weighted sum

Upload: debasishmee5808

Post on 11-Oct-2015

14 views

Category:

Documents


0 download

DESCRIPTION

WS

TRANSCRIPT

  • Struct Multidisc Optim (2010) 41:853862DOI 10.1007/s00158-009-0460-7

    RESEARCH PAPER

    The weighted sum method for multi-objectiveoptimization: new insightsR. Timothy Marler Jasbir S. Arora

    Received: 17 April 2009 / Revised: 17 September 2009 / Accepted: 17 November 2009 / Published online: 12 December 2009c Springer-Verlag 2009

    Abstract As a common concept in multi-objective opti-mization, minimizing a weighted sum constitutes an inde-pendent method as well as a component of other methods.Consequently, insight into characteristics of the weightedsum method has far reaching implications. However, despitethe many published applications for this method and the lit-erature addressing its pitfalls with respect to depicting thePareto optimal set, there is little comprehensive discussionconcerning the conceptual significance of the weights andtechniques for maximizing the effectiveness of the methodwith respect to a priori articulation of preferences. Thus,in this paper, we investigate the fundamental significanceof the weights in terms of preferences, the Pareto optimalset, and objective-function values. We determine the factorsthat dictate which solution point results from a particularset of weights. Fundamental deficiencies are identified interms of a priori articulation of preferences, and guidelinesare provided to help avoid blind use of the method.

    Keywords Multi-objective optimization Weighted sum Preferences

    R. T. Marler (B)Center for Computer Aided Design, College of Engineering;111 ERF, The University of Iowa, Iowa City, IA, 52242, USAe-mail: [email protected]

    J. S. AroraCenter for Computer Aided Design, College of Engineering;238 ERF, The University of Iowa, Iowa City, IA, 52242, USAe-mail: [email protected]

    1 Introduction

    Despite deficiencies with respect to depicting the Paretooptimal set, the weighted sum method for multi-objectiveoptimization (MOO) continues to be used extensively notonly to provide multiple solution points by varying theweights consistently, but also to provide a single solu-tion point that reflects preferences presumably incorporatedin the selection of a single set of weights. However, themethods effectiveness in this latter capacity has not beenstudied thoroughly. To be sure, there are many differentapproaches for determining the weights (Marler and Arora2004), but ultimately, however, these are all just differentprocesses for organizing ones preferences and priorities.Instead of proposing yet another algorithm for translatingpreferences into weights, we focus on the mathematicalcharacteristics of the solution and on the fundamental con-ceptual meaning of the weights. Furthermore, we determinethe factors that dictate which Pareto optimal solution pointresults from a particular set of weights. Through this anal-ysis, we augment the currently available literature and pro-vide new insight into the nature of the method and new ideasfor maximizing its effectiveness with respect to articulatingpreferences.

    New understanding of the fundamentals presented in thispaper, is useful in setting the weights directly (without theuse of an additional algorithm). In addition, understandinginadequacies that are inherent in the method is instructive,and in this vein, we present deficiencies that cannot beavoided and that have not been discussed in the currentliterature.

    Setting weights is just one approach to articulating pref-erences, and it is applicable to many different methods.Consequently, understanding how the weights affect thesolution to the weighted sum method has implications

  • 854 R.T. Marler, J.S. Arora

    concerning other approaches that involve similar method-parameters.

    1.1 Overview of the formulation

    The general multi-objective optimization (MOO) problemis posed as follows:

    Minimizex

    : F (x) = [F1 (x) , F2 (x) , , Fk (x)]Tsubject to : g j (x) 0; j = 1, 2, . . . ,m

    (1)

    where k is the number of objective functions and m is thenumber of inequality constraints. x En is a vector ofdesign variables, and F (x) Ek is a vector of objectivefunctions Fi (x) : En E1. The feasible design spaceis defined as X = {x g j (x) 0, j = 1, 2, ...,m

    }. The

    feasible criterion space is defined as Z = {F (x) |x X }.The idea of a solution for (1) can be unclear, because

    a single point that minimizes all objectives simultaneouslyusually does not exist. Consequently, the idea of Pareto opti-mality is used to describe solutions for MOO problems. Asolution point is Pareto optimal if it is not possible to movefrom that point and improve at least one objective functionwithout detriment to any other objective function. Alterna-tively, a point is weakly Pareto optimal if it is not possibleto move from that point and improve all objective functionssimultaneously.

    Solving a particular MOO formulation can serve as anecessary and/or sufficient condition for Pareto optimality,and a methods performance in this capacity is important.However, these characterizations veer slightly from theirmeaning in terms of single-objective optimization. If a for-mulation provides a necessary condition, then for a pointto be Pareto optimal, it must be a solution to that formu-lation. Consequently, the method can yield every Paretooptimal point as a solution. However, some solutions to aformulation that provides a necessary condition may not bePareto optimal. Alternatively, if a formulation provides asufficient condition, then its solution is always Pareto opti-mal, although such a formulation may not be able to captureall of the Pareto optimal points.

    Typically, there are infinitely many Pareto optimal solu-tions for a multi-objective problem. Thus, it is often neces-sary to incorporate user preferences for various objectives inorder to determine a single suitable solution. With methodsthat incorporate a priori articulation of preferences, the userindicates preferences before running the optimization algo-rithm and subsequently allows the algorithm to determinea single solution that presumably reflects such preferences.Alternatively, with a posteriori articulation of preferences,one manually selects a single solution from a representa-tion of the Pareto optimal set. This paper focuses on theuse of the weighted sum method for a priori articulation,

    while its use for a posteriori articulation has been addressedthoroughly in the literature.

    Using the weighted sum method to solve the problem in(1) entails selecting scalar weights wi and minimizing thefollowing composite objective function:

    U =k

    i=1wi Fi (x) (2)

    If all of the weights are positive, as assumed in thisstudy, then minimizing (2) provides a sufficient conditionfor Pareto optimality, which means the minimum of (2)is always Pareto optimal (Zadeh 1963; Goicoechea et al.1982). Although some literature indicates that ki=1 wi = 1and w 0, if any one of the weights is zero, there is apotential for the solution to be only weakly Pareto optimal.

    With regards to the design space, minimizing theweighted sum provides a necessary condition if the multi-objective problem is convex, which means the feasibledesign space is convex (each constraint is convex) and all ofthe objective functions are convex (Geoffrion 1968; Koski1985; Miettinen 1999). A function defined on a convex setis convex if and only if the Hessian matrix of the functionis positive semidefinite or positive definite at all points inthe set. Thus, if the Hessian matrix for each constraint andfor each objective function is positive semidefinite or posi-tive definite, then the weighted sum method can provide allPareto optimal points (the complete Pareto optimal curve).

    With regards to the criterion space, because the Paretooptimal set is always on the boundary of the feasible crite-rion space Z, some studies suggest that Z must be convexfor the weighted sum method to yield all of the Pareto opti-mal points (Stadler 1995; Athan and Papalambros 1996).However, Lin (1975) demonstrates that, in fact, if Z isonly p-directionally convex (Holtzman and Halkin 1966)for any definitely negative vector p, then the weighted sumapproach provides a necessary condition. The term de-nitely negative implies p < 0. p-directionally convex isdefined as follows:

    Definition 1 p-Directionally Convex: Given a nonzero vec-tor p Ek , Z Ek is said to be p-directionally convex ifgiven any two different points in Z, F1 and F2, and any twopositive scalars, w1 and w2, with w1 + w2 = 1, there is apositive number such that w1F1 + w2F2 + pZ.

    This means that only the portion of the feasible criterionspace that is visible when looking in the direction of p fromwithin the feasible space needs to be convex. Therefore,only the Pareto optimal hypersurface needs to be convex forthe weighted sum method to provide a necessary conditionfor Pareto optimality. Currently, the authors do not know ofany method for predicting whether or not the Pareto optimalhypersurface is convex.

  • The weighted sum method for multi-objective optimization 855

    1.2 Review of the literature and motivation

    Following the introduction of the weighted sum methodby Zadeh (1963), the method has been mentioned promi-nently in the literature. However, most general treatments ofMOO simply outline the weighted sum approach and indi-cate that it provides a Pareto optimal solution. Intricacies ofthe method and of the solutions that it yields are typicallynot discussed. In particular, the significance of the weightsis not thoroughly explored, and thus, despite the presence ofmany algorithms for determining weight values, no funda-mental guidelines have been presented for selecting weightsfor accurate a priori articulation of preferences.

    The weighted sum method is often presented strictly asa tool, especially over the past few years, and literatureregarding examples of applications is extensive. However,the focus is on the application, and the problems tend tobe limited to those with just two objective functions. Forinstance, as development for a new approach, Koski andSilvennoinen (1987) provide an early application and usethe weighted sum method to obtain multiple Pareto opti-mal solutions with a systematic change in the weights,while minimizing the volume and the nodal displacementof a four-bar space truss. Kassaimah et al. (1995) usethe method for the two-objective optimization of laminatedplates, where critical buckling shear stress is maximized anddeflection is minimized. A few different somewhat arbitrarysets of weights are considered, and the corresponding solu-tions are compared, but the method itself is not studied andthe articulation of preferences is not considered. Proos et al.(2001) apply the method to topology optimization, mini-mizing compliance and maximizing the first mode of thenatural frequency, for two-dimensional plane stress prob-lems. Again, the weights are altered to yield a representationof the Pareto optimal set. Saramago and Steffen (1998) usea weighted sum to combine two objective functions in anoptimization-based approach to predicting robotic motion,but the weights have the same value and are thus irrelevant.

    Marler and Arora (2005) study a three-objective prob-lem, but they use the weighted sum method as a platform forstudying various function-transformation methods and theiraffect on the depiction of the Pareto optimal set. It is foundthat a convex combination of functions is advantageous fordepicting the Pareto optimal set, as apposed to determin-ing a single solution. In addition, it is shown that sometransformation methods can actually be detrimental to theprocess of obtaining a diverse and uniform spread of Paretooptimal points, and criteria are presented for determiningwhen certain transformation methods (in conjunction withthe weighted sum method) may not perform well.

    Some literature provides a more substantial look at themethod, focusing on its two primary deficiencies whenrepresenting the Pareto optimal set for a posteriori articu-

    lation of preferences. First, many authors demonstrate themethods inability to capture Pareto optimal points that lieon non-convex portions of the Pareto optimal curve (Koski1985; Stadler and Dauer 1992; Stadler 1995; Chen et al.1999; Athan and Papalambros 1996; Huang et al. 2007).Secondly, it is well known that the method does not pro-vide an even distribution of points in the Pareto optimal set,with a consistent change in weights, where even implies aconsistent Euclidean distance between consecutive solutionpoints. Das and Dennis (1997) evaluate why this deficiencyoccurs, and for a two-objective problem, they derive theexplicit form of the unique Pareto optimal curve for whichan even distribution of weights yields an even distribution ofPareto solutions. In general, however, the points that resultfrom a systematic selection of weights may cluster together.These two deficiencies are often answered with alternativemethods for representing a Pareto optimal set (Das andDennis 1998; Tappeta et al. 2000; Messac and Mattson2002; Messac et al. 2003; Huang et al. 2007; Zhang et al.2008). Such alternatives can be effective and valuable, butthey are usually independent of the weighted sum method.

    Although the difficulties mentioned above are well doc-umented, they concern the methods use for yielding acomplete Pareto optimal set rather than a single solution (apriori articulation of preferences). Criteria for maximizingthe effectiveness of the method in the latter respect are notavailable. In fact, there are few, if any numerical studies ofthe weighted sum method that focus on a priori articulationof preferences. Essentially, there is no fundamental unifyinganalysis of the method.

    Many researchers have, however, developed systematicapproaches to selecting weights, surveys of which are pro-vided by Eckenrode (1965), Hobbs (1980), Hwang andYoon (1981), and Voogd (1983). We identify two broadclasses of approaches, but with these approaches, weightsmay still be set ineffectively in terms of accurately articulat-ing preferences. With rating methods, the decision-makerassigns independent values of relative importance to eachobjective function. With ranking methods (Yoon and Hwang1995), which can be viewed as a subset of rating methods,the different objective functions are ordered by importance.Then, the least important objective receives a weight of one,and integer weights with consistent increments are assignedto objectives that are more important. The same approachis used with categorization methods, in which differentobjectives are grouped in broad categories such as highlyimportant and moderately important. Ratio questioning orpaired comparison methods are also common and providesystematic means to rate objective functions by comparingtwo objectives at a time. In this vein, Saaty (1977, 2003),and Saaty and Hu (1998) provide an eigenvalue methodfor determining weights, which involves k(k 1)/2 pair-wise comparisons between objective functions. This yields

  • 856 R.T. Marler, J.S. Arora

    a comparison matrix, and the eigenvalues of the matrix areused as the weights. Rao and Roy (1989) discuss a methodfor determining weights based on fuzzy set theory. Whiledisregarding preferences, Gennert and Yuille (1988) suggestthat the weights should be selected to maximize the finalminimum value of the weighted sum function. For the mostpart, algorithms for determining weights that reflect userpreferences either fall into the category of rating methodsor paired comparison methods.

    Although the above-mentioned algorithms provide validprocesses for determining the weights, there is little rigor-ous investigation as to what the weights actually represent.Thus, this paper studies the significance of the weightsthemselves. By considering the fundamental significance ofthe weights from different viewpoints, we uncover intrinsicflaws in the rating methods and paired comparison meth-ods for a priori articulation of preferences. In addition, weidentify which factors govern which solution is determinedfrom a specific set of weights, and we explain why evenwith an effective algorithm that helps one decide on valuesfor the weights, a weighted sum is not necessarily capableof incorporating preferences completely.

    2 Interpretation of the weights

    One often views the weights as general gauges of relativeimportance for each objective function. However, selectinga set of weights that reflects preference towards one objec-tive or another can be difficult, because preferences tendto be indistinct. In addition, even with full knowledge ofthe objectives and satisfactory selection of weights, the finalsolution may not necessarily reflect the intended preferencesthat are supposedly incorporated in the weights. Nonethe-less, the specific Pareto optimal point that is provided as thesolution depends on which weights are used, so it is impor-tant to determine how the weights relate to preferences,to the Pareto optimal set, and to the individual objectivefunctions. These relationships are explored in this section.

    2.1 Preferences

    First, the relationship between preferences and weights isexamined. Concurrently, the theoretical foundation of rat-ing methods is studied. We discover that such methodsallow one to inadvertently set the weights according to therelative magnitude of the objective functions rather thanthe relative importance of the objectives. In this way, wedemonstrate how function-transformation can be advanta-geous with a priori articulation of preferences. In addition,we reveal inherent deficiencies in the weighted sum methodwith respect to its capacity for incorporating preferences.

    A preference function is an abstract function (of pointsin the criterion space) in the mind of the decision-maker,which perfectly incorporates his/her preferences. MostMOO methods that involve minimizing a single aggregatedobjective function, attempt to approximate the preferencefunction with some mathematical representation, called autility function. The gradients of the preference functionP[F(x)] and the utility function in (2) are given respectivelyas follows:

    xP [F (x)] =k

    i=1

    PFi

    xFi (x) (3)

    xU =k

    i=1wixFi (x) (4)

    Each component of the gradient xP qualitatively repre-sents how the decision-makers satisfaction changes witha change in the design point and a consequent change infunction values. Comparing (3) and (4) suggests that if theweights are selected properly, then the utility function canhave a gradient that is parallel to the gradient of the prefer-ence function. This is significant because the purpose of anyutility function is to approximate the preference function,so imposing similarities between these two functions suchas parallel gradients, can result in a more accurate repre-sentation of ones preferences. If the utility function shouldideally be the same as the preference function, then certainlythe gradients of the two functions should be the same aswell.

    The above relationships indicate that wi representsP/Fi . Conceptually, P/Fi is the approximate changein the preference-function value (change in a decision-makers satisfaction) that results from a change in theobjective-function value for Fi . In this way, P/Fi pro-vides a mathematical definition or metric for the importanceof Fi . However, it only makes sense to consider the impor-tance of an objective or change in preference-function value,in relative terms. Thus, the value of a weight is signifi-cant relative to the values of other weights; the independentabsolute magnitude of a weight is irrelevant in terms ofpreferences.

    2.1.1 Deciencies in the method

    Although (3) and (4) provide a general, conceptual idea ofwhat the weights represent in terms of preferences, usingthem to set precise values for weights can be difficult. Infact, a fundamental deficiency in the weighted sum methodis that it can be difcult to discern between setting weights tocompensate for differences in objective-function magnitudes

  • The weighted sum method for multi-objective optimization 857

    and setting weights to indicate the relative importance of anobjective as is done with the rating methods. For practicalpurposes, P/Fi can be approximated as P/Fi , andwhen objective functions have different ranges and ordersof magnitude, an appropriate value for Fi may not beapparent. For example, consider two objective functions tobe minimized, where function-one is stress, which for thesake of argument ranges between 10,000 and 40,000 psi,and function-two is displacement, which ranges between 1and 2 in. If displacement were to decrease by one unit, theconsequence would be a significant improvement; P/F2would be relatively high. Concurrently, if stress were todecrease by a unit of one, the consequence would be negligi-ble; P/F1 would be insignificant. The result would be arelatively large value for w2, compared to w1. However, thisresult does not necessarily reflect the relative importance ofthe objectives; the values for the weights have been dictatedpurely by the relative magnitudes of the objective functions.

    Alternatively, one could consider percentage changes inthe objective functions rather than an absolute change ofunity as discussed above. However, then the question arisesas to what one should consider when evaluating percent-ages. Should Fi be a percentage of the average value forFi , the maximum value for Fi , the minimum value for Fi ,or some factor of the range of values for Fi? The choice isarbitrary.

    Consequently, one may set the weights to reflectobjective-importance directly, but this approach can alsolead to difficulties. With the previous example, suppose dis-placement is twice as important as stress. Considering therelative value of the weights, if w1 = 1, then w2 = 2.However, given the relative magnitudes of the objectivefunctions, stress would still dominate the weighted sum, andthe use of w2 = 2 would be irrelevant. Again, the magni-tudes of the objective functions affect how the weights areset, and the articulation of preferences becomes blurred atbest.

    This difficulty arises because what dictate the solution tothe weighted sum are the relative magnitudes of the termswi Fi , in (2), not just the magnitudes of the functions orweights alone. The value of a weight is signicant not onlyrelative to other weights but also relative to its correspond-ing objective function. This is a critical idea, although it isoften overlooked. With many weighted methods, includingthe weighted sum method, preferences are articulated a pri-ori by using weights to induce the dominance of a particularterm in a formulation (and the objective function associatedwith the term). However, the process of selecting weightsand thus indicating preferences is complicated if this dom-inance is instead, induced by relative function-magnitude.Thus, when using weights to represent the relative impor-tance of the objectives, transforming the functions so thatthey all have similar magnitudes and do not naturally dom-

    inate the aggregate objective function can help one set theweights to reect preferences more accurately.

    As an example demonstrating the significance of the rel-ative values of the weights, consider the following MOOproblem:

    Find : x = [x1, x2]Tto minimize

    x:

    F1 = 20 (x1 0.75)2 + (2x2 2)2F2 = (x1 2.5)2 + (x2 1.5)2

    (5)

    The design space for this problem is shown in Fig. 1, andthe criterion space is shown in Fig. 2.

    When F1 is minimized independently, it has a minimumvalue of zero, and F2 has a value of 3.3125. When F2 is min-imized independently, it has a minimum value of zero, andF1 has a value of 62.25. F1 has a range between zero and62.25, and F2 has a range between zero and 3.3125. Thus,the range of values for F1 is much larger than that of F2. Ifthis problem is solved using the weighted sum method, andweights are selected with relative values to suggest equalimportance of eth two objectives (i.e. w1 = w2 = 1), thenthe solution is x = (0.833, 1.100) and F = (0.179, 2.938).Yet, even when considered to be equally important, F1 hasa value close to its absolute minimum, and F2 has a valueclose to its maximum. If the weights are set to suggest thatF1 is twice as important as F1 (i.e. w1 = 1, w2 = 2), thenthe solution is x = (0.909, 1.167) and F = (0.179, 2.938).Even when one weight is twice as large as the other, thesolution does not change a significant amount. However,when the objective functions are divided by their respectivemaxima and thus normalized, then it becomes easy to set theweights such that they are significant relative to each otherand relative to the objective functions values. Then, with theobjective functions normalized such that they both a have arange between zero and one, if the weights are again set

    0 1 2 3 4-1

    0

    1

    2

    3

    4

    x1

    x2

    F1

    F2

    A

    D

    B

    C

    Pareto OptimalSet (Segment BC)

    Fig. 1 Example problem design space

  • 858 R.T. Marler, J.S. Arora

    20 40 60 80 100F1

    1

    2

    3

    4

    5

    6F2

    A

    D

    C

    B

    Pareto OptimalSet (Segment BC)

    Fig. 2 Example problem criteria space

    such that w1 = w2 = 1, the solution is x = (1.598, 1.412)and F = (15.054, 0.822).

    Further study of (2), (3), and (4) reveals another funda-mental deficiency in the weighted sum method: the weightedsum utility function is only a linear approximation (in thecriterion space) of the preference function. Equation (2)is a linear function of the objective functions; it is themost basic approximation of ones preference function. Theconsequences of this characteristic are explained as fol-lows. P/Fi may actually be a nonlinear function ofthe objective functions, changing from point to point inthe criterion space. However, using a weighted sum asa utility function to approximate the preference function,inherently assumes that this term is constant, representedby a scalar weight. Thus, values that are selected for theweights are only locally relevant. That is, they are notnecessarily appropriate for the complete criterion space,only for the point at which they are determined, presum-ably the starting point (initial guess). As the design pointchanges with an iterative optimization process, the valuesof the objective functions change, and consequently, prefer-ences (between various points in the criterion space) maychange in a nonlinear fashion. For instance, in the con-text of the stress-displacement example described above,minimizing displacement may be more important than min-imizing stress only until stress reaches a relatively highvalue, suggesting potential failure. Then, minimizing stressbecomes more critical, and its associated weighting valueshould be increased. Anticipating such a change and alteringweighting values while an optimization problem is runningis impractical. Thus, decision-makers implicitly assume thatP/Fi is constant, when they determine a set of weights toreflect the relative importance of the objectives. This meansthat one assumes that the gradient of the preference function(in the criterion space) is constant, which is not necessarilythe case. This is why even with a process that enables one todetermine acceptable values for the weights, the final solu-tion to the weighted sum problem may not accurately reflectinitial preferences that were supposedly incorporated in theweights.

    It is interesting to note that from a computational per-spective, nonlinear composite objective functions often helpto find Pareto optimal solutions on non-convex Pareto opti-mal sets more effectively (Messac et al. 2000a, b; Marler2009). Such functions typically involve various types ofparameters, not just weighting factors. Thus, they providea more complex and versatile utility function for alteringpreferences.

    2.2 Pareto optimal set

    In this section and the next, factors that affect the solutionto the weighted sum method are determined. In addition,fundamental characteristics of the weights are presentedwith regard to the Pareto optimal set, whereas the pre-vious section concerned fundamental characteristics withrespect to preferences. The characteristics discussed hereprovide a theoretical foundation for the paired comparisonmethods. We point out a fundamental deficiency with suchmethods for articulating preferences, and we discover thatfunction transformations can actually be detrimental withthese methods.

    Figure 3 depicts a general illustrative model of a Paretooptimal set in the criterion space, for two objective func-tions.

    The weights represent the gradient of U in (2) withrespect to the vector function F(x), shown as follows:

    FU =

    UF1UF2

    =

    {w1w2

    }(6)

    The Pareto optimal solution (for a given set of weights) isfound by determining where the U -contour with the lowest

    Z

    F1

    F2 Pareto

    Optimal Set

    Direction of Decreasing U-values ( U F )

    Utility Function Contours ( ) ( )1 1 2 2U w F w F= +x x

    Fig. 3 Pareto optimal set in the criterion space

  • The weighted sum method for multi-objective optimization 859

    possible value intersects the boundary of the feasible crite-rion space. The U -contours are linear and are thus tangentto the Pareto optimal curve at the solution point. Changesin the relative values of the weights result in changes in theorientation of the U -contours. The subsequent change in thesolution depends on the shape of the Pareto optimal curve inthe criterion space. This has significant repercussions whentrying to provide an even spread of Pareto optimal points bysolving a series of weighted sum problems, for a posterioriarticulation of preferences. In fact, some more complicatedmethods have been developed so that the shape of the Paretooptimal hypersurface has a minimal affect the accuracy withwhich it is depicted (Marler and Arora 2004).

    In terms of higher dimensions (more than two objec-tive functions), the minimum value of (2) for a fixed setof weights determines a supporting hyperplane to Z (i.e.U (F) = constant) (Zadeh 1963; Gembicki 1974), and thenormal of the hyperplane is the vector of weights. ThePareto optimal hypersurface is always on the boundary of Z,so the supporting hyperplane is tangent to the hypersurfaceat the point that minimizes the weighted sum.

    Steuer (1989) uses such a hyperplane to provide aninterpretation of the weights, which we suggest forms thebasis for paired comparison methods. This interpretationis explained as follows. w1 = 1 is considered a referenceweight, and F1 is the reference function (theoretically, anyfunction can be used as a reference). Then, i represents theamount by which Fi must increase in order to compensatefor a decrease (improvement) of 1 in F1, while remainingon the hyperplane. Each weight is then defined as

    wi = limi0

    1i

    (7)

    If one assumes 1 = 1, then each weight is approximated as

    wi 1i

    (8)

    Thus, wi represents the trade-off between Fi and the ref-erence function, at the solution point to the weighted sumproblem. This is the idea behind paired comparison methodsfor setting weights.

    To clarify this idea, note that one can draw the same con-clusion using Fig. 3 for the case of two objective functions.Considering that the solution to the weighted sum problemis always Pareto optimal, the slope of the Pareto optimalcurve in Fig. 3 is determined as follows:

    dF2dF1

    = w1w2

    (9)

    The left side of (9) can be approximated as F2/F1. Then,assuming w1 = 1 is again a reference weight and F1 = 1(comparable to using 1 = 1 as above), (9) is reduced to

    w2 1F2

    (10)

    There is no negative sign in (8), because the numerator isassumed to represent a decrease of one.

    When using the concepts described above, the objectivefunctions should not be transformed. This is because using(7) involves comparing the different objective functions andevaluating trade-offs (how much of a loss in an objectiveis one willing to sacrifice for a gain in another objective).In contrast to the idea of evaluating relative importance,which was discussed with regard to the rating methods,assigning values to the weights based on trade-offs can behampered by function transformations. Trade-off compar-isons between functions are easier when the functions retaintheir original physical significance and original units. How-ever, most transformation methods yield unitless functions,the values of which have little or no physical significance.

    When using the stress-displacement example, one maybe willing to accept an increase in displacement of 0.002 in.if stress were to reduce by 5,000 psi. If stress is the referencefunction, then a decrease (improvement) of 1 psi must cor-respond to a displacement increase of 5,000 0.002 = 10.Therefore, if the reference weight is w1 = 1, then w2 = 10.However, if the objectives are transformed or normalizedand thus become unitless, it becomes difficult to conductany type of trade-off analysis; it is difficult to conceptualizewhat level of increase in normalized displacement one mightaccept in exchange for a reduction in normalized stress ofone.

    Although (7) is derived based on the mathematical char-acteristics of (2), it provides another general method forarticulating preferences, and here we point out an inher-ent deficiency in this method. Note that (7) is a relationshipthat necessarily exists at the solution point for the weightedsum. However, a users decisions concerning trade-offscan change depending on the point at which the differ-ent objective functions are evaluated. One may be forcedto determine trade-offs based on values for the differentfunctions when the functions are evaluated at a point otherthan the solution point to which (7) applies. Therefore, aswith (3) and (4), (7) only provides an approximate meansfor incorporating preferences in the weighted sum utilityfunction a priori.

    Figure 3 and (6) provide reinforcement of the conclu-sions drawn in Section 2.1 concerning the significance ofthe weights. Note that the direction of the gradient of Uin the criterion space, and thus the orientation of the U-contours and corresponding solution point, depends on therelative values (not the absolute values) of the weights. In

  • 860 R.T. Marler, J.S. Arora

    addition, the orientation of the U-contours relative to thePareto optimal set also depends on the range of objective-function values for points in the set. That is, for a given setof weights (a given gradient vector for the weighted sum),scaling an objective function changes the Pareto optimalsolution point that the weighted sum method provides. Thus,as suggested earlier, the value of a weight is significant rel-ative to the values of other weights and relative to the valueof its corresponding objective function.

    2.3 Objective functions

    Whereas the discussion above considers the significanceof weights in terms of the Pareto optimal set, this sectionfocuses on the relationship between a set of weights andthe objective function values. First, consider a problem withtwo unconstrained objective functions:U = w1F1 (x) + w2F2 (x) (11)For U to have a minimum, it is necessary that the gradientof U be equal to zero, as follows:

    xU = w1xF1 + w2xF2 = 0 (12)Assuming the weights are positive and noting that mini-mizing a weighted sum always provides a Pareto optimalsolution, (12) indicates that at a Pareto optimal point, thegradients of the two objective functions are co-linear andpoint in opposite directions. Essentially, the linear com-bination of the gradients equals zero. In compliance withthe definition of Pareto optimality, this suggests that mov-ing from a solution that satisfies (12), in order to improve afunction, is detrimental to at least one other function.

    General constrained problems are considered usingKarushKuhnTucker necessary conditions, which are sim-plified as follows:

    k

    i=1wixFi =

    ma

    i=1ixgi (13)

    where ma is the number of active constraints and i isthe Lagrange multiplier for gi(x). Equation (13) indicatesthat a linear combination of the objective function gradi-ents is equal and opposite to a linear combination of theconstraint gradients (for active constraints), at a Pareto opti-mal point. If a constraint is active at the solution point for aweighted sum, which necessarily provides a Pareto optimalpoint, then that constraint forms part of the Pareto optimalset. Equations (12) and (13) suggest that the Pareto opti-mal point resulting from the use of a specic set of weightsdepends on active constraints that form part of the Paretooptimal set and on the relationship between the gradients ofthe different objective functions.

    3 Discussion and conclusions

    This paper has provided insight into how the weighted summethod works and has explored the significance of theweights with respect to preferences, the Pareto optimal set,and the objective-function values. We have revealed fun-damental deficiencies with respect to preferences, and wehave shown that although the weighted sum method is easyto use, it provides only a linear approximation of the pref-erence function. Thus, the solution may not preserve onesinitial preferences no matter how the weights are set, a cru-cial idea that is often overlooked. The solution depends onmultiple factors, one of which is the relative magnitude ofthe objective functions. However, when setting the weights,only the relative importance of the objectives should beconsidered, not the relative magnitudes of the function val-ues. In this way, we argue that function-transformation canbe helpful with a priori articulation of preferences, whenthe weights are used to represent the relative importanceof the objectives (as with rating methods). This findingprovides an important guideline that should always be con-sidered when setting weights and/or transforming objectivefunctions. Alternatively, we argue that if the weights areviewed as representing trade-offs between objective func-tions (as with paired comparison methods), then retainingthe original units of the objectives, without transformation,is advantageous.

    Much of the discussion in this paper pertains to a pri-ori articulation of preferences, and although some literaturesuggests that the weights be set such that

    ki=1 wi = 1 and

    w 0, which results in a convex combination of objectivefunctions, when using the weighted sum for a priori articula-tion of preferences, there is no need to place any restrictionon the values of the weights other than w 0, whichensures Pareto optimality. In fact, using unrestricted weightscan be beneficial, making the determination of appropriateweight values easier. However, when systematically alter-ing the weights for a posteriori articulation of preferences,using a convex combination of objective function can helpavoid repeating weighting vectors in terms of relative values(Marler and Arora 2005). Thus, weights should be treateddifferently when used for a priori articulation than they arewhen used for a posteriori articulation.

    Conclusions based on this work and guidelines for settingweights are presented as follows:

    1) When setting weights directly for a priori articulationof preferences, the value of a weight must be sig-nificant relative to other weights and relative to itscorresponding objective function.

    2) When using a ranking method to set weights, all objec-tive functions should be transformed such that they havesimilar ranges.

  • The weighted sum method for multi-objective optimization 861

    3) When using a paired comparison method to set weightsor when viewing the weights as trade-offs, objectivefunctions should not be transformed/normalized.

    4) When articulating well-understood preferences withthe paired comparison methods, unrestricted positiveweights should be used.

    5) The weighted sum provides only a basic approximationof ones preference function. It is fundamentally inca-pable of incorporating complex preference information.Thus, even if one determines acceptable values for theweights a priori, the final solution may not accuratelyreflect initial preferences.

    6) The Pareto optimal solution that results from a specificset of weights depends on constraints that form part ofthe Pareto optimal set, the relationships between thegradients of the different objective functions, the rel-ative magnitudes of the objective functions, and theshape of the Pareto optimal hypersurface.

    7) The weighted sum method provides a basic and easy-to-use approach for multi-objective optimization andis useful as such, but alternate methods should beconsidered for the following cases:

    a. Non-convex or potentially non-convex problemswhere it is critical to depict the complete Paretooptimal set thoroughly

    b. When one wishes to depict the Pareto optimal setfor even a convex problem thoroughly, but needsto minimize computational expenses (minimize thenumber of times (1) must be solved)

    c. When it is critical to articulate complex preferencesaccurately

    References

    Athan TW, Papalambros PY (1996) A note on weighted criteria meth-ods for compromise solutions in multi-objective optimization.Eng Optim 27:155176

    Chen W, Wiecek MM, Zhang J (1999) Quality utilitya compromiseprogramming approach to robust design. J Mech Des 121:179187

    Das I, Dennis JE (1997) A closer look at drawbacks of minimiz-ing weighted sums of objectives for Pareto set generation inmulticriteria optimization problems. Struct Optim 14:6369

    Das I, Dennis JE (1998) Normal-boundary intersection: a new methodfor generating the Pareto surface in nonlinear multicriteria opti-mization problems. SIAM J Optim 8:631657

    Eckenrode RT (1965) Weighting multiple criteria. Manage Sci 12:180192

    Gembicki FW (1974) Performance and sensitivity optimization: avector index approach. PhD dissertation, Case Western ReserveUniversity, Cleveland, OH

    Gennert MA, Yuille AL (1988) Determining the optimal weights inmultiple objective function optimization. In: Second internationalconference on computer vision (held in Los Alamos, CA), Insti-tute of Electrical and Electronics Engineers, Piscataway, NJ, pp8789

    Geoffrion AM (1968) Proper efficiency and the theory of vectormaximization. J Math Anal Appl 22:618630

    Goicoechea A, Hansen DR, Duckstein L (1982) Multiobjective deci-sion analysis with engineering and business applications. Wiley,New York

    Hobbs BF (1980) A comparison of weighting methods in power plantsiting. Decis Sci 11:725737

    Holtzman JM, Halkin H (1966) Directional convexity and the maxi-mum principle for discrete systems. SIAM J Control 4:263275

    Huang C-H, Galuski J, Bloebaum CL (2007) Multi-objective Paretoconcurrent subspace optimization for multidisciplinary design.AIAA J 45:18941906

    Hwang C-L, Yoon K (1981) Multiple attribute decision making, meth-ods and applications: a state-of-the-art survey. In: Beckmann M,Kunzi HP (eds) Lecture notes in economics and mathematicalsystems, no 186. Springer, Berlin

    Kassaimah SA, Mohamed AM, Kolkailah FA (1995) Bi-criteria opti-mum design of laminated plates under uniform load and shear.In: Proceedings of the 27th international SAMPLE technicalconference (held in Albuquerque, NM), 27, pp 731737

    Koski J (1985) Defectiveness of weighting method in multicrite-rion optimization of structures. Commun Appl Numer Methods1:333337

    Koski J, Silvennoinen R (1987) Norm methods and partial weightingin multicriterion optimization of structures. Int J Numer MethodsEng 24:11011121

    Lin JG (1975) Three methods for determining Pareto-optimal solu-tions of multiple-objective problems. In: Ho YC, Mitter SK (eds)Directions in large-scale systems. Plenum, New York

    Marler RT (2009) A study of multi-objective optimization methods forengineering applications. VDM, Saarbrucken

    Marler RT, Arora JS (2004) Survey of multi-objective optimizationmethods for engineering. Struct Multidiscipl Optim 26:369395

    Marler RT, Arora JS (2005) Transformation methods for multi-objective optimization. Eng Optim 37:551569

    Messac A, Mattson CA (2002) Generating well-distributed sets ofPareto points for engineering design using physical programming.Eng Optim 3:431450

    Messac A, Sukam CP, Melachrinoudis E (2000a) Aggregate objectivefunctions and Pareto frontiers: required relationships and practicalimplications. Optim Eng 1:171188

    Messac A, Sundararaj GJ, Tappeta RV, Renaud JE (2000b) Abilityof objective functions to generate points on nonconvex Paretofrontiers. AIAA J 38(6):10841091

    Messac A, Ismail-Yahaya A, Mattson CA (2003) The normalized nor-mal constraint method for generating the Pareto frontier. StructMultidiscipl Optim 25:8698

    Miettinen K (1999) Nonlinear multiobjective optimization. KluwerAcademic, Boston

    Proos KA, Steven GP, Querin OM, Xie YM (2001) Multicriterionevolutionary structural optimization using the weighted and theglobal criterion methods. AIAA J 39:20062012

    Rao JR, Roy N (1989) Fuzzy set theoretic approach of assigningweights to objectives in multicriteria decision making. Int J SystSci 20:13811386

    Saaty TL (1977) A scaling method for priorities in hierarchies, multi-ple objectives and fuzzy sets. J Math Psychol 15:234281

    Saaty TL (2003) Decision-making with the AHP: why is the principaleigenvalue necessary. Eur J Oper Res 145:8591

    Saaty TL, Hu G (1998) Ranking by eigenvector versus other methodsin the analytic hierarchy process. Appl Math Lett 11:121125

    Saramago SFP, Steffen V Jr (1998) Optimization of the trajectory plan-ning of robot manipulators taking into account the dynamics ofthe system. Mech Mach Theory 33:883894

    Stadler W (1995) Caveats and boons of multicriteria optimization.Microcomput Civ Eng 10:291299

  • 862 R.T. Marler, J.S. Arora

    Stadler W, Dauer JP (1992) Multicriteria optimization in engineering:a tutorial and survey. In: Kamat MP (ed) Structural optimiza-tion: status and promise. American Institute of Aeronautics andAstronautics, Washington, DC

    Steuer RE (1989) Multiple criteria optimization: theory, computation,and application. Krieger, Malabar

    Tappeta RV, Renaud JE, Messac E, Sundararaj GJ (2000) Interactivephysical programming: tradeoff analysis and decision making inmulticriteria optimization. AIAA J 38:917926

    Voogd H (1983) Multicriteria evaluation for urban and regional plan-ning. Pion, London

    Yoon KP, Hwang C-L (1995) Multiple attribute decision making, anintroduction. Sage, London

    Zadeh LA (1963) Optimality and non-scalar-valued performance crite-ria. IEEE Trans Automat Contr AC-8:5960

    Zhang K-S, Han Z-H, Li W-J, Song W-P (2008) Bilevel adaptiveweighted sum method for multidisciplinary multi-objective opti-mization. AIAA J 46:26112622

    The weighted sum method for multi-objective optimization: new insightsAbstractIntroductionOverview of the formulationReview of the literature and motivation

    Interpretation of the weightsPreferencesDeficiencies in the method

    Pareto optimal setObjective functions

    Discussion and conclusionsReferences

    /ColorImageDict > /JPEG2000ColorACSImageDict > /JPEG2000ColorImageDict > /AntiAliasGrayImages false /CropGrayImages true /GrayImageMinResolution 150 /GrayImageMinResolutionPolicy /OK /DownsampleGrayImages true /GrayImageDownsampleType /Bicubic /GrayImageResolution 150 /GrayImageDepth -1 /GrayImageMinDownsampleDepth 2 /GrayImageDownsampleThreshold 1.50000 /EncodeGrayImages true /GrayImageFilter /DCTEncode /AutoFilterGrayImages true /GrayImageAutoFilterStrategy /JPEG /GrayACSImageDict > /GrayImageDict > /JPEG2000GrayACSImageDict > /JPEG2000GrayImageDict > /AntiAliasMonoImages false /CropMonoImages true /MonoImageMinResolution 1200 /MonoImageMinResolutionPolicy /OK /DownsampleMonoImages true /MonoImageDownsampleType /Bicubic /MonoImageResolution 600 /MonoImageDepth -1 /MonoImageDownsampleThreshold 1.50000 /EncodeMonoImages true /MonoImageFilter /CCITTFaxEncode /MonoImageDict > /AllowPSXObjects false /CheckCompliance [ /None ] /PDFX1aCheck false /PDFX3Check false /PDFXCompliantPDFOnly false /PDFXNoTrimBoxError true /PDFXTrimBoxToMediaBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXSetBleedBoxToMediaBox true /PDFXBleedBoxToTrimBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXOutputIntentProfile (None) /PDFXOutputConditionIdentifier () /PDFXOutputCondition () /PDFXRegistryName (http://www.color.org?) /PDFXTrapped /False

    /SyntheticBoldness 1.000000 /Description >>> setdistillerparams> setpagedevice