</strong> <strong>optimal</strong> <strong>design</strong> methods are <strong>review</strong>ed<br />

extensively <strong>and</strong> an example <strong>of</strong> the <strong>robust</strong> <strong>design</strong><br />

<strong>of</strong> a passive vibration absorber is used to demonstrate<br />

the application <strong>of</strong> these techniques to vibration analysis.<br />

2. The concept <strong>of</strong> <strong>robust</strong> <strong>design</strong><br />

The <strong>robust</strong> <strong>design</strong> method is essential to improving<br />

engineering productivity, <strong>and</strong> early work can be traced<br />

back to the early 1920s when Fisher <strong>and</strong> Yates [5] developed<br />

the statistical <strong>design</strong> <strong>of</strong> experiments (DOE) approach<br />

to improve the yield <strong>of</strong> agricultural crops in<br />

Engl<strong>and</strong>. In the 1950s <strong>and</strong> early 1960s, Taguchi developed<br />

the foundations <strong>of</strong> <strong>robust</strong> <strong>design</strong> to meet the challenge<br />

<strong>of</strong> producing high-quality products. In 1980, he<br />

applied his methods in the American telecommunica-<br />

Fig. 1. Different types <strong>of</strong> performance variations.<br />

tions industry <strong>and</strong> since then the Taguchi <strong>robust</strong> <strong>design</strong><br />

method has been successfully applied to various industrial<br />

fields such as electronics, automotive products,<br />

photography, <strong>and</strong> telecommunications [6–8].<br />

The fundamental definition <strong>of</strong> <strong>robust</strong> <strong>design</strong> is described<br />

as A product or process is said to be <strong>robust</strong> when<br />

it is insensitive to the effects <strong>of</strong> sources <strong>of</strong> variability, even<br />

through the sources themselves have not been eliminated<br />

[9]. In the <strong>design</strong> process, a number <strong>of</strong> parameters can<br />

affect the quality characteristic or performance <strong>of</strong> the<br />

product. Parameters within the system may be classified<br />

as signal factors, noise factors <strong>and</strong> control factors. Signal<br />

factors are parameters that determine the range <strong>of</strong><br />

configurations to be considered by the <strong>robust</strong> <strong>design</strong>.<br />

Noise factors are parameters that cannot be controlled<br />

by the <strong>design</strong>er, or are difficult <strong>and</strong> expensive to control,<br />

<strong>and</strong> constitute the source <strong>of</strong> variability in the system.<br />

Control factors are the specified parameters that the <strong>design</strong>er<br />

has to optimise to give the least sensitivity <strong>of</strong> the<br />

response to the effect <strong>of</strong> the noise factors. For example,<br />

in an automotive body in white, uncertainty may arise in<br />

the panel thicknesses given as the noise factors. The<br />

geometry is then optimised through control factors<br />

describing the panel shape, for the different configurations<br />

to be analysed determined by the signal factors,<br />

such as different applied loads or the response at different<br />

frequencies.<br />

A P-diagram [6] may be used to represent different<br />

types <strong>of</strong> parameters <strong>and</strong> their relationships. Fig. 1 shows<br />

the different types <strong>of</strong> performance variations, where the<br />

large circles denote the target <strong>and</strong> the response distribution<br />

is indicated by the dots <strong>and</strong> the associated probability<br />

density function. The aim <strong>of</strong> <strong>robust</strong> <strong>design</strong> is to make


the system response close to the target with low variations,<br />

without eliminating the noise factors in the system,<br />

as illustrated in Fig. 1(d).<br />

Suppose that y = f(s,z,x) denotes the vector <strong>of</strong> responses<br />

for a particular set <strong>of</strong> factors, where s, z <strong>and</strong> x<br />

are vectors <strong>of</strong> the signal, noise <strong>and</strong> control factors.<br />

The noise factors are uncertain <strong>and</strong> generally specified<br />

probabilistically. Thus z, <strong>and</strong> hence f, are r<strong>and</strong>om variables.<br />

The range <strong>of</strong> configurations are specified by a<br />

set <strong>of</strong> signal factors, V, <strong>and</strong> thus s 2 V. One possible<br />

mathematical description <strong>of</strong> <strong>robust</strong> <strong>design</strong> is then<br />

^x ¼ arg min<br />

x<br />

h i<br />

max<br />

s2V Ez kfðs; z; xÞ tk 2<br />

ð1Þ<br />

subject to the constraints,<br />

gjðs; z; xÞ 6 0 for j ¼ 1; ...; m; ð2Þ<br />

where E denotes the expected value (over the uncertainty<br />

due to z) <strong>and</strong> t is the vector <strong>of</strong> target responses, which<br />

may depend on the signal factors, s. Note that x is the<br />

parameter vector that is adjusted to obtain the <strong>optimal</strong><br />

solution, given by Eq. (1) as the solution that gives the<br />

smallest worse case response over the range <strong>of</strong> configurations,<br />

or signal factors, in the set V. The key step in<br />

the <strong>robust</strong> <strong>design</strong> problem is the specification <strong>of</strong> the<br />

objective function, <strong>and</strong> once this has been done the tools<br />

<strong>of</strong> statistics (such as the analysis <strong>of</strong> variance) <strong>and</strong> the <strong>design</strong><br />

<strong>of</strong> experiments (such as orthogonal arrays) may be<br />

used to calculate the solution.<br />

For convenience, the <strong>robust</strong> <strong>design</strong> approaches will<br />

be classified into three categories, namely Taguchi, optimisation<br />

<strong>and</strong> stochastic optimisation methods. Further<br />

details are given in the following sections. The stochastic<br />

optimisation methods include those that propagate the<br />

uncertainty through the model <strong>and</strong> hence use accurate<br />

statistics in the optimisation. Also included are optimisation<br />

methods that are inherently stochastic, such as<br />

the genetic algorithm. These methods are generally time<br />

consuming, <strong>and</strong> the statistical representation may be<br />

simplified at the expense <strong>of</strong> accuracy, for example by<br />

considering only a linear perturbation from the nominal<br />

values <strong>of</strong> the parameters. The Taguchi methods take this<br />

a step further, using only a limited number <strong>of</strong> discrete<br />

values for the parameters <strong>and</strong> using the <strong>design</strong> <strong>of</strong> experiments<br />

methodology to limit the number <strong>of</strong> times the<br />

model has to be run. The response data then requires<br />

analysis <strong>of</strong> variance techniques to determine the sensitivity<br />

<strong>of</strong> the response to the noise <strong>and</strong> control factors.<br />

3. The Taguchi based methods<br />

TaguchiÕs approach to the product <strong>design</strong> process<br />

may be divided into three stages: system <strong>design</strong>, parameter<br />

<strong>design</strong>, <strong>and</strong> tolerance <strong>design</strong> [6]. System <strong>design</strong> is<br />

the conceptual <strong>design</strong> stage where the system configura-<br />

C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326 317<br />

tion is developed. Parameter <strong>design</strong>, sometimes called<br />

<strong>robust</strong> <strong>design</strong>, identifies factors that reduce the system<br />

sensitivity to noise, thereby enhancing the systemÕs<br />

<strong>robust</strong>ness. Tolerance <strong>design</strong> specifies the allowable<br />

deviations in the parameter values, loosening tolerances<br />

if possible <strong>and</strong> tightening tolerances if necessary [9].<br />

TaguchiÕs objective functions for <strong>robust</strong> <strong>design</strong> arise<br />

from quality measures using quadratic loss functions.<br />

In the extension <strong>of</strong> this definition to <strong>design</strong> optimisation,<br />

Taguchi suggested the signal-to-noise ratio (SNR),<br />

10 log 10(MSD), as a measure <strong>of</strong> the mean squared deviation<br />

(MSD) in the performance. The use <strong>of</strong> SNR in system<br />

analysis provides a quantitative value for response<br />

variation comparison. Maximizing the SNR results in<br />

the minimization <strong>of</strong> the response variation <strong>and</strong> more <strong>robust</strong><br />

system performance is obtained. Suppose we have<br />

only one response variable, y, <strong>and</strong> only one configuration<br />

<strong>of</strong> the system (so the signal factor may be neglected).<br />

Then for any set <strong>of</strong> control factors, x, the<br />

noise factors are represented by n sets <strong>of</strong> parameters,<br />

leading to the n responses, yi. Although there are many<br />

possible SNR ratios, only two will be considered here.<br />

The target is best SNR. This SNR quantifies the deviation<br />

<strong>of</strong> the response from the target, t, <strong>and</strong> is<br />

SNR ¼ 10 log 10ðMSDÞ<br />

¼ 10 log 10<br />

1<br />

n<br />

X<br />

ðyi i<br />

tÞ 2<br />

¼ 10 log 10ðS 2 þðy tÞ 2 Þ ð3Þ<br />

where S is the population st<strong>and</strong>ard deviation. Eq. (3) is<br />

essentially a sampled version <strong>of</strong> the general optimisation<br />

criteria given in Eq. (1). Note that the second form indicates<br />

that the MSD is the summation <strong>of</strong> population variance<br />

<strong>and</strong> the deviation <strong>of</strong> the population mean from the<br />

target. If the control parameters are chosen such that<br />

y ¼ t (the population mean is the target value), then<br />

the MSD is just the population variance. If the population<br />

st<strong>and</strong>ard deviation is related to the mean, then the<br />

MSD may also be scaled by the mean to give<br />

SNR ¼ 10 log 10ðMSDÞ ¼ 10log 10<br />

y<br />

¼ 10log10 2<br />

S 2 : ð4Þ<br />

The smaller the better SNR. This SNR considers the<br />

deviation from zero <strong>and</strong>, as the name suggests, penalises<br />

large responses. Thus<br />

SNR ¼ 10 log 10ðMSDÞ ¼ 10log 10<br />

!<br />

S 2<br />

y 2<br />

1<br />

n<br />

X<br />

i<br />

y 2<br />

i<br />

!<br />

ð5Þ<br />

This is equivalent to the Target-is-Best SNR, with<br />

t =0.


318 C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326<br />

The most important task in TaguchiÕs <strong>robust</strong> <strong>design</strong><br />

method is to test the effect <strong>of</strong> the variability in different<br />

experimental factors using statistical tools. The requirement<br />

to test multiple factors means that a full factorial<br />

experimental <strong>design</strong> that describes all possible conditions<br />

would result in a large number <strong>of</strong> experiments.<br />

Taguchi solved this difficulty by using orthogonal arrays<br />

(OA) to represent the range <strong>of</strong> possible experimental<br />

conditions. After conducting the experiments, the data<br />

from all experiments are evaluated using the analysis<br />

<strong>of</strong> variance (ANOVA) <strong>and</strong> the analysis <strong>of</strong> mean<br />

(ANOM) <strong>of</strong> the SNR, to determine the optimum levels<br />

<strong>of</strong> the <strong>design</strong> variables. The optimisation process consists<br />

<strong>of</strong> two steps; maximizing the SNR to minimize<br />

the sensitivity to the effects <strong>of</strong> noise, <strong>and</strong> adjusting the<br />

mean response to the target response.<br />

TaguchiÕs techniques were based on direct experimentation.<br />

However, <strong>design</strong>ers <strong>of</strong>ten use a computer to simulate<br />

the performance <strong>of</strong> a system instead <strong>of</strong> actual<br />

experiments. Ragsdell <strong>and</strong> dÕEntremont [10] developed<br />

a non-linear code that applied TaguchiÕs concepts to <strong>design</strong><br />

optimisation. In some cases, the <strong>optimal</strong> <strong>design</strong> is<br />

the least <strong>robust</strong>, <strong>and</strong> <strong>design</strong>ers have to make a trade<strong>of</strong>f<br />

between target performance <strong>and</strong> <strong>robust</strong>ness. Ideally,<br />

one should optimise the expected performance over a<br />

range <strong>of</strong> variations <strong>and</strong> uncertainties in the noise factors.<br />

Although TaguchiÕs contributions to the philosophy<br />

<strong>of</strong> <strong>robust</strong> <strong>design</strong> are almost unanimously considered to<br />

be <strong>of</strong> fundamental importance, there are certain limitations<br />

<strong>and</strong> inefficiencies associated with his methods.<br />

Box <strong>and</strong> Fung [11] pointed out that the orthogonal array<br />

method does not always yield the <strong>optimal</strong> solution <strong>and</strong><br />

suggested that non-linear optimisation techniques<br />

should be employed when a computer model <strong>of</strong> the <strong>design</strong><br />

exists. Better results were achieved on a Wheatstone<br />

Bridge circuit <strong>design</strong> problem used by Taguchi. Montgomery<br />

[12] demonstrated that the inner array used for<br />

the control factors in the TaguchiÕs approach <strong>and</strong> the<br />

outer array used for noise factors, is <strong>of</strong>ten unnecessary<br />

<strong>and</strong> results in a large number <strong>of</strong> experiments. Tsui [13]<br />

showed that the Taguchi method does not necessarily<br />

find an accurate solution for <strong>design</strong> problems with<br />

highly non-linear behaviour. An excellent survey <strong>of</strong><br />

these controversies was the panel discussion edited by<br />

Nair [14].<br />

TaguchiÕs approach has been extended in a number<br />

<strong>of</strong> ways. DÕErrico <strong>and</strong> Zaino [15] implemented a modification<br />

<strong>of</strong> the Taguchi method using Gaussian–Hermite<br />

quadrature integration. Yu <strong>and</strong> Ishii [16] used the fractional<br />

quadrature factorial method for systems with significant<br />

nonlinear effects. Otto <strong>and</strong> Antonsson [17]<br />

addressed <strong>robust</strong> <strong>design</strong> optimisation with constraints,<br />

using constrained optimisation methods. Ramakrishnan<br />

<strong>and</strong> Rao [18,19] formulated the <strong>robust</strong> <strong>design</strong> problem<br />

as a non-linear optimization problem with TaguchiÕs loss<br />

function as the objective. They used a Taylor series<br />

expansion <strong>of</strong> the objective function about the mean values<br />

<strong>of</strong> the <strong>design</strong> variables. Other significant publications<br />

in this area include those by Moh<strong>and</strong>as <strong>and</strong><br />

S<strong>and</strong>gren [20], S<strong>and</strong>gren [21], <strong>and</strong> Belegundu <strong>and</strong> Zhang<br />

[22]. Chang et al. [23] extended TaguchiÕs parameter <strong>design</strong><br />

to the notion <strong>of</strong> conceptual <strong>robust</strong>ness. Lee et al.<br />

[24] developed <strong>robust</strong> <strong>design</strong> in discrete <strong>design</strong> space<br />

using the Taguchi method. The orthogonal array based<br />

on the Taguchi concept was utilized to arrange the discrete<br />

variables, <strong>and</strong> <strong>robust</strong> solutions for unconstrained<br />

optimization problems were found.<br />

4. Optimisation methods<br />

The optimisation procedures aim to minimise the<br />

objective functions, such as Eq. (1), directly. The uncertainty<br />

in the noise factors means that the system performance<br />

is a r<strong>and</strong>om variable. One option for the <strong>robust</strong><br />

optimisation is to minimize both the deviation in the<br />

mean value, jlf tj, <strong>and</strong> the variance, r2 f , <strong>of</strong> the performance<br />

function, subject to the constraints, where<br />

lfðx; sÞ ¼Ez½fðs; z; xÞŠ;<br />

r 2<br />

f ðx; sÞ ¼Ez ðfðs; z; xÞ lfðx; sÞÞ 2<br />

h i<br />

: ð6Þ<br />

The quantities <strong>of</strong> mean <strong>and</strong> the st<strong>and</strong>ard deviation <strong>of</strong><br />

system performance, for given signal <strong>and</strong> control factors,<br />

may be calculated if the joint probability density<br />

function (PDF) <strong>of</strong> the noise factors is known. For most<br />

practical applications these PDFs are unknown, but <strong>of</strong>ten<br />

it is assumed that all variables have independent normal<br />

distributions. In this case the joint PDF becomes a<br />

product <strong>of</strong> the individual PDFs. However, evaluating<br />

Eq. (6) is extremely time consuming <strong>and</strong> computationally<br />

expensive <strong>and</strong> approximations using TaylorÕs series<br />

expansions about the mean noise <strong>and</strong> control factors, z<br />

<strong>and</strong> x, may be used. If only the linear terms are retained<br />

in the expansion then the mean <strong>and</strong> variance <strong>of</strong> the response<br />

are readily computed in terms <strong>of</strong> the mean <strong>and</strong><br />

variance <strong>of</strong> the noise factors.<br />

The constraints must also be satisfied. For a worst<br />

case analysis the constraints must be satisfied for all values<br />

<strong>of</strong> control <strong>and</strong> noise factors, <strong>and</strong> the constraint in<br />

Eq. (2) may be approximated [25] as<br />

gjðz; xÞþ X ogj Dzi þ<br />

ozi i<br />

X ogj Dxl 6 0 ð7Þ<br />

oxl l<br />

where the dependence on the signal factors has been<br />

made implicit. The derivatives are evaluated at z <strong>and</strong><br />

x, <strong>and</strong> Dzi <strong>and</strong> Dxl represent the deviations <strong>of</strong> the elements<br />

<strong>of</strong> z <strong>and</strong> x from these means. Because <strong>of</strong> the absolute<br />

values this approximation is likely to be very<br />

conservative. For a statistical analysis the constraint is<br />

not always satisfied, <strong>and</strong> the probability that the con-


straints are satisfied must be chosen a priori. The constraints<br />

become,<br />

gjðz; xÞþkrgjðz; xÞ 6 0 ð8Þ<br />

where rgj is an approximation to the st<strong>and</strong>ard deviation<br />

<strong>of</strong> the j th constraint <strong>and</strong> k is a constant that reflects the<br />

probability that the constraint will be satisfied. For<br />

example, k = 3 means that a constraint is satisfied<br />

99.865% <strong>of</strong> the time, if the constraint distributions are<br />

normal. The constraint variance for a linear Taylor series<br />

is<br />

r 2<br />

gjðz; xÞ ¼X<br />

i<br />

og j<br />

ozi<br />

2<br />

r 2<br />

X<br />

þ zi<br />

l<br />

og j<br />

oxl<br />

2<br />

ðDxlÞ 2 : ð9Þ<br />

The noise factors are assumed to be stochastic, whereas<br />

the control factors are used to optimise the system <strong>and</strong><br />

therefore must remain as parameters in the constraints.<br />

Parkinson et al. [25] proposed a general approach for<br />

<strong>robust</strong> <strong>optimal</strong> <strong>design</strong> <strong>and</strong> addressed two main issues.<br />

The first issue was <strong>design</strong> feasibility, where the procedures<br />

are developed to account for tolerances during <strong>design</strong><br />

optimisation such that the final <strong>design</strong> will remain<br />

feasible despite variations in parameters or variables.<br />

The second issue was the control <strong>of</strong> the transmitted variation<br />

by minimizing sensitivities or by trading <strong>of</strong>f controllable<br />

<strong>and</strong> uncontrolled tolerances. The calculation<br />

<strong>of</strong> the transmitted variation was based on well-known<br />

results developed for the analysis <strong>of</strong> tolerances [26,27].<br />

Parkinson [28] also discussed the feasibility <strong>robust</strong>ness<br />

<strong>and</strong> sensitivity <strong>robust</strong>ness for <strong>robust</strong> mechanical <strong>design</strong><br />

using engineering models. Variability was defined in<br />

terms <strong>of</strong> tolerances. For a worst-case analysis, a tolerance<br />

b<strong>and</strong> was defined, whereas for a statistical analysis<br />

the ± 3r lim<strong>its</strong> for the variable or parameter were used.<br />

A number <strong>of</strong> engineering examples showed that the<br />

method works well if the tolerances are small. For larger<br />

tolerances or for strongly non-linear problems, higher<br />

order Taylor series expansions must be used. A second<br />

order worst-case model was described by Emch <strong>and</strong> Parkinson<br />

[29] <strong>and</strong> a second order model using statistical<br />

analysis was presented by Lewis <strong>and</strong> Parkinson [30].<br />

The disadvantage <strong>of</strong> higher order models is that the formulae<br />

to evaluate the response become quite complicated<br />

<strong>and</strong> computationally expensive.<br />

Sundaresan et al. [31] applied a Sensitivity Index<br />

optimisation approach to determine a <strong>robust</strong> optimum,<br />

but the drawback was the difficulty in choosing the<br />

weighting factors for the mean performance <strong>and</strong><br />

variance. Mulvery et al. [32] treated <strong>robust</strong> <strong>design</strong> as a<br />

bi-objective non-linear programming problem <strong>and</strong> employed<br />

a multi-criteria optimisation approach to generate<br />

a complete <strong>and</strong> efficient solution set to support<br />

decision-making. Chen et al. [33] developed a <strong>robust</strong> <strong>design</strong><br />

methodology to minimize variations caused by the<br />

noise <strong>and</strong> control factors, by setting the factors to zero<br />

in turn.<br />

C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326 319<br />

A <strong>robust</strong> <strong>design</strong> is one that attempts to optimise both<br />

the mean <strong>and</strong> variance <strong>of</strong> the performance, <strong>and</strong> is therefore<br />

a multi-objective <strong>and</strong> non-deterministic problem.<br />

Optimisation <strong>of</strong> the mean <strong>of</strong>ten conflicts with minimizing<br />

the variance, <strong>and</strong> a trade-<strong>of</strong>f decision between them<br />

is needed to choose the best <strong>design</strong>. The conventional<br />

weighted sum (WS) methods to determine this trade<strong>of</strong>f<br />

have serious drawbacks for the Pareto set generation<br />

in multi-objective optimisation problems [34]. The Pareto<br />

set [35] is the set <strong>of</strong> <strong>design</strong>s for which there is no other<br />

<strong>design</strong> that performs better on all objectives. Using<br />

weighted sum methods, it may be impossible to achieve<br />

some Pareto solutions <strong>and</strong> there is no assurance that the<br />

best one is selected, even if all <strong>of</strong> the Pareto points<br />

are available. Chen et al. [36] used a combination <strong>of</strong><br />

multi-objective mathematical programming methods<br />

<strong>and</strong> the principles <strong>of</strong> decision analysis to address the<br />

multi-objective optimisation in <strong>robust</strong> <strong>design</strong>. The compromise<br />

programming (CP) approach, that is the Tchebycheff<br />

or min–max method, replaced the conventional<br />

WS method. The advantages <strong>of</strong> the CP method over<br />

the WS approach in locating the efficient multi-objective<br />

<strong>robust</strong> <strong>design</strong> solution (Pareto points) were illustrated<br />

both theoretically <strong>and</strong> through example problems. Chen<br />

et al. [37] made the bi-objective <strong>robust</strong> <strong>design</strong> optimisation<br />

perspective more powerful by using a physical programming<br />

approach [38–40], where each objective was<br />

controlled with more flexibility than by using CP. Recently,<br />

Messac <strong>and</strong> Ismail-Yahaya [41] formulated the<br />

<strong>robust</strong> <strong>design</strong> optimization problem from a fully multiobjective<br />

perspective, again using the physical programming<br />

method. This method allows the <strong>design</strong>er to<br />

express independent preferences for each <strong>design</strong> metric,<br />

<strong>design</strong> metric variation, <strong>design</strong> variable, <strong>design</strong> variable<br />

variation, <strong>and</strong> parameter variation.<br />

An alternative to CP is the preference aggregation<br />

method, <strong>and</strong> Dai <strong>and</strong> Mourelatos [42] discussed the fundamental<br />

differences <strong>of</strong> these approaches. The compromise<br />

programming approach is a technique for efficient<br />

optimisation to recover an entire Pareto frontier, <strong>and</strong><br />

does not attempt to select one point on that frontier.<br />

Preference aggregation, on the other h<strong>and</strong>, can select<br />

the proper trade-<strong>of</strong>f between nominal performance <strong>and</strong><br />

variability before calculating a Pareto frontier. CP relies<br />

on an efficient algorithm to generate an entire Pareto<br />

frontier <strong>of</strong> <strong>robust</strong> <strong>design</strong>s, while preference aggregation<br />

selects the best trade-<strong>of</strong>f before performing any<br />

calculation.<br />

Mattson <strong>and</strong> Messac [43] gave a brief literature survey<br />

on how various <strong>robust</strong> <strong>design</strong> optimisation methods<br />

h<strong>and</strong>le constraint satisfaction. The focus was the variation<br />

in constraints caused by variations in the controlled<br />

<strong>and</strong> uncontrolled parameters. The constraints are considered<br />

as two types: equality <strong>and</strong> inequality constraints.<br />

Discussions on inequality constraint satisfaction were<br />

given by Balling et al. [44], Parkinson et al. [25], Du


320 C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326<br />

<strong>and</strong> Chen [45], Lee <strong>and</strong> Park [46] <strong>and</strong> Wang <strong>and</strong> Kodialyam<br />

[47]. Research on equality constraint problems are<br />

limited to three approaches; to relax the equality constraint<br />

[48,49,41], to satisfy the equality constraint in a<br />

probabilistic sense [50,51], or to remove the equality<br />

constraint through substitution [52].<br />

5. Stochastic optimization<br />

A slightly different strategy for <strong>robust</strong> <strong>design</strong> optimisation<br />

is based on stochastic optimisation. The stochastic<br />

nature <strong>of</strong> the optimisation arises from incorporating<br />

uncertainty into the procedure, either as the parameter<br />

uncertainty through the noise factors, or because <strong>of</strong><br />

the stochastic nature <strong>of</strong> the optimisation procedure.<br />

The earliest work on stochastic optimization can be<br />

traced back to the 1950s [53] <strong>and</strong> detailed information<br />

may be obtained from recent books [54,55]. The objective<br />

<strong>of</strong> stochastic optimization is to minimize the expectation<br />

<strong>of</strong> the sample performance as a function <strong>of</strong> the <strong>design</strong><br />

parameters <strong>and</strong> the r<strong>and</strong>omness in the system. Eggert<br />

<strong>and</strong> Mayne [56] gave an introduction to probabilistic<br />

optimisation using successive surrogate probability density<br />

functions. Some authors [57–59] employed the concept<br />

exploration method for <strong>robust</strong> <strong>design</strong> optimisation<br />

by creating surrogate global response surface models <strong>of</strong><br />

computationally expensive simulations. The response<br />

surface methodology (RSM) is a set <strong>of</strong> statistical techniques<br />

used to construct an empirical model <strong>of</strong> the relationship<br />

between a response <strong>and</strong> the levels <strong>of</strong> some<br />

input variables, <strong>and</strong> to find the <strong>optimal</strong> responses. Lucas<br />

[60] <strong>and</strong> Myers et al. [61] considered the RSM as an alternative<br />

to TaguchiÕs <strong>robust</strong> <strong>design</strong> method. Monte Carlo<br />

simulation generates instances <strong>of</strong> r<strong>and</strong>om variables<br />

according to their specified distribution types <strong>and</strong> characteristics,<br />

<strong>and</strong> although accurate response statistics<br />

may be obtained, the computation is expensive <strong>and</strong> time<br />

consuming. Mavris <strong>and</strong> B<strong>and</strong>te [62] combined the response<br />

model with a Monte Carlo simulation to construct<br />

cumulative distribution functions (CDFs) <strong>and</strong><br />

probability density functions (PDFs) for the objective<br />

function <strong>and</strong> constraints. All these methods depend on<br />

the sampling <strong>of</strong> the statistics, whereby the probabilistic<br />

distributions <strong>of</strong> the stochastic input sets are required.<br />

One concern is that the response surface approximations<br />

might not generate the accurate sensitivities required for<br />

<strong>robust</strong> <strong>design</strong> [48]. The stochastic finite element method<br />

[63] provides a powerful tool for the analysis <strong>of</strong> structures<br />

with parameter uncertainty. Chakraborty <strong>and</strong><br />

Dey [64] proposed a stochastic finite element method in<br />

the frequency domain for analysis <strong>of</strong> structural dynamic<br />

problems involving uncertain parameters. The uncertain<br />

structural parameters are modelled as homogeneous<br />

Gaussian stochastic fields <strong>and</strong> discretised by the local<br />

averaging method. Numerical examples were presented<br />

to demonstrate the accuracy <strong>and</strong> efficiency <strong>of</strong> the proposed<br />

method. Chen [65] used Monte Carlo simulation<br />

<strong>and</strong> quadratic programming. Schueller [66] gave a recent<br />

<strong>review</strong> on structural stochastic analysis.<br />

Optimisation approaches that are inherently stochastic<br />

include techniques such as simulated annealing, neural<br />

networks <strong>and</strong> evolutionary algorithms (EA) (genetic<br />

algorithms, evolutionary programming <strong>and</strong> evolution<br />

strategies (ES)), <strong>and</strong> these have been applied to multiobjective<br />

optimisation problems [67–69]. These techniques<br />

do not require the computation <strong>of</strong> gradients,<br />

which is important if the objective function relies on estimating<br />

moments <strong>of</strong> the response r<strong>and</strong>om variables.<br />

Gupta <strong>and</strong> Li [70] applied mathematical programming<br />

<strong>and</strong> neural networks to <strong>robust</strong> <strong>design</strong> optimisation, <strong>and</strong><br />

the numerical examples showed that the approach is able<br />

to solve highly non-linear <strong>design</strong> optimisation problems<br />

in mechanical <strong>and</strong> structural <strong>design</strong>. S<strong>and</strong>gren <strong>and</strong> Cameron<br />

[71] used a hybrid combination <strong>of</strong> a genetic algorithm<br />

<strong>and</strong> non-linear programming for <strong>robust</strong> <strong>design</strong><br />

optimization <strong>of</strong> structures with variations in loading,<br />

geometry <strong>and</strong> material properties. Parkinson [72] employed<br />

a genetic algorithm for <strong>robust</strong> <strong>design</strong> to directly<br />

obtain a global minimum for the variability <strong>of</strong> a <strong>design</strong><br />

function by varying the nominal <strong>design</strong> parameter values.<br />

The method proved effective <strong>and</strong> more efficient than<br />

conventional optimisation algorithms. Additional studies<br />

are required before these methods are suitable for<br />

application to large-scale optimisation problems.<br />

6. Applications in dynamics<br />

The most successful applications <strong>of</strong> <strong>robust</strong> <strong>design</strong> are<br />

found in the fields <strong>of</strong> mechanical <strong>design</strong> engineering (static<br />

performance) <strong>and</strong> process systems, <strong>and</strong> there have<br />

been few applications to the <strong>robust</strong>ness <strong>of</strong> dynamic performance.<br />

Seki <strong>and</strong> Ishii [73] applied the <strong>robust</strong> <strong>design</strong><br />

concept to the dynamic <strong>design</strong> <strong>of</strong> an optical pick-up<br />

actuator focusing on shape synthesis using computer<br />

models <strong>and</strong> <strong>design</strong> <strong>of</strong> experiments. The response in the<br />

first bending <strong>and</strong> torsion modes were selected as measures<br />

<strong>of</strong> undesirable vibration energy. The objective<br />

functions were defined as the signal-to-noise ratios <strong>of</strong> response<br />

frequencies <strong>and</strong> the sensitivities were derived<br />

from the <strong>design</strong> <strong>of</strong> experiments using an orthogonal array.<br />

Hwang et al. [74] optimised the vibration displacements<br />

<strong>of</strong> an automobile rear view mirror system for<br />

<strong>robust</strong>ness, defined by the Taguchi concept.<br />

In this paper the <strong>robust</strong> <strong>design</strong> approach is applied to<br />

the dynamics <strong>of</strong> a tuned vibration absorber due to<br />

parameter uncertainty, using the optimisation approach<br />

through non-linear programming. The objective is to<br />

determine the stiffness, mass <strong>and</strong> damping parameters<br />

<strong>of</strong> the absorber, to minimize the displacements <strong>of</strong> the<br />

main system over a large range <strong>of</strong> excitation frequencies,


despite uncertainty in the mass <strong>and</strong> stiffness properties<br />

<strong>of</strong> the main system.<br />

The principle <strong>of</strong> the vibration absorber was attributed<br />

to Frahm [75] who found that a natural frequency<br />

<strong>of</strong> a structure could be split into two frequencies by<br />

attaching a small spring–mass system tuned to the same<br />

frequency as the structure. Den Hartog <strong>and</strong> Ormonroyd<br />

[76] developed a theoretical analysis <strong>of</strong> the vibration absorber<br />

<strong>and</strong> showed that a damped vibration absorber<br />

could control the vibration over a wide frequency range.<br />

Brock [77] <strong>and</strong> Den Hartog [78] gave criteria for the efficient<br />

optimum operation <strong>of</strong> a tuned vibration absorber,<br />

such as the relationship <strong>of</strong> the frequency <strong>and</strong> mass ratios<br />

between the absorber <strong>and</strong> main system <strong>and</strong> the relationship<br />

between the damping <strong>and</strong> mass ratio. Jones et al.<br />

[79] successfully <strong>design</strong>ed prototype vibration absorbers<br />

for two bridges using these criteria. However, the effect<br />

<strong>of</strong> parameter uncertainty <strong>and</strong> variations in the dynamic<br />

performance have not been considered.<br />

We consider the forced vibration <strong>of</strong> the two degree <strong>of</strong><br />

freedom system shown in Fig. 2. The original one degree<br />

<strong>of</strong> freedom system, referred to as the main system, consists<br />

<strong>of</strong> the mass m1 <strong>and</strong> the spring k1, <strong>and</strong> the added the<br />

system, referred to as the absorber, consists <strong>of</strong> the mass<br />

m2, the spring k2 <strong>and</strong> the damper c2. The equations <strong>of</strong> motion are<br />

m1€q 1 þ c2ð_q 1 _q 2Þþk1q1 þ k2ðq1 q2Þ¼f0 sinðXtÞ<br />

m2€q 2 þ c2ð_q 2 _q 1Þþk2ðq2 q1Þ¼0 ð10Þ<br />

Inman [80] <strong>and</strong> Smith [81] should be consulted for further<br />

details <strong>of</strong> the modelling <strong>and</strong> analysis <strong>of</strong> vibration<br />

absorbers. Solving these equations for the steady-state<br />

solution, gives the amplitude <strong>of</strong> vibration <strong>of</strong> the two<br />

masses as<br />

q 1 max ¼ f0<br />

q 2 max ¼ f0<br />

The equations may be non-dimensionalised using a<br />

ÔstaticÕ deflection <strong>of</strong> main system, defined by<br />

q1st ¼ f0<br />

k1<br />

c 2 2 X2 þðk2 m2X 2 Þ 2<br />

c 2 2 X2 ðk1 m1X 2 m2X 2 Þ 2 þðk2m2X 2<br />

c 2 2 X2 þ k 2<br />

2<br />

c 2 2 X2 ðk1 m1X 2 m2X 2 Þ 2 þðk2m2X 2<br />

ð12Þ<br />

to give the non-dimensional displacement <strong>of</strong> mass m1<br />

as<br />

C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326 321<br />

F ðX;m1;k1Þ¼ q 1max<br />

q 1st<br />

¼ k1 c 2<br />

2 X2 þðk2 m2X 2 Þ 2<br />

þ k2m2X 2<br />

k 1<br />

.<br />

ðk1 m1X 2 Þðk2 m2X 2 Þ 2 1=2<br />

c 2<br />

2 X2 ðk1 m1X 2 m2X 2 Þ 2<br />

ð13Þ<br />

Suppose a steel box girder footbridge may be represented<br />

by a single degree <strong>of</strong> freedom system with mass<br />

m1 = 17500 kg <strong>and</strong> stiffness k1 = 3.0 MN/m. The worst<br />

case <strong>of</strong> dynamic loading is considered to be equivalent<br />

to a sinusoidal loading with a constant amplitude <strong>of</strong><br />

0.48 kN at the natural frequency <strong>of</strong> the bridge. To simulate<br />

the environmental changes, the stiffness k1 <strong>and</strong> the<br />

mass m1 are allowed to undergo 10% variations<br />

(k1 ± Dk1, m1 ± Dm<br />

qffiffiffiffi<br />

1), <strong>and</strong> a wide excitation frequency<br />

k1<br />

b<strong>and</strong> ð1 6 X 6 3 Þ is considered. This ensures that<br />

m1 the absorber works well for a wide range <strong>of</strong> possible<br />

excitations. The frequency X is the signal factor, <strong>and</strong><br />

the range <strong>of</strong> this frequency gives the set V. Using the<br />

worst-case formulation, the st<strong>and</strong>ard deviations <strong>of</strong> the<br />

ðk1 m1X 2 Þðk2 m2X 2 ÞÞ 2<br />

ðk1 m1X 2 Þðk2 m2X 2 ÞÞ 2<br />

! 1=2<br />

! 1=2<br />

m 1<br />

mass <strong>and</strong> stiffness in the main system are rk1<br />

q 1<br />

f sin(Ωt)<br />

0<br />

Fig. 2. Forced vibration <strong>of</strong> a two degree <strong>of</strong> freedom system.<br />

ð11Þ<br />

¼ 1<br />

3 Dk1<br />

<strong>and</strong> rm 1 ¼ 1<br />

3 Dm1. This st<strong>and</strong>ard deviation is calculated<br />

for a uniform distribution over the 10% parameter variations<br />

<strong>and</strong> is used for the <strong>robust</strong> <strong>design</strong> optimisation.<br />

However the original intervals are used in the simulations<br />

to demonstrate the effectiveness <strong>of</strong> the solutions.<br />

The mean response function <strong>and</strong> the st<strong>and</strong>ard deviation<br />

k 2<br />

c 2<br />

m 2<br />

q 2


322 C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326<br />

<strong>of</strong> the maximum non-dimensional displacement <strong>of</strong> mass<br />

m1 may be calculated using the first-order Taylor<br />

approximation as<br />

2<br />

3<br />

6<br />

lf ¼E6 4<br />

7<br />

max qffiffiffiðF ðX;m1;k1ÞÞ7<br />

5<br />

16X63<br />

k 1<br />

m 1<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

rf ¼E max qffiffiffi<br />

oF ðX;m1;k1Þ<br />

ok1<br />

2<br />

r2 k þ 1<br />

oF ðX;m1;k1Þ<br />

om1<br />

2<br />

r2 2<br />

3<br />

0s<br />

1<br />

6<br />

4<br />

@<br />

7<br />

A7<br />

m1 5<br />

16X63<br />

k 1<br />

m 1<br />

ð14Þ<br />

Here the expected value is evaluated for the uncertain<br />

mass <strong>and</strong> stiffness parameters m1 <strong>and</strong> k1, <strong>and</strong> F is defined<br />

in Eq. (13). The <strong>design</strong> variables for the absorber are m 2,<br />

k2 <strong>and</strong> c2, with lower <strong>and</strong> upper bounds given by<br />

m2 2 [10, 1750], c2 2 [10, 2000], k2 2 [100, 10 6 ]. Therefore,<br />

the <strong>robust</strong> <strong>design</strong> <strong>of</strong> a suitable tuned vibration absorber<br />

may be formulated as follows:<br />

Minimize: ½lfðm2; c2; k2Þ; rfðm2; c2; k2ÞŠ<br />

Subject to: j Dk1 j6 0 1k1; j Dm1 j6 0:1m1<br />

10 6 m2 6 1750; 10 6 c2 6 2000;<br />

100 6 k2 6 10 6<br />

ð15Þ<br />

The first step <strong>of</strong> <strong>robust</strong> <strong>design</strong> is to seek the ideal <strong>design</strong><br />

(Utopia point) by a minimisation <strong>of</strong> lf <strong>and</strong> rf individually<br />

as single objective functions. The ideal <strong>design</strong> obtained<br />

is denoted by ½l f ; r f Š. Since there are only two<br />

objective functions, lf <strong>and</strong> rf, the two functions are combined<br />

into a single objective function, G, by the conventional<br />

weighted sum method discussed earlier. The<br />

objective function is then<br />

G ¼ a lf þð1 aÞ<br />

lf rf<br />

ð16Þ<br />

rf where the weighting factor a 2 [0,1] represents the relative<br />

importance <strong>of</strong> the two objectives. G is minimised<br />

for a between 0 <strong>and</strong> 1 in steps <strong>of</strong> 0.1, <strong>and</strong> Fig. 3 shows<br />

the results in the objective space, formed by the normalized<br />

values <strong>of</strong> lf versus rf. The trade-<strong>of</strong>f between the<br />

mean <strong>and</strong> the st<strong>and</strong>ard deviation can clearly be observed.<br />

Points 1 <strong>and</strong> 11 denote the two utopia points,<br />

showing the minimized mean (a = 1) <strong>and</strong> variance<br />

(a = 0) response optimization, respectively. The other<br />

points are the optimized results from different weighting<br />

<strong>of</strong> the mean <strong>and</strong> variance objection functions. This<br />

weighted sum approach has similarities to the L-curve<br />

method used in regularization [82]. The 11 optimised<br />

solutions for the absorber parameters (mass, stiffness<br />

<strong>and</strong> damping) <strong>and</strong> the recommended solution from<br />

vibration textbooks [80,81], are denoted RD1–RD11<br />

<strong>and</strong> BS, <strong>and</strong> are listed in Table 1. To visualise all <strong>of</strong> these<br />

responses over the excitation frequency b<strong>and</strong> <strong>of</strong> interest,<br />

the non-dimensional displacements <strong>of</strong> mass m1 for these<br />

different parameter sets are plotted in Fig. 4, where TD<br />

denotes the text book solution.<br />

*<br />

σ / σ<br />

f f<br />

2<br />

1.5<br />

1<br />

1<br />

2<br />

3<br />

4<br />

5 6<br />

7<br />

8 9 10 11<br />

1 1.5 2 2.5<br />

*<br />

µ / µ<br />

f f<br />

Fig. 3. Efficient solutions <strong>of</strong> a <strong>robust</strong> multi-objective<br />

optimization.<br />

Table 1<br />

The <strong>robust</strong> optimization solutions for the absorber <strong>design</strong>,<br />

together with the textbook solution<br />

Name a m2 (kg) k2 (N/m) c2 (N s/m) m2m1<br />

RD1 1 203.72 34509 2000 0.011641<br />

RD2 0.9 204.05 34576 2000 0.01166<br />

RD3 0.8 209.79 35532 2000 0.011988<br />

RD4 0.7 239.12 40421 2000 0.013664<br />

RD5 0.6 251.27 42421 2000 0.014358<br />

RD6 0.5 259.71 43817 2000 0.01484<br />

RD7 0.4 310.37 52188 2000 0.017736<br />

RD8 0.3 348.83 58483 2000 0.019933<br />

RD9 0.2 410.44 68483 1999.1 0.023454<br />

RD10 0.1 421.27 70238 2000 0.024072<br />

RD11 0 510.93 84592 2000 0.029196<br />

BS (book<br />

solution)<br />

175 29393 272.2 0.01<br />

q 1max / q 1st<br />

10<br />

8<br />

6<br />

4<br />

2<br />

15 0<br />

10<br />

Design Number<br />

5<br />

0<br />

0<br />

10<br />

20<br />

Ω (rad/s)<br />

RD1<br />

RD2<br />

RD3<br />

RD4<br />

RD5<br />

RD6<br />

RD7<br />

RD8<br />

RD9<br />

RD10<br />

RD11<br />

TD<br />

Fig. 4. Robust <strong>design</strong> results compared with the traditional<br />

solution.<br />

30<br />

40


Most <strong>of</strong> the damping values for the <strong>robust</strong> <strong>design</strong><br />

cases (except for the RD9 case) are selected as the upper<br />

limit (2000 Ns/m). Larger damping values are able to reduce<br />

the amplitude <strong>of</strong> the peak response very effectively,<br />

although in practice the amount <strong>of</strong> damping that can be<br />

added is limited. The optimum mass ratio (m2/m1) is between<br />

1% <strong>and</strong> 3%, <strong>and</strong> the value <strong>of</strong> the mass for the RD1<br />

case, where only the mean response was minimized, is<br />

close to that <strong>of</strong> the BS case. With the decreased weighting<br />

on the mean response <strong>and</strong> correspondingly increasing<br />

weighting on the response variance, the absorber<br />

mass is gradually increased, <strong>and</strong> the response shown in<br />

Fig. 4 becomes flatter. Although the peak response is decreased<br />

in these cases, the response near the main system<br />

natural frequency increases.<br />

To demonstrate the effectiveness <strong>of</strong> the <strong>robust</strong> <strong>design</strong><br />

approach, Monte Carlo simulation [83] is used to evaluate<br />

the possible response variations <strong>of</strong> the main system<br />

due to the mass <strong>and</strong> stiffness uncertainty in the main system,<br />

at the different optimum points. In Monte Carlo<br />

simulation a r<strong>and</strong>om number generator produces samples<br />

<strong>of</strong> the noise factors (in this case m 1 <strong>and</strong> k 1), which<br />

are then used to calculate the response over the frequency<br />

range <strong>of</strong> interest for a given set <strong>of</strong> control factors.<br />

The r<strong>and</strong>om number generator approximates the<br />

PDF <strong>of</strong> the noise factors for a large number <strong>of</strong> samples.<br />

The response variations then give some indication <strong>of</strong> the<br />

<strong>robust</strong>ness <strong>of</strong> the <strong>design</strong> given by those particular set <strong>of</strong><br />

control factors. Four representative cases, namely RD1<br />

(a = 1), RD6 (a = 0.5), RD11 (a = 0) <strong>and</strong> BS, are investigated,<br />

<strong>and</strong> frequency response variations are plotted in<br />

Figs. 5–8, together with the corresponding mean response,<br />

shown by the solid line. The amplitudes <strong>of</strong> the<br />

response for the three <strong>robust</strong> <strong>design</strong> cases (RD1, RD6,<br />

RD11) are much smaller than those from the traditional<br />

textbook <strong>design</strong> (BS). The RD1 case has the lowest minimized<br />

mean response <strong>and</strong> the most sensitive variations<br />

q 1max / q 1st<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 10 20<br />

Ω (rad/s)<br />

30 40<br />

Fig. 5. Monte Carlo simulation <strong>of</strong> the response variation due to<br />

the uncertainty (BS).<br />

C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326 323<br />

q 1max / q 1st<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 10 20 30 40<br />

Ω (rad/s)<br />

Fig. 6. Monte Carlo simulation <strong>of</strong> the response variation due to<br />

the uncertainty (RD1).<br />

q 1max / q 1st<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 10 20<br />

Ω (rad/s)<br />

30 40<br />

Fig. 7. Monte Carlo simulation <strong>of</strong> the response variation due to<br />

the uncertainty (RD11).<br />

q 1max / q 1st<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 10 20 30 40<br />

Ω (rad/s)<br />

Fig. 8. Monte Carlo simulation <strong>of</strong> the response variation due to<br />

the uncertainty (RD6).


324 C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326<br />

while the RD11 case has the highest mean response <strong>and</strong><br />

the least sensitive variations. The RD6 case is a compromise<br />

between the RD1 <strong>and</strong> RD11 cases, with a reasonable<br />

mean response <strong>and</strong> insensitive variations due to<br />

the uncertainty. The textbook solution, shown in Fig.<br />

5, is most sensitive to the mass <strong>and</strong> stiffness variations<br />

in the main system, although increasing the damping<br />

in the absorber would improve this performance.<br />

7. Concluding remarks<br />

The state-<strong>of</strong>-art approaches to <strong>robust</strong> <strong>design</strong> optimisation<br />

have been extensively <strong>review</strong>ed. It is noted that<br />

<strong>robust</strong> <strong>design</strong> is a multi-objective <strong>and</strong> non-deterministic<br />

problem. The objective is to optimise the mean <strong>and</strong> minimize<br />

the variability in the performance response that results<br />

from uncertainty represented through noise<br />

variables. The <strong>robust</strong> <strong>design</strong> approaches can generally<br />

be classified as statistical-based methods <strong>and</strong> optimisation<br />

methods. Most <strong>of</strong> the Taguchi based methods use<br />

direct experimentation <strong>and</strong> the objective functions for<br />

the optimisation are expressed as the signal to noise<br />

ratio (SNR) using the Taguchi method. Using the<br />

orthogonal array technique, the analysis <strong>of</strong> variance<br />

<strong>and</strong> analysis <strong>of</strong> mean <strong>of</strong> the SNR are used to evaluate<br />

the optimum <strong>design</strong> variables to ensure that the system<br />

performance is insensitive to the effects <strong>of</strong> noise, <strong>and</strong><br />

to tune the mean response to the target. The optimisation<br />

approaches for <strong>robust</strong> <strong>design</strong> are based on non-linear<br />

programming methods. The objective functions<br />

simultaneously optimise both the mean performance<br />

<strong>and</strong> the variance in performance. A trade-<strong>of</strong>f decision<br />

must be made, to choose the best <strong>design</strong> with the maximum<br />

<strong>robust</strong>ness. Recently, novel techniques such as<br />

simulated annealing, neural networks <strong>and</strong> the field <strong>of</strong><br />

evolutionary algorithms have been applied to solving<br />

the resulting multi-objective optimisation problem.<br />

The application <strong>of</strong> <strong>robust</strong> <strong>design</strong> optimisation in<br />

structural dynamics is very rare. This paper considers<br />

the forced vibration <strong>of</strong> a two degree <strong>of</strong> freedom system<br />

as an example to illustrate the <strong>robust</strong> <strong>design</strong> <strong>of</strong> a vibration<br />

absorber. The objective is to minimize the displacement<br />

response <strong>of</strong> the main system within a wide b<strong>and</strong><br />

<strong>of</strong> excitation frequencies. The <strong>robust</strong>ness <strong>of</strong> the response<br />

due to uncertainty in the mass <strong>and</strong> stiffness <strong>of</strong><br />

the main system was also considered, <strong>and</strong> the maximum<br />

mean displacement response <strong>and</strong> the variations<br />

caused by the mass <strong>and</strong> stiffness uncertainty were minimized<br />

simultaneously. Monte Carlo simulation demonstrated<br />

significant improvement in the mean response<br />

<strong>and</strong> variation compared with the traditional solution<br />

recommended from vibration textbooks. The results<br />

show that <strong>robust</strong> <strong>design</strong> methods have great potential<br />

for application in structural dynamics to deal with<br />

uncertain structures.<br />

Acknowledgments<br />

The authors acknowledge the support <strong>of</strong> the EPSRC<br />

(UK) through grants GR/R34936 <strong>and</strong> GR/R26818.<br />

Pr<strong>of</strong>. <strong>Friswell</strong> acknowledges the support <strong>of</strong> a Royal<br />

Society-Wolfson Research Merit Award.<br />

References<br />

[1] Harry MJ. The nature <strong>of</strong> Six Sigma Quality. Shaumburg,<br />

IL: Motorola University Press; 1997.<br />

[2] P<strong>and</strong>e PS, Neuman RP, Cavanagh RR. The Six Sigma<br />

way: how GE, Motorola, <strong>and</strong> other top companies are<br />

honing their performance. New York: McGraw-Hill;<br />

2000.<br />

[3] Siddall JN. A new approach to probability in engineering<br />

<strong>design</strong> <strong>and</strong> optimisation. ASME J Mech, Transmiss,<br />

Autom Des 1984;106:5–10.<br />

[4] Doltsinis I, Kang Z. Robust <strong>design</strong> <strong>of</strong> structures using<br />

optimization methods. Comput Methods Appl Mech Eng<br />

2004;193:2221–37.<br />

[5] Fisher RA. Design <strong>of</strong> experiments. Edinburgh: Oliver &<br />

Boyd; 1951.<br />

[6] Phadke MS. Quality engineering using <strong>robust</strong><br />

<strong>design</strong>. Englewood Clifffs, NJ: Prentice-Hall; 1989.<br />

[7] Taguchi G. Quality engineering through <strong>design</strong> optimization.<br />

White Plains, NY: Krauss International Publications;<br />

1986.<br />

[8] Taguchi G. System <strong>of</strong> experimental <strong>design</strong>. IEEE J<br />

1987;33:1106–13.<br />

[9] Fowlkes WY, Creveling CM. Engineering methods for<br />

<strong>robust</strong> product <strong>design</strong>: using Taguchi methods in technology<br />

<strong>and</strong> product development. Reading, MA: Addison-<br />

Wesley Publishing Company; 1995.<br />

[10] Ragsdell KM, dÕEntremont KL. Design for latitude using<br />

TOPT. ASME Adv Des Autom 1988;DE-14:265–72.<br />

[11] Box G, Fung C. Studies in quality improvement: minimizing<br />

transmitted variation by parameter <strong>design</strong>. In: Report<br />

No. 8, Center for Quality Improvement, University <strong>of</strong><br />

Wisconsin; 1986.<br />

[12] Montgomery DC. Design <strong>and</strong> analysis <strong>of</strong> experiments. 3rd<br />

ed. New York: John Wiley & Sons; 1991.<br />

[13] Tsui KL. An overview <strong>of</strong> Taguchi method <strong>and</strong> newly<br />

developed statistical methods for <strong>robust</strong> <strong>design</strong>. IIE Trans<br />

1992;24:44–57.<br />

[14] Nair VN. TaguchiÕs parameter <strong>design</strong>: a panel discussion.<br />

Technometrics 1992;34:127–61.<br />

[15] DÕErrico JR, Zaino Jr NA. Statistical tolerancing using a<br />

modification <strong>of</strong> TaguchiÕs method. Technometrics<br />

1988;30(4):397–405.<br />

[16] Yu J, Ishii K. A <strong>robust</strong> optimization procedure for systems<br />

with significant non-linear effects. In: ASME Design<br />

Automation Conference, Albuquerque, NM 1993; DE-<br />

65-1:371–8.<br />

[17] Otto JN, Antonsson EK. Extensions to the Taguchi<br />

method <strong>of</strong> product <strong>design</strong>. ASME J Mech Des<br />

1993;115:5–13.<br />

[18] Ramakrishnan B, Rao SS. A <strong>robust</strong> optimization approach<br />

using TaguchiÕs loss function for solving nonlinear opti-


mization problems. ASME Adv Des Autom 1991;DE-<br />

32(1):241–8.<br />

[19] Ramakrishnan B, Rao SS. Efficient strategies for the<br />

<strong>robust</strong> optimization <strong>of</strong> large-scale nonlinear <strong>design</strong> problems.<br />

ASME Adv Des Autom 1994;DE-69(2):25–35.<br />

[20] Moh<strong>and</strong>as SU, S<strong>and</strong>gren E. Multi-objective optimization<br />

dealing with uncertainty. In: Proceedings <strong>of</strong> the 15th<br />

ASME Design Automation Conference Montreal, vol. DE<br />

19-2; 1989.<br />

[21] S<strong>and</strong>gren E. A multi-objective <strong>design</strong> tree approach for<br />

optimization under uncertainty. In: Proceedings <strong>of</strong> the 15th<br />

ASME Design Automation Conference Montreal, vol. 2;<br />

1989.<br />

[22] Belegundu AD, Zhang S. Robustness <strong>of</strong> <strong>design</strong> through<br />

minimum sensitivity. J Mech Des 1992;114:213–7.<br />

[23] Chang T, Ward AC, Lee J, Jacox EH. Distributed <strong>design</strong><br />

with conceptual <strong>robust</strong>ness: a procedure based on TaguchiÕs<br />

parameter <strong>design</strong>. Concurrent Product Des ASME<br />

1994;DE 74:19–29.<br />

[24] Lee KH, Eom IS, Park GJ, Lee WI. Robust <strong>design</strong> for<br />

unconstrained optimization problems using the Taguchi<br />

method. AIAA J 1996;34(5):1059–63.<br />

[25] Parkinson A, Sorensen C, Pourhassan N. A general<br />

approach for <strong>robust</strong> <strong>optimal</strong> <strong>design</strong>. J Mech Des ASME<br />

Trans 1993;115:74–80.<br />

[26] Cox NE. In: Cornell J, Shapiro S. editors. Volume II: how<br />

to perform statistical tolerance analysis, the ASQC basic<br />

reference in quality control: statistical techniques; 1986.<br />

[27] Bjorke O. Computer-aided tolerancing. New York:<br />

ASME Press; 1989.<br />

[28] Parkinson A. Robust mechanical <strong>design</strong> using engineering<br />

models. J Mech Des ASME Trans 1995;117:48–54.<br />

[29] Emch G, Parkinson A. Robust <strong>optimal</strong> <strong>design</strong> for worstcase<br />

tolerances. ASME J Mech Des 1994;116:1019–25.<br />

[30] Lewis L, Parkinson A. Robust <strong>optimal</strong> <strong>design</strong> with a<br />

second-order tolerance model. Res Eng Des 1994;6:25–37.<br />

[31] Sunderasan S, Ishii K, Houser DR. A <strong>robust</strong> optimization<br />

procedure with variation on <strong>design</strong> variables <strong>and</strong> constraints.<br />

ASME Adv Des Autom 1993;DE 65(1):379–86.<br />

[32] Mulvey J, V<strong>and</strong>erber R, Zenios S. Robust optimization <strong>of</strong><br />

large-scale systems. Operat Res 1995;43:264–81.<br />

[33] Chen W, Allen JK, Mistree F, Tsui K-L. A procedure for<br />

<strong>robust</strong> <strong>design</strong>: minimizing variations caused by noise factors<br />

<strong>and</strong> control factors. ASME J Mech Des 1996;118:478–85.<br />

[34] Das I, Dennis JE. A closer look at drawbacks <strong>of</strong><br />

minimizing weighted sums <strong>of</strong> objectives for Pareto set<br />

generation in multi-criteria optimization problems. Struct<br />

Optim 1997;14:63–9.<br />

[35] Steuer RE. Multiple criteria optimisation: theory, computation<br />

<strong>and</strong> application. New York: John Wiley; 1986.<br />

[36] Chen W, Wiecek MM, Zhang J. Quality utility: a<br />

compromise programming approach to <strong>robust</strong> <strong>design</strong>.<br />

ASME J Mech Des 1999;121(2):179–87.<br />

[37] Chen W, Sahai A, Messac A, Sundararaj GJ. Exploration<br />

<strong>of</strong> the effectiveness <strong>of</strong> physical programming in <strong>robust</strong><br />

<strong>design</strong>. J Mech Des 2000;122:155–63.<br />

[38] Messac A. Physical programming: effective optimization<br />

for computational <strong>design</strong>. AIAA J 1996;34:149–58.<br />

[39] Messac A. Control-structure integrated <strong>design</strong> with closedform<br />

<strong>design</strong> metrics using physical programming. AIAA J<br />

1998;36:855–64.<br />

C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326 325<br />

[40] Messac A. From the dubious construction <strong>of</strong> objective<br />

functions to the application <strong>of</strong> physical programming.<br />

AIAA J 2000;38:155–63.<br />

[41] Messac A, Ismail-Yahaya A. Multi-objective <strong>robust</strong> <strong>design</strong><br />

using physical programming. Struct Multidisciplin Optimiz<br />

2002;23(5):357–71.<br />

[42] Dai Z, Mourelatos ZP. Robust <strong>design</strong> using preference<br />

aggregation methods. In: ASME 2003 Design Engineering<br />

Technical Conferences <strong>and</strong> Computer <strong>and</strong> Information in<br />

Engineering Conference, Chicago, IL, USA, DETC2003/<br />

DAC-48715.<br />

[43] Mattson C, Messac A. H<strong>and</strong>ing equality constraints in<br />

<strong>robust</strong> <strong>design</strong> optimization. In: 44th AIAA/ASME/ASCE/<br />

AHS Structures, Structure Dynamics, <strong>and</strong> Materials Conference.<br />

2003, Paper No. AIAA-2003-1780.<br />

[44] Balling RJ, Free JC, Parkinson AR. Consideration <strong>of</strong><br />

worst-case manufacturing tolerances in <strong>design</strong> optimization.<br />

ASME J Mech, Trans, Autom Des 1986;108:438–41.<br />

[45] Du X, Chen W. Towards a better underst<strong>and</strong>ing <strong>of</strong><br />

modeling feasibility <strong>robust</strong>ness in engineering <strong>design</strong>.<br />

ASME J Mech Des 2001;122:385–94.<br />

[46] Lee KH, Park GJ. Robust optimization considering<br />

tolerances <strong>of</strong> <strong>design</strong> variables. Comput Struct 2001;79:<br />

77–86.<br />

[47] Wang L, Kodialyam S. An efficient method for probabilistic<br />

<strong>and</strong> <strong>robust</strong> Design with non-normal distributions. In:<br />

AIAA 43rd AIAA/ASME/ASCE/AHS Structures, Structure<br />

Dynamics, <strong>and</strong> Materials Conference. 2002, Paper No.<br />

AIAA-2002-1754.<br />

[48] Su J, Renaud JE. Automatic differentiation in <strong>robust</strong><br />

optimization. AIAA J 1997;35:1072–9.<br />

[49] Fares B, Noll D, Apkarian P. Robust control via sequential<br />

semidefinite programming. SIAM J Contr Optimiz<br />

2002;23(5):357–71.<br />

[50] Sunderasan S, Ishii K, Houser DR. Design optimization<br />

for <strong>robust</strong>ness using performance simulation programs.<br />

ASME Adv Des Autom 1991;DE 32(1):249–56.<br />

[51] Putko MM, Newman PA, Taylor AC, Green LL.<br />

Approach for uncertainty propagation <strong>and</strong> <strong>robust</strong> <strong>design</strong><br />

in CFD using sensitivity derivatives. In: AIAA 42nd<br />

AIAA/ASME/ASCE/AHS Structures, Structural Dynamics,<br />

<strong>and</strong> Materials Conference. 2001. Paper No. AIAA-<br />

2001-2528.<br />

[52] Das I. Robust optimization for constrained, nonlinear<br />

programming problems. Eng Optim 2000;32(5):585–<br />

618.<br />

[53] Beale EML. On minimizing a convex function subject<br />

to linear inequalities. J R Statist Soc 1955;17B:173–<br />

84.<br />

[54] Birge JR, Louveaux FV. Introduction to stochastic programming.<br />

New York, NY: Springer; 1997.<br />

[55] Kall P, Wallace SW. Stochastic programming. New York,<br />

NY: Wiley; 1994.<br />

[56] Eggert RJ, Mayne RW. Probabilistic <strong>optimal</strong>-<strong>design</strong> using<br />

successive surrogate probability density-functions. ASME<br />

J Mech Des 1993;115(3):385–91.<br />

[57] Chen W, Allen JK, Mavis D, Mistree F. A concept<br />

exploration method for determining <strong>robust</strong> top-level specifications.<br />

Eng Optimiz 1996;26:137–58.<br />

[58] Koch PN, Barlow A, Allen JK, Mistree F. Configuring turbine<br />

propulsion systems using <strong>robust</strong> concept exploration.


326 C. Zang et al. / Computers <strong>and</strong> Structures 83 (2005) 315–326<br />

In: ASME Design Automation Conference. Irvine, CA, 96-<br />

DETC/DAC-1285; 1996.<br />

[59] Simpson TW, Chen W, Allen JK, Mistree F. Conceptual<br />

<strong>design</strong> <strong>of</strong> a family <strong>of</strong> products through the use <strong>of</strong> the <strong>robust</strong><br />

concept exploration method. In AIAA/NASA/USAF/<br />

ISSMO Symposium on Multidisciplinary Analysis <strong>and</strong><br />

Optimisation, Bellevue, Washington, vol. 2(2); 1996. p.<br />

1535–45.<br />

[60] Lucas JM. How to achieve a <strong>robust</strong> process using response<br />

surface methodology. J Quality Technol 1994;26:248–60.<br />

[61] Myers RH, Khuri AI, Vining G. Response surface alternative<br />

to the Taguchi <strong>robust</strong> parameter <strong>design</strong> approach. J<br />

Am Stat 1992;46:131–9.<br />

[62] Mavris DN, B<strong>and</strong>te O. A probabilistic approach to<br />

multivariate constrained <strong>robust</strong> <strong>design</strong> simulation. 1997.<br />

Paper No. AIAA-97-5508.<br />

[63] Elishak<strong>of</strong>f I, Ren YJ, Shinozuka M. Improved finite<br />

element method for stochastic structures. Chaos, Solitons<br />

& Fractals 1995;5:833–46.<br />

[64] Chakraborty CS, Dey SS. A stochastic finite element<br />

dynamic analysis <strong>of</strong> structures with uncertain parameters.<br />

Int J Mech Sci 1998;40:1071–87.<br />

[65] Chen S-P. Robust <strong>design</strong> with dynamic characteristics<br />

using stochastic sequential quadratic programming. Eng<br />

Optimiz 2003;35(1):79:89.<br />

[66] Schueller GI. Computational stochastic mechanics recent<br />

advances. Comput Struct 2001;79:2225–34.<br />

[67] Hajela P. Nongradient methods in multidisciplinary <strong>design</strong><br />

optimisationstatus <strong>and</strong> potential. J Aircraft 1999;36(1):<br />

255–65.<br />

[68] Kalyanmoy DC. Multi-objective optimization using evolutionary<br />

algorithms. West Sussex, Engl<strong>and</strong>: John Wiley<br />

& Sons; 2001.<br />

[69] Fouskakis D, Draper D. Stochastic optimisation: a <strong>review</strong>.<br />

Int Statist Rev 2002;70(3):315–49.<br />

[70] Gupta KC, Li J. Robust <strong>design</strong> optimization with mathematical<br />

programming neural networks. Comput Struct<br />

2000;76:507–16.<br />

[71] S<strong>and</strong>gren E, Cameron TM. Robust <strong>design</strong> optimization <strong>of</strong><br />

structures through consideration <strong>of</strong> variation. Comput<br />

Struct 2002;80:1605–13.<br />

[72] Parkinson DB. Robust <strong>design</strong> employing a genetic algorithm.<br />

Qual Reliab Eng Int 2000;16:201–8.<br />

[73] Seki K, Ishii K. Robust <strong>design</strong> for dynamic performance:<br />

optical pick-up example. In: Proceedings <strong>of</strong> the 1997<br />

ASME Design Engineering Technical Conference <strong>and</strong><br />

Design Automation Conference, Sacrament, CA, Paper<br />

No. 97-DETC/DAC3978; 1997.<br />

[74] Hwang KH, Lee KW, Park GJ. Robust optimization <strong>of</strong> an<br />

automobile rearview mirror for vibration reduction. Struct<br />

Multidisc Optim 2001;21:300–8.<br />

[75] Frahm H. Device for damping vibrations <strong>of</strong> bodies. US<br />

patent No. 989958. 1911.<br />

[76] Den Hartog JP, Ormondroyd J. Theory <strong>of</strong> the dynamic<br />

absorber. Trans ASME 1928;APM50-7:11–22.<br />

[77] Brock J. A note on the damped vibration absorber. ASME<br />

J Appl Mech 1946;68(4):A-284.<br />

[78] Den Hartog JP. Mechanical vibration. Boston: McGraw-<br />

Hill; 1947.<br />

[79] Jones RT, Pretlove AJ, Eyre R. Two case studies <strong>of</strong> the use<br />

<strong>of</strong> tuned vibration absorbers on footbridges. Struct Eng<br />

1981;59B:27–32.<br />

[80] Inman DJ. Engineering vibration. Englewood Cliffs,<br />

NJ: Prentice-Hall; 1994.<br />

[81] Smith JW. Vibration <strong>of</strong> structures: applications in civil<br />

engineering <strong>design</strong>. London: Chapman & Hall; 1988.<br />

[82] Hansen PC. Analysis <strong>of</strong> discrete ill-posed problems by<br />

means <strong>of</strong> the L-curve. SIAM Rev 1992;34:561–80.<br />

[83] Law AM, Kelton WD. Simulation modeling <strong>and</strong> analysis.<br />

3rd ed. Boston: McGraw-Hill; 2000.

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

Saved successfully!

Ooh no, something went wrong!