b

DiscoverSearch
About
My stuff
Multiplicative Approximations, Optimal Hypervolume Distributions, and the Choice of the Reference Point
2013·arXiv
Abstract
Abstract

Many optimization problems arising in applications have to consider several objective functions at the same time. Evolutionary algorithms seem to be a very natural choice for dealing with multi-objective problems as the population of such an algorithm can be used to represent the trade-offs with respect to the given objective functions. In this paper, we contribute to the theoretical understanding of evolutionary algorithms for multi-objective problems. We consider indicator-based algorithms whose goal is to maximize the hypervolume for a given problem by distributing  µ points onthe Pareto front. To gain new theoretical insights into the behavior of hypervolume-based algorithms we compare their optimization goal to the goal of achieving an optimal multiplicative approximation ratio. Our studies are carried out for different Pareto front shapes of bi-objective problems. For the class of linear fronts and a class of convex fronts, we prove that maximizing the hypervolume gives the best possible approximation ratio when assuming that the extreme points have to be included in both distributions of the points on the Pareto front. Furthermore, we investigate the choice of the reference point on the approximation behavior of hypervolume-based approaches and examine Pareto fronts of different shapes by numerical calculations.

Multi-objective optimization [15] deals with the task of optimizing several objective functions at the same time. Here, several attributes of a given problem are employed as objective functions and are used to define a partial order, called preference order, on the solutions, for which the set of minimal (maximal) elements is sought. Usually, the objective functions are conflicting, which means that improvements with respect to one function can only be achieved when impairing the solution quality with respect to another objective function. Due to this, such problems usually do not have a single optimal function value. Instead, there is a set of optimal objective vectors which represents the different trade-offs of the different objective functions. Solutions that cannot be improved with respect to any function without impairing another one are called Pareto-optimal solutions. The objective vectors associated with these solutions are called Pareto-optimal objective vectors and the set of all these objective vectors constitutes the Pareto front.

In contrast to single-objective optimization, in multi-objective optimization the task is not to compute a single optimal solution but a set of solutions representing the different trade-offs with respect to the given objective functions. Most of the best-known single-objective polynomially solvable problems like shortest path or minimum spanning tree become NP-hard when at least two weight functions have to be optimized at the same time. In this sense, multi-objective optimization is generally considered as more difficult than single-objective optimization.

Another, more promising, approach to deal with multi-objective optimization problems is to apply general stochastic search algorithms that evolve a set of possible solutions into a set of solutions that represent the trade-offs with respect to the objective functions. Well-known approaches in this field are evolutionary algorithms [2] and ant colony optimization [14]. Especially, multi-objective evolutionary algorithms (MOEAs) have been shown to be very successful when dealing with multi-objective problems [11, 12]. Evolutionary algorithms work with a set of solutions called population which is evolved over time by applying crossover and mutation operators to produce new possible solutions for the underlying multi-objective problem. Due to this populationbased approach, they are in a natural way well-suited for dealing with multi-objective optimization problems.

A major problem when dealing with multi-objective optimization problems is that the number of different trade-offs may be too large. This implies that not all trade-offs can be computed efficiently, i. e., in polynomial time. In the discrete case the Pareto front may grow exponentially with respect to the problem size and may be even infinite in the continuous case. In such a case, it is not possible to compute the whole Pareto front efficiently and the goal is to compute a good approximation consisting of a not too large set of Pareto-optimal solutions. It has been observed empirically that MOEAs are able to obtain good approximations for a wide range of multi-objective optimization problems.

The aim of this paper is to contribute to the theoretical understanding of

MOEAs in particular with respect to their approximation behavior. Many researchers have worked on how to use evolutionary algorithms for multi-object-ive optimization problems and how to find solutions being close to the Pareto front and covering all parts of the Pareto front. However, often the optimization goal remains rather unclear as it is not stated explicitly how to measure the quality of an approximation that a proposed algorithm should achieve.

One popular approach to achieve the mentioned objectives is to use the hypervolume indicator [32] for measuring the quality of a population. This approach has gained increasing interest in recent years (see e. g. [4, 21, 23, 34]). The hypervolume indicator implicitly defines an optimization goal for the population of an evolutionary algorithm. Unfortunately, this optimization goal is rarely understood from a theoretical point of view. Recently, it has been shown in [1] that the slope of the front determines which objective vectors maximize the value of the hypervolume when dealing with continuous Pareto fronts. The aim of this paper is to further increase the theoretical understanding of the hypervolume indicator and examine its approximation behavior.

As multi-objective optimization problems often involve a vast number of Pareto-optimal objective vectors, multi-objective evolutionary algorithms use a population of fixed size and try to evolve the population into a good approximation of the Pareto front. However, often it is not stated explicitly what a good approximation for a given problem is. One approach that allows a rigorous evaluation of the approximation quality is to measure the quality of a solution set with respect to its approximation ratio [27]. We follow this approach and examine the approximation ratio of a population with respect to all objective vectors of the Pareto front.

The advantage of the approximation ratio is that it gives a meaningful scalar value which allows us to compare the quality of solutions between different functions, different population sizes, and even different dimensions. This is not the case for the hypervolume indicator. A specific dominated volume does not give a priori any information how well a front is approximated. Also, the hypervolume measures the space relative to an arbitrary reference point (cf. Section 2.1). This (often unwanted) freedom of choice not only changes the distribution of the points, but also makes the hypervolumes of different solutions measured relative to a (typically dynamically changing) reference point very hard to compare.

Our aim is to examine whether a given solution set of  µsearch points maximizing the hypervolume (called the optimal hypervolume distribution) gives a good approximation measured with respect to the approximation ratio. We do this by investigating two classes of objective functions having two objectives each and analyze the optimal distribution for the hypervolume indicator and the one achieving the optimal approximation ratio.

In a first step, we assume that both sets of  µpoints have to include both optimal points regarding the given two single objective functions. We point out situations where maximizing the hypervolume provably leads to the best approximation ratio achievable by choosing  µPareto-optimal solutions. After these theoretical investigations, we carry out numerical investigations to see how the shape of the Pareto front influences the approximation behavior of the hypervolume indicator and point out where the approximation given by the hypervolume differs from the best one achievable by a solution set of  µpoints. These initial theoretical and experimental results investigating the correlation between the hypervolume indicator and multiplicative approximations have been published as a conference version in [20].

This paper extends its conference version in Section 4 to the case where the optimal hypervolume distribution is dependent on the chosen reference point. The reference point is a crucial parameter when applying hypervolume-based algorithms. It determines the area in the objective space where the algorithm focuses its search. As the hypervolume indicator itself, it is hard to understand the impact of the choice of the reference point. Different studies have been carried out on this topic and initial results on the optimal hypervolume distribution in the dependence of the reference point have been obtained in [1] and [10]. We provide new insights into how the choice of the reference point may affect the approximation behavior of hypervolume-based algorithms. In our studies, we relate the optimal hypervolume distribution with respect to a given reference to the optimal approximation ratio obtainable when having the freedom to choose the  µpoints arbitrarily.

The rest of the paper is structured as follows. In Section 2, we introduce the hypervolume indicator and our notation of approximations. Section 3 gives analytic results for the approximation achievable by the hypervolume indicator under the assumption that both extreme points have to be included in the two distributions and reports on our numerical investigations Pareto fronts having different shapes. In Section 4, we generalize our results and study the impact of the reference point on the optimal hypervolume distribution and relate this choice to the best possible overall approximation ratio when choosing  µpoints. Finally, we finish with some concluding remarks.

image

In this paper, we consider bi-objective maximization problems  P : S → R2for an arbitrary decision space S. We are interested in the so-called Pareto front of P, which consists of all maximal elements of P(S) with respect to the weak Pareto dominance relation. We restrict ourselves to problems with a Pareto front that can be written as  {(x, f(x)) | x ∈ [xmin, xmax]}where  f : [xmin, xmax] →R is a continuous, differentiable, and strictly monotonically decreasing function. This allows us to denote with f not only the actual function  f : [xmin, xmax] →R, but also the front  {(x, f(x)) | x ∈ [xmin, xmax]}itself. We assume further that xmin > 0and  f(xmax) > 0hold.

We intend to find a solution set  X∗ = {x∗1, x∗2, . . . , x∗µ}of  µPareto-optimal search points  (x∗i , f(x∗i ))that constitutes a good approximation of the front f.

image

Figure 1: Point distribution X = {1, 1.6, 2} for the linear front  f : [1, 2] → [1, 2] withf(x) = 3 − x, which achieves a hypervolume of HYP(X) = 1.865 with respect to the reference point r = (0.5, 0.25) and an approximation ratio of APP(X) = 1.25. The shaded areas show the dominated portion of the objective space and the approximated portion of the objective space, respectively.

2.1 Hypervolume indicator

The hypervolume (HYP) measures the volume of the dominated portion of the objective space. It was first introduced for performance assessment in multi-objective optimization by Zitzler and Thiele [32]. Later on it was used to guide the search in various hypervolume-based evolutionary optimizers [4, 16, 21, 24, 30, 34].

Geometrically speaking, the hypervolume indicator measures the volume of the dominated space of all solutions contained in a solution set  X ⊆ Rd. This space is truncated at a fixed footpoint called the reference point  r = (r1, r2, . . . , rd). The hypervolume  HYPr(Y )of a solution set Y in dependence of a given reference point  r = (r1, r2, . . . , rd)is then defined as

image

with VOL(·)being the usual Lebesgue measure (see Figure 1(a) for an illustration).

The hypervolume indicator is a popular second-level sorting criterion in many recent multi-objective evolutionary algorithms for several reasons. Besides having a very intuitive interpretation, it is also the only common indicator that is strictly Pareto-compliant [33]. Strictly Pareto-compliant means that given two solution sets A and B the indicator values A higher than B if the solution set A dominates the solution set B. It has further been shown by Bringmann and Friedrich [9] that the worst-case approximation factor of all possible Pareto fronts obtained by any hypervolume-optimal set of fixed size  µis asymptotically equal to the best worst-case approximation factor achievable by any set of size  µ.

In the last years, the hypervolume has become very popular and several algorithms have been developed to calculate it. The first one was the Hypervolume by Slicing Objectives (HSO) algorithm, which was suggested independently by Zitzler [29] and Knowles [22]. For  d ≤ 3it can be solved in (asymptotically optimal) time O(n log n) [19]. The currently best asymptotic runtime for  d ∈ {4, 5, 6}is  O(n(d−1)/2 log n)[28]. The best known bound for large dimensions  d ≥ 7is  O(n(d+2)/3)[5].

On the other hand, Bringmann and Friedrich [6] proved that all hypervolume algorithms must have a superpolynomial runtime in the number of objectives (unless P = NP). Assuming the widely accepted exponential time hypothesis, the runtime must even be at least  nΩ(d)[8]. As this dashes the hope for fast and exact hypervolume algorithms, there are several estimation algorithms [3, 6, 7] for approximating the hypervolume based on Monte Carlo sampling.

2.2 Approximations

In the following, we define our notion of approximation in a formal way. Let X = {x1, . . . , xµ}be a solution set and f a function that describes the Pareto front. We call a Pareto front convex if the function defining the Pareto front is a convex function. Otherwise, we call the Pareto front concave. Note that this differs from the notation used in [20].

The approximation ratio APP(X) of a solution set X with respect to f is de-fined according to [27] as follows.

Definition 1 Let  f : [xmin, xmax] → Rand  X = {x1, x2, . . . , xµ}. The solution set X is a  δ-approximation of f iff for each  x ∈ [xmin, xmax]there is an  xi ∈ Xwith

image

where  δ ∈ R, δ ≥ 1. The approximation ratio of X with respect to f is defined as

image

Figure 1(b) shows the area of the objective space that a certain solution set X δ-approximates for  δ = 1.25. Note that this area covers the entire Pareto front f. Since the objective vector (1.25, 1.75) is not  δ-approximated for all  δ < 1.25, the approximation ratio of X is 1.25.

Our definition of approximation is similar to the definition of multiplicative  ε-dominance given in [25]. In this paper, an algorithmic framework for discrete multi-objective optimization is proposed which converges to a  (1 + ε)-approximation of the Pareto front.

The goal of this paper is to relate the above definition of approximation to the optimization goal implicitly defined by the hypervolume indicator. Using the

hypervolume, the choice of the reference point decides which parts of the front are covered. In this section we avoid the additional influence of the reference point by considering only solutions where both extreme points have to be included. The influence of the reference point is studied in Section 4. All the functions that we consider in this paper have positive and bounded domains and codomains. Furthermore, the functions that are under consideration don’t have infinite or zero derivative at the extremes. Hence, choosing the reference point  r = (r1, r2)for appropriate  r1, r2 ≤ 0ensures that the points xminand  xmaxare contained in an optimal hypervolume distribution. A detailed calculation on how to choose the reference point such that  xminand  xmaxare contained in an optimal hypervolume distribution is given in [1]. Assuming that  xminand  xmaxhave to be included in the optimal hypervolume distribution, the value of the volume is in this section independent of the choice of the reference point. Therefore, we write HYP(X) instead of  HYPr(X).

image

sets of fixed size  µ. To make this more formal, let  X(µ, f)be the set of all subsets of

image

of cardinality  µwhich contain  (xmin, f(xmin))and  (xmax, f(xmax)). We want to compare two specific solution sets from X called optimal hypervolume distribution and optimal approximation distribution defined as follows.

Definition 2 The optimal hypervolume distribution

image

consists of  µpoints that maximize the hypervolume with respect to f. The optimal approximation distribution

image

consists of  µpoints that minimize the approximation ratio with respect to f. For brevity, we will also use  Xopt(µ, f)in Figures 5–7 as a short form to refer to both sets XHYPopt (µ, f)and  XAPPopt (µ, f).

Note that “optimal hypervolume distributions” are also called “optimal µ-distributions” [1, 10] or “maximum hypervolume set” [9] in the literature.

We want to investigate the approximation ratio obtained by a solution set maximizing the hypervolume indicator in comparison to an optimal one. For this, we first examine conditions for an optimal approximation distribution XAPPopt (µ, f). Later on, we consider two classes of functions f on which the optimal hypervolume distribution  XHYPopt (µ, f)is equivalent to the optimal ap- proximation distribution  XAPPopt (µ, f)and therefore provably leads to the best achievable approximation ratio.

3.1 Optimal approximations

We now consider the optimal approximation ratio that can be achieved placing µpoints on the Pareto front given by the function f. The following lemma states a condition which allows to check whether a given set consisting of  µpoints achieves an optimal approximation ratio for a given function f.

Lemma 1 Let  f : [xmin, xmax] → Rbe a Pareto front and  X = {x1, . . . , xµ}be an arbitrary solution set with  x1 = xmin, xµ = xmax, and  xi ≤ xi+1for all  1 ≤ i < µ. If there is a constant  δ > 1and a set  = Z = {z1, . . . , zµ−1}with  xi ≤ zi ≤ xi+1and  δ = zixi = f(zi)f(xi+1)for all  1 ≤ i < µ, then  X = XAPPopt (µ, f)is the optimal approximation distribution with approximation ratio  δ.

Proof. We assume that a better approximation ratio than  δcan be achieved by choosing a different set of solutions  X′ = {x′1, . . . , x′µ}with  x′1 = xmin, x′µ =xmax, and  x′i ≤ x′i+1, 1 ≤ i < µ, and show a contradiction.

the set X. Each  ziis approximated by a factor of  δ. Hence, in order to obtain a better approximation than the one achieved by the set X, the points  zihave to be approximated within a ratio of less than  δ. We now assume that there is a point  zifor which a better approximation is achieved by the set  X′. Getting a better approximation of  zithan  δmeans that there is at least one point  x′j ∈ X′with  xi < x′j < xi+1as otherwise  ziis approximated within a ratio of at least

image

We assume w. l. o. g. that  j ≤ i + 1and show that there is at least one point z with  z ≤ zithat is not approximated by a factor of  δor that  x′1 > xminholds. To approximate all points z with  xi−1 ≤ z ≤ x′jby a factor of  δ, the inequality xi−1 < x′j−1 < xihas to hold as otherwise  zi−1is approximated within a ratio of more than  δby  X′. We iterate the arguments. In order to approximate all points in  xi−s ≤ y ≤ xi−s+1, xi−s < x′j−s < xi−s+1has to hold as otherwise zi−sis not approximated within a ratio of  δby  X′. Considering  s = j −1either one of the points  z, xi−j+1 ≤ y ≤ xi−j+2is not approximated within a ratio of δby  X′or  xmin = x1 ≤ xi−j+1 < x′1holds, which contradicts the assumption that  X′includes  xminand constitutes an approximation better than  δ.

The case j > i + 1 can be handled symmetrically, by showing that either x′n < xmaxor there is a point  z ≥ zi+1that is not approximated within a ratio of  δby  X′. This completes the proof.

We will use this lemma in the rest of the paper to check whether an approximation obtained by the hypervolume indicator is optimal as well as use these ideas to identify sets of points that achieve an optimal approximation ratio.

3.2 Analytic results for linear fronts

The distribution of points maximizing the hypervolume for linear fronts has already been investigated in [1, 17]. Therefore, we start by considering the

image

Figure 2: Optimal point distribution  XHYPopt (12, f) = XAPPopt (12, f)for (a) the linear front f : [1, 2] → [1, 2] with f(x) = 3 − xand (b,c) the convex fronts  f : [1, c] → [1, c]with f(x) = c/x. The respective optimal hypervolume distributions and optimal approximation distributions are equivalent in all three cases.

hypervolume indicator with respect to the approximation it achieves when the Pareto front is given by a linear function

image

where c < 0 and  d > 1 − care arbitrary constants.

Auger et al. [1] and Emmerich et al. [17] have shown that the maximum hypervolume of  µpoints on a linear front is reached when the points are distributed in an equally spaced manner. We assume that the reference point is chosen such that the extreme points of the Pareto front are included in the optimal distribution of the  µpoints on the Pareto front, that is,  x1 = xmin = 1and  xµ = xmax = (1 − d)/chold. The maximal hypervolume is achieved by choosing

image

due to Theorem 6 in [1]. The following theorem shows that the optimal approximation distribution coincides with the optimal hypervolume distribution.

Theorem 2 Let  f : [1, (1 − d)/c] → [1, c + d]be a linear function  f(x) = c · x + dwhere c < 0 and  d > 1 − care arbitrary constants. Then

image

Proof. We determine the approximation ratio that the optimal hypervolume distribution  XHYPopt (µ, f) = {x1, . . . , xµ}using  µpoints achieves. Let  ˜x, xi <˜x < xi+1, be a Pareto-optimal x-coordinate. The approximation given by  xiand  xi+1is

image

as f is monotonically decreasing. Furthermore, as f is monotonically decreasing, the worst-case approximation is attained for a point  ˜x, xi < ˜x < xi+1, if

image

and resolving the Equation 2 with respect to  ˜x, we get

image

For the approximation ratio we get

image

Hence, the worst-case approximation is independent of the choice of i and the same for all intervals  [xi, xi+1]of the Pareto front. Lemma 1 implies that the hypervolume achieves the best possible approximation ratio on the class of linear fronts.

Figure 2 (a) shows the optimal distribution for  f(x) = 3 − xand  µ = 12.

3.3 Analytic results for a class of convex fronts

We now consider the distribution of  µpoints on a convex front maximizing the hypervolume. In contrast to the class of linear functions where an optimal approximation can be achieved by distributing the  µpoints in an equally spaced manner along the front, the class of functions considered in this section requires that the points are distributed exponentially to obtain an optimal approximation.

As already argued we want to make sure that optimal hypervolume distribution includes  xminand  xmax. For the class of convex fronts that we consider, this can be achieved by choosing the reference point r = (0, 0).

The hypervolume of a set of points  X = {x1, . . . , xµ}, where w. l. o. g.  x1 ≤x2 ≤ · · · ≤ xµ, is then given by

image

We consider a Pareto front given by the function

image

where c > 1 is an arbitrary constant. Then we get

image

Hence, to maximize the hypervolume we have to find  µpoints minimizing

image

Setting  x1 = 1and  xµ = cminimizes h, since  x1and  xµoccur just in the first and last term of h, respectively. Furthermore, we have  1 = x1 < x2 < . . . <xµ = cas the equality of two points implies that one of them can be exchanged for another unchosen point on the Pareto front and thereby increases the hypervolume.

We work under these assumptions and aim to find a set of points X that minimizes the function h. To do this, we consider the gradient vector given by the partial derivatives

image

This implies that h can be minimized by setting

image

From the last equation we get

image

The following theorem shows that the optimal approximation distribution coincides with the optimal hypervolume distribution.

Theorem 3 Let  f : [1, c] → [1, c]be a convex front with f(x) = c/x where c > 1 is an arbitrary constant. Then

image

Proof. We determine the approximation ratio that the optimal hypervolume distribution  XHYPopt (µ, f) = {x1, . . . , xµ}using  µpoints achieves. As f is mono- tonically decreasing, the worst-case approximation is attained for a point x, xi < x < xi+1, if

image

holds. Substituting the coordinates and function values, we get

image

Hence, the set of search points maximizing the hypervolume achieves an approximation ratio of

image

We have seen that the requirements of Lemma 1 are fulfilled. Hence, an application of Lemma 1 shows that the hypervolume indicator achieves an optimal approximation ratio when the Pareto front is given by  f : [1, c] → [1, c]with f(x) = c/x where  c ∈ R>1is any constant.

Figure 2 shows the optimal distribution for  µ = 12and c = 2 as well as c = 200.

3.4 Numerical evaluation for fronts of different shapes

The analysis of the distribution of an optimal set of search points tends to be hard or is impossible for more complex functions. Hence, resorting to numerical analysis methods constitutes a possible escape from this dilemma. This section is dedicated to the numerical analysis of a larger class of functions.

Our goal is to study the optimal hypervolume distribution for different shapes of Pareto fronts and investigate how the shape of such a front influ-ences the approximation behavior of the hypervolume indicator. We examine a family of fronts of the shape  xpwhere p > 0 is a parameter that determines the degree of the polynomial describing the Pareto front. Furthernmore, we allow scaling in both dimensions.

The Pareto fronts that we consider can be defined by a function of the form fp : [x1, xµ] → [yµ, y1]with

image

image

Figure 3: Optimal point distributions for symmetric front  f sym2. Note that the optimal hypervolume distribution and the optimal approximation distribution differ in this case. The set of points maximizing the hypervolume yields an approximation ratio of  APP(XHYPopt (12, f sym2 )) ≈ 1.025, which is 0.457%larger than the optimal approximation ratio  APP(XAPPopt (12, f sym2 )) ≈ 1.021.

We use the notation  yi = f(xi)for the function value  f(xi)of a point  xi. As we assume the reference point to be sufficiently negative, the leftmost point (x1, y1)and the rightmost point  (xµ, yµ)are always contained in the optimal hypervolume distribution as well as in the optimal approximation. We will mainly concentrate on two parameter sets of  fp, that is,

image

Note, that choosing p = 1 corresponds to the well-known test function DTLZ1 [13]. For p = 2 the shape of the front corresponds to functions DTLZ2, DTLZ3, and DTLZ4.

Our goal is to study the optimal hypervolume distribution for our parametrized family of Pareto fronts and relate it to an optimal multiplicative approximation. Therefore, we calculate for different functions  fpand  µ ≥ 3

the set of  µpoints  XHYPopt (µ, fp)which maximizes the dominated hyper- volume, and

the set of  µpoints  XAPPopt (µ, fp)which minimizes the multiplicative ap- proximation ratio.

As in Section 3, we assume that both extreme have to be included in both distributions. For the optimal hypervolume distribution, it suffices to find the x2, x3, . . . , xµ−1that maximize the dominated hypervolume, that is, the solutions of

image

We solve the arising nonlinear continuous optimization problem numerically by means of sequential quadratic programming [18].

image

Figure 4: Optimal point distributions for asymmetric front  f asy2 . Note that the optimal hypervolume distribution and the optimal approximation distribution differ in this case. The set of points maximizing the hypervolume yields an approximation ratio of  APP(XHYPopt (12, f asy2 )) ≈ 1.038, which is 0.839%larger than the optimal approxi- mation ratio  APP(XAPPopt (12, f asy2 )) ≈ 1.030.

In the optimal multiplicative approximation, we have to solve the following system of nonlinear equations

image

with auxiliary variables  z1, . . . , zµ−1due to Lemma 1. The numerical solution of this system of equations can be determined easily by any standard computer algebra system. We used the Optimization package of Maple 15.

In the following, we present the results that have been obtained by our numerical investigations. We first examine the case of  f2. Figures 3 and 4 show different point distributions for  f2. It can be observed that the hypervolume distribution differs from the optimal distribution. Figures 3(a) and 3(b) show the distributions for the symmetric front

image

with  (x1, y1) = (1, 2)and  (xµ, yµ) = (2, 1). Figures 4(a) and 4(b) show the asymmetric front

image

It can be observed that the relative positions of the hypervolume points stay the same in Figures 3(a) and 4(a) while the relative positions achieving an optimal approximation change with scaling (cf. Figures 3(b) and 4(b)). Hence, the relative position of the points maximizing the hypervolume is robust with respect to scaling. But as the optimal point distribution for a multiplicative

image

Figure 5: Approximation ratio of the optimal hypervolume distribution ( ) and the optimal approximation distribution ( ) depending on the scaling  xµof the fronts  fp(cf. Definition 2). We omit the values of the y-axis as we are only interested in the relative comparison ( vs. ) for each front  fp. Note that as analytically predicted in Theorem 2, both curves coincide in (c) for the linear function  f1independent of the scaling.

approximation is dependent on the scaling, the hypervolume cannot achieve the best possible approximation quality.

In the example of Figures 3 and 4 the optimal multiplicative approximation factor for the symmetric and asymmetric case is 1.021 (Figure 3(b)) and 1.030 (Figure 4(b)), respectively, while the hypervolume only achieves an approximation of 1.025 (Figure 3(a)) and 1.038 (Figure 4(a)), respectively. Therefore in the symmetric and asymmetric case of  f2the hypervolume is not calculating the set of points with the optimal multiplicative approximation.

We have already seen that scaling the function has a high impact on the optimal approximation distribution but not on the optimal hypervolume distribution. We want to investigate this effect in greater detail. The influence of scaling the parameter  xµ ≥ 2of different functions  fp : [1, xµ] → [1, 2]is depicted in Figure 5 for p = 1/3, 1/2, 1, 2, 3. For fixed  µ = 3it shows the achieved approximation ratio. As expected, the larger the asymmetry (xµ) the larger the approximation ratios. For concave fronts (p > 1) the approximation ratios seem to converge quickly for large enough  xµ. The approximation of  f2tends towards the golden ratio√5 − 1 ≈ 1.236for the optimal approximation and  4/3 ≈ 1.333for the optimal hypervolume. For  f3they tend towards 1.164 and 1.253, respectively. Hence, for  f2and  f3the hypervolume is never more

image

Figure 6: Approximation ratio of the optimal hypervolume distribution ( ) and the optimal approximation distribution ( ) depending on the number of points  µ for sym-metric and asymmetric fronts  fpand different parameters p (cf. Definition 2). We omit the values of the y-axis as we are only interested in the relative comparison ( vs. ) for each front  fp. Note that (c) and (h) show that the approximation ratio of the optimal hypervolume distribution  APP(XHYPopt (µ, f sym1 ))and the optimal approximation distribution  APP(XHYPopt (µ, f sym1 ))are equivalent for all examined  µ.That maximizing the hypervolume yields the optimal approximation ratio can also be observed for all symmetric  f symp with µ = 3in (a)–(e).

image

Figure 7: Approximation ratio of the optimal hypervolume distribution ( ) and the optimal approximation distribution ( ) depending on the convexity/concavity parameter p for symmetric and asymmetric fronts  fpand different population sizes  µ (cf.Definition 2). The x-axis is scaled logarithmically. We omit the values of the y-axis as we are only interested in the relative comparison ( vs. ) for each front  fp andpopulation size  µ. Note that (a) shows that the approximation ratio of the optimal hypervolume distribution  APP(XHYPopt (3, f symp ))and the optimal approximation distribution  APP(XAPPopt (3, f symp ))are equivalent for all examined p.

than 8% worse than the optimal approximation. This is different for the convex fronts (p < 1). There, the ratio between the hypervolume and the optimal approximation appears divergent.

Another important question is how the choice of the population size influ-ences the relation between an optimal approximation and the approximation achieved by an optimal hypervolume distribution. We investigate the influ-ence of the choice of  µon the approximation behavior in greater detail. Figure 6 shows the achieved approximation ratios depending on the number of points  µ. For symmetric  fp’s with  (x1, y1) = (yµ, xµ)and  µ = 3the hypervolume achieves an optimal approximation distribution for all p > 0. The same holds for the linear function  f1independent of the scaling implied by  (x1, y1)and  (yµ, xµ).

For larger populations, the approximation ratio of the hypervolume distribution and the optimal distribution decreases. However, the performance of the hypervolume measure is especially poor even for larger  µfor convex asymmetric fronts, that is,  f asypwith p < 1 (e.g. Figures 6(f) and 6(g)). Our investigations show that the approximation of an optimal hypervolume distribution may differ significantly from an optimal one depending on the choice of p. An important issue is whether the front is convex or concave [26]. The hypervolume was thought to prefer convex regions to concave regions [31] while [1] showed that the density of points only depends on the slope of the front and not on convexity or concavity. To illuminate the impact of convex vs. concave further, Figure 7 shows the approximation ratios depending on p. As expected, for p = 1 the hypervolume calculates the optimal approximation. However, the influence of the p is very different for the symmetric and the asymmetric test function. For  f sympthe convex (p < 1) fronts are much better approximated by the hypervolume than the concave (p > 1) fronts (cf. Figure 7(a)–(d)). For f asypthis is surprisingly the other way around (cf. Figure 7(e)–(h)).

In all previous investigations, we have not considered the impact of the reference point. To allow a fair comparison we assumed that the optimal approximation distribution and the optimal hypervolume distribution have to include both extreme points. This is clearly not optimal when considering the optimal approximation distribution. Therefore, we relax our assumption and allow any set consisting of  µpoints and raise the question how the optimal approximation distribution looks in this case. Considering the hypervolume indicator, the question arises whether this optimal approximation distribution can be achieved by choosing a certain reference point. Therefore, the goal of this section is to examine the impact of the reference point for determining optimal approximation distributions.

For this we have to redefine parts of the notation. We mark all variables with a hat (like  �) to make clear that we do not require the extreme points to be included anymore.

image

of

image

of cardinality  µ, where we do not assume that  (xmin, f(xmin))and  (xmax, f(xmax))have to be necessarily contained. We also have to redefine the notion of optimal hypervolume distributions and optimal approximation distribution.

Definition 3 The optimal hypervolume distribution

image

consists of  µpoints that maximize the hypervolume with respect to f. The optimal approximation distribution

image

consists of  µpoints that minimize the approximation ratio with respect to f.

4.1 Optimal approximations

Similar to Lemma 1, the following lemma states conditions for an optimal approximation distribution which does not have to contain the extreme points.

Lemma 4 Let  f : [xmin, xmax] → Rbe a Pareto front and  X = {x1, . . . , xµ}a solution set with  xi < xi+1for all  1 ≤ i < µ. If there is a ratio  δ > 1and a set Z = {z1, . . . , zµ−1}with  xi < zi < xi+1for all  1 ≤ i < µsuch that

image

then  X = �XAPPopt (µ, f)is the optimal approximation distribution with approximation ratio  δ.

Proof. For each  i, 1 ≤ i ≤ µ − 1, ziis the worst approximated point in the interval  [xi, xi+1]. Furthermore,  z0 = xminis the worst approximated point in the interval  [xmin, x1]and  zµ = xmaxis the worst approximated point in the interval  [xµ, xmax]. This implies that the approximation ratio of X is

image

Assume  x′i < xi. Consider the point  z′i = δ · x′i. Since  z′i = δ · x′i < δ · xi = zi, we derive  x′i+1 < xi+1as otherwise f(z′i)f(x′i+1) ≥ f(z′i)f(xi+1) > δwould contradict our assumption that  X′achieves an approximation ratio of at most  δ. Repeating the argument  (µ−i)-times leads to  x′µ < xµ, which gives  δ·x′µ < δ·xµ = xmax. This implies that the approximation of  xmaxby  X′is xmaxx′µ > δwhich contradicts the assumption that  X′achieves an approximation ratio of at most  δ.

Assume  x′i > xi. Then all points within  (zi−1, f −1(δ · f(x′i)))are not  δ- approximated. The interval is not empty since  f −1(δ · f(x′i)) > x′i > xi >zi−1due to  δ > 1and f strictly monotonically decreasing. We have another contradiction.

Altogether, we get that  X = �XAPPopt (µ, f)is the unique set achieving an approximation ratio of at most  δand therefore an optimal approximation distribution.

The previous lemma can be used to compute the overall optimal approximation distribution of  µfor a given function describing the Pareto front. In the following, we will use this to compare it to the optimal hypervolume distribution depending on the chosen reference point. Again we consider the class of linear fronts and the class of convex fronts given in Section 3.

4.2 Analytic results for linear fronts

We first consider linear fronts. The optimal multiplicative approximation factor can be easily determined with Lemma 4 as shown in the following theorem.

Theorem 5 Let  f : [1, (1 − d)/c] → [1, c + d]be a linear function  f(x) = c · x + dwhere c < 0 and  d > 1 − care arbitrary constants. Then

image

Proof. Using Lemma 4, we get the following system of  2 (µ+1)linear equations

image

The unique solution of this system of linear equations is

image

which proves the claim.

It remains to analyze the approximation factor achieved by an optimal hypervolume distribution. The impact of the reference point for the class of linear functions has been investigated by Brockhoff in [10]. Using his results, we can conclude the following theorem.

Theorem 6 Let  f : [1, (1 − d)/c] → [1, c + d]be a linear function  f(x) = c · x + dwhere c < 0 and  d > 1 − care arbitrary constants. Let  µ ≥ 2and

image

Then the optimal hypervolume distribution with respect to the reference point r is

image

where

image

Theorem 2 follows immediately from Theorem 3 of Brockhoff [10] by translating their minimization setting into our maximization setting. Knowing the set of points which maximize the hypervolume, we can now determine the achieved approximation depending on the chosen reference point.

Theorem 7 Let  f : [1, (1 − d)/c] → [1, c + d]be a linear function  f(x) = c · x + dwhere c < 0 and  d > 1 − care arbitrary constants. Let  µ ≥ 2and  M1and  M2defined as in Theorem 6, then

image

where

image

Proof. We want to determine the approximation ratio of the optimal hypervolume distribution �XHYPopt (µ, r, f) = {x1, . . . , xµ}as defined in Theorem 6. For this, we distinguish between three cases. The approximation ratio of the inner points  ˜xwith  x1 ≤ ˜x ≤ xµcan be determined as in the proof of Theorem 2. It suffices to plug the definition of  xiand  xi+1from Theorem 6 into equation 2. Let  ˜xbe the solution of this linear equation. Then the inner approximation factor is

image

which is independent of i.

It remains to determine the outer approximation factors. The approximation factor of the points  ˜xwith  1 ≤ ˜x ≤ x1is maximized for  ˜x = 1. The left approximation factor is therefore

image

The approximation factor of the points  ˜xwith  xµ ≤ ˜x ≤ (1−d)/cis maximized for  ˜x = xµ. The right approximation factor is therefore

image

The overall approximation factor is then the largest approximation factor of the three parts, that is,  max{Aℓ, Ac, Ar}.

4.3 Analytic results for a class of convex fronts

We now consider convex fronts and investigate the overall optimal multiplicative approximation first which does not have to include the extreme points. The following theorem shows how such an optimal approximations looks like and will serve later for the comparison to an optimal hypervolume distribution in dependence of the chosen reference point.

Theorem 8 Let  f : [1, c] → [1, c]be a convex front with f(x) = c/x where c > 1 is an arbitrary constant. Then

image

where  xi = c

image

Figure 8: Position of the extremal points of optimal hypervolume distributions depending on the reference point for convex functions  f : [1, c] → [1, c] with f(x) = c/x andc > 1. Depending on the position of the reference point, the leftmost point of a optimal hypervolume distribution  x1is either at the border (x1 = xmin = 1) or inside the domain (x1 > xmin = 1). Similarly, the rightmost point  xµis either at the border (xµ = xmax = c) or inside the domain (xµ < xmax = c). (Note that the figure looks very similar for linear functions.)

Proof. Using Lemma 4, we have  z0 = xmin = 1and  zµ = xmax = c. Furthermore,

image

We have  zi = δxi, 1 ≤ i ≤ µ, and

image

for  1 ≤ i < µ. This implies  xi = δ2i−1, 1 ≤ i ≤ µ. Furthermore,

image

This implies  xi = c

2µ , 1 ≤ i ≤ µand  APP( �XAPPopt (µ, f)) = c12µwhich com-pletes the proof.

Now, we consider the optimal hypervolume distribution depending on the choice of the reference point and compare it to the optimal multiplicative ap- proximation.

Theorem 9 Let  f : [1, c] → [1, c]be a convex front with f(x) = c/x where c > 1 is an arbitrary constant. Then

image

where  xi, 1 ≤ i ≤ µ, depend on the choice of the reference point r as follows:

image

2. If r1 ≤c rµ2and r2 ≤c rµ1, then x1 > 1, xµ< c, xi = (ci rµ−i+11 /ri2)1/(µ+1)for  1 ≤ i ≤µ, and

image

3. If  r2 ≥ c−1/(µ−1), r2 ≤ c, and  r2 ≥ crµ1, then  x1 = 1, xi = (c/r2)(i−1)/µfor 1 ≤ i ≤ µ, and

image

4. If r1 ≥ c−1/(µ−1), r1 ≤c and r1 ≥c rµ2, then xµ =c, xi = r1(c/r1)i/µfor 1 ≤ i ≤µ, and

image

Proof. In order to proof Theorem 9, we distinguish four cases, namely whether x1 = 1or  x1 > 1and whether  xµ = cor  xµ < c. Figure 8 gives an illustration of the four cases.

The first case  x1 = 1and  xµ = ccorresponds to the previous situation where we required that both extreme points are included. The statement of Theorem 9 for this case follows immediately from equations 3 and 4 in Section 3.3.

The second case  x1 > 1and  xµ < cis more involved. First note that we consider only points that have a positive contribution with respect to the given reference point. Therefore, we assume that  r1 < x1and  r2 < f(xµ)holds.

The hypervolume of a set of points  X = {x1, . . . , xµ}, where w. l. o. g.  x1 ≤x2 ≤ · · · ≤ xµ, with respect to a reference point  r = (r1, r2)with  r1 < x1and

r2 < f(xµ)is then given by

image

In order to maximize the hypervolume, we consider the function

image

and compute its partial derivatives.

We have  1 < x1 < x2 < . . . < xµ < cas the equality of two points implies that one of them can be exchanged for another and thereby increases the hypervolume. We work under these assumptions and aim to find a set of points X that minimizes the function h. To do this, we consider the gradient vector given by the partial derivatives

image

This implies that h can be minimized by setting

image

As we can assume  x1 ≥ 1and  xµ ≤ c, we get for  r2 ≤ crµ1and  r1 ≤ crµ2that

image

for  1 ≤ i ≤ µ. It now remains to determine the achieved approximation factor. For this, we proceed as in Theorem 3 and use that

image

This gives an approximation factor of the inner points of

image

For the upper end points the approximation is

image

For the lower end points the approximation is

image

Hence the overall approximation factor in the second case is

image

The third case  x1 = 1and  xµ < cfixes only the left end of the front. Here, we consider the function

image

Not that in contrast to the second case,  h(·)does not depend on  r1. We can assume without loss of generality that  1 = x1 < x2 < . . . < xµ < c. The partial derivatives are therefore

image

This implies that h can be minimized by setting

image

Starting with  x2µ = x2µwe get,

image

and

image

Again using that  x2 ≥ 1and  xµ ≤ cand assuming that  r2 ≤ cand  r2 ≥ c− 1µ−1, we get

image

This results in an approximation factor for the inner points of

image

For the upper end points the approximation is

image

For the lower end points the approximation is

image

Hence the overall approximation factor in the third case is

image

The fourth case  x1 > 1and  xµ = cfixes the right end of the front. We consider the function

image

and compute its partial derivatives

image

This implies that h can be minimized by setting

image

we get

image

Using that  x1 ≥ 1and  xµ−1 ≤ cand assuming  r2 ≥ c−1/(µ−1)and  r2 ≤ cgives

image

This results in an approximation factor for the inner points of

image

For the upper end points the approximation is

image

For the lower end points the approximation is

image

Hence the overall approximation factor for the fourth case is

image

which finishes the proof.

4.4 Numerical evaluation for two specific fronts

We now use the theoretical results of this Section 4 on the approximation factor depending on the reference point and study two specific fronts as an example.

image

Figure 9: Approximation factor of optimal hypervolume distribution depending on the reference point for the linear function  f : [1, 2] → [1, 2] with f(x) = 3 − x for apopulation size of  µ = 10. The right figure shows a closeup view of the area around the reference reference point with the best approximation ratio, which is marked with a red dot.

First, we consider the linear front  f : [1, 2] → [1, 2]with  f(x) = 3 − x. A plot of this front is shown in Figure 2 (a). For  µ = 10, Theorem 5 gives that the optimal distribution

image

achieves the (optimal) approximation of  APP( �XAPPopt (µ, f)) = 31/30.With The- orem 7 we can now also determine the approximation factor of optimal hypervolume distributions depending on the reference point r. For some specific

image

Figure 10: Approximation factor of optimal hypervolume distribution depending on the reference point for the convex function  f : [1, 2] → [1, 2] with f(x) = 2/x for apopulation size of  µ = 10. The right figure shows a closeup view of the area around the reference point with the best approximation ratio, which is marked with a red dot.

reference points we get

image

Figure 9 shows a plot of the approximation factor depending on the reference point. We observe that if  r2 > 32 r1 − 30or  r1 > 32 r2 − 30, the approximation factor is only determined by the inner approximation factor  Ac(cf. Theorem 7). Moreover, for  r1 > 10 r2−8and  r2 > 10 r1−8the approximation factor only depends on  r1and  r2, respectively. For  r ≤ (8/9, 8/9)it is constant. The optimal

image

Let us now consider a specific convex function  f : [1, c] → [1, c]with f(x) = c/x and c = 2 for a population size of  µ = 10. The function is shown in Figure 2(b). Theorem 8 gives that the optimal distribution

image

achieves the (optimal) approximation factor of

image

With Theorem 9 we can determine the approximation factor of optimal hypervolume distributions depending on the reference point r. Figure 10 shows the behavior of the approximation factor depending on the choice of the reference point r. We observe that for  r2 > c rµ1and  r1 > c rµ2, the approximation factor only depends on  r1and  r2, respectively. For

image

the approximation factor is invariably

image

The optimal approximation factor is achieved for the reference point

image

Evolutionary algorithms have been shown to be very successful for dealing with multi-objective optimization problems. This is mainly due to the fact that such problems are hard to solve by traditional optimization methods. The use of the population of an evolutionary algorithm to approximate the Pareto front seems to be a natural choice for dealing with these problems. The use of the hypervolume indicator to measure the quality of a population in an evolutionary multi-objective algorithm has become very popular in recent years. Understanding the optimal distribution of a population consisting of  µindividuals is a hard task and the optimization goal when using the hypervolume indicator is rather unclear. Therefore, it is a challenging task to understand the optimization goal by using the hypervolume indicator as a quality measure for a population.

We have examined how the hypervolume indicator approximates Pareto fronts of different shapes and related it to the best possible approximation ratio. We started by considering the case where we assumed that the extreme points with respect to the given objective functions have to be included in both distributions. Considering linear fronts and a class of convex fronts we have pointed out that the hypervolume indicator gives provably the best multiplicative approximation ratio that is achievable. To gain further insights into the optimal hypervolume distribution and its relation to multiplicative approximations, we carried out numerical investigations. These investigations point out that the shape as well the scaling of the objectives heavily influences the approximation behavior of the hypervolume indicator. Examining fronts with different shapes we have shown that the approximation achieved by an optimal set of points with respect to the hypervolume may differ from the set of  µpoints achieving the best approximation ratio.

After having obtained these results, we analyzed the impact of the reference points on the hypervolume distribution and compared the multiplicative approximation ratio obtained by this indicator to the overall optimal approximation that does not have to contain the extreme points. Our investigations show that also in this case the hypervolume distribution can lead to an overall optimal approximation when the reference point is chosen in the right way for the class of linear and convex functions under investigation. Furthermore, our results point out the impact of the choice of the reference point with respect to the approximation ratio that is achieved as shown in Figures 9 and 10.

Our results provide insights into the connection of the optimal hypervolume distribution and approximation ratio for special classes of functions describing the the Pareto fronts of multi-objective problems having two objectives. For future work, it would be interesting to obtain results for broader classes of functions as well as problems having more than 2 objectives.

[1] A. Auger, J. Bader, D. Brockhoff, and E. Zitzler. Theory of the hypervol- ume indicator: Optimal  µ-distributions and the choice of the reference point. In 10th International Workshop on Foundations of Genetic Algorithms (FOGA ’09), pages 87–102. ACM Press, 2009.

[2] T. B¨ack, D. B. Fogel, and Z. Michalewicz, editors. Handbook of Evolutionary Computation. IOP Publishing and Oxford University Press, Bristol, UK, 1997.

[3] J. Bader and E. Zitzler. HypE: An algorithm for fast hypervolume-based many-objective optimization. Evolutionary Computation, 19:45–76, 2011.

[4] N. Beume, B. Naujoks, and M. T. M. Emmerich. SMS-EMOA: Multiob- jective selection based on dominated hypervolume. European Journal of Operational Research, 181:1653–1669, 2007.

[5] K. Bringmann. An improved algorithm for Klee’s measure problem on fat boxes. Computational Geometry: Theory and Applications, 45(5-6):225–233, 2012.

[6] K. Bringmann and T. Friedrich. Approximating the volume of unions and intersections of high-dimensional geometric objects. Computational Geometry: Theory and Applications, 43:601–610, 2010.

[7] K. Bringmann and T. Friedrich. Approximating the least hypervolume contributor: NP-hard in general, but fast in practice. Theoretical Computer Science, 425:104–116, 2012.

[8] K. Bringmann and T. Friedrich. Parameterized average-case complexity of the hypervolume indicator. In 15th Annual Conference on Genetic and Evolutionary Computation Conference (GECCO), 2013.

[9] K. Bringmann and T. Friedrich. Approximation quality of the hypervol- ume indicator. Artificial Intelligence, 195:265–290, 2013.

[10] D. Brockhoff. Optimal  µ-distributions for the hypervolume indicator for problems with linear bi-objective fronts: Exact and exhaustive results. In 8th International Conference on Simulated Evolution And Learning (SEAL ’10), volume 6457 of LNCS, pages 24–34, 2010.

[11] C. A. Coello Coello, D. A. Van Veldhuizen, and G. B. Lamont. Evolutionary Algorithms for Solving Multi-Objective Problems. Kluwer Academic Publishers, New York, 2002.

[12] K. Deb. Multi-objective optimization using evolutionary algorithms. Wiley, Chichester, UK, 2001.

[13] K. Deb, L. Thiele, M. Laumanns, and E. Zitzler. Scalable multi-objective optimization test problems. In IEEE Congress on Evolutionary Computation (CEC ’02), pages 825–830, 2002.

[14] M. Dorigo and T. St¨utzle. Ant Colony Optimization. MIT Press, 2004.

[15] M. Ehrgott. Multicriteria Optimization. Springer, 2nd edition, 2005.

[16] M. T. M. Emmerich, N. Beume, and B. Naujoks. An EMO algorithm us- ing the hypervolume measure as selection criterion. In Third International Conference on Evolutionary Multi-Criterion Optimization (EMO ’05), pages 62–76. Springer, 2005.

[17] M. T. M. Emmerich, A. H. Deutz, and N. Beume. Gradientbased/evolutionary relay hybrid for computing Pareto front approximations maximizing the S-metric. In Fourth International Workshop on Hybrid Metaheuristics (HM ’07), volume 4771 of Lecture Notes in Computer Science, pages 140–156. Springer, 2007.

[18] R. Fletcher. Practical methods of optimization. Wiley-Interscience, New York, NY, USA, 1987.

[19] C. M. Fonseca, L. Paquete, and M. L´opez-Ib´a˜nez. An improved dimension-sweep algorithm for the hypervolume indicator. In IEEE Congress on Evolutionary Computation (CEC ’06), pages 1157–1163, 2006.

[20] T. Friedrich, C. Horoba, and F. Neumann. Multiplicative approximations and the hypervolume indicator. In 11th Annual Conference on Genetic and Evolutionary Computation (GECCO ’09), pages 571–578. ACM Press, 2009.

[21] C. Igel, N. Hansen, and S. Roth. Covariance matrix adaptation for multi- objective optimization. Evolutionary Computation, 15:1–28, 2007.

[22] J. D. Knowles. Local-Search and Hybrid Evolutionary Algorithms for Pareto Optimization. PhD thesis, Department of Computer Science, University of Reading, UK, 2002.

[23] J. D. Knowles and D. Corne. Properties of an adaptive archiving algorithm for storing nondominated vectors. IEEE Trans. Evolutionary Computation, 7:100–116, 2003.

[24] J. D. Knowles, D. W. Corne, and M. Fleischer. Bounded archiving using the Lebesgue measure. In IEEE Congress on Evolutionary Computation (CEC ’03), volume 4, pages 2490–2497. IEEE Press, 2003.

[25] M. Laumanns, L. Thiele, K. Deb, and E. Zitzler. Combining convergence and diversity in evolutionary multiobjective optimization. Evolutionary Computation, 10(3):263–282, 2002.

[26] G. Lizarraga-Lizarraga, A. Hernandez-Aguirre, and S. Botello-Rionda. G- metric: An m-ary quality indicator for the evaluation of non-dominated sets. In 10th Annual Conference on Genetic and Evolutionary Computation (GECCO ’08), pages 665–672. ACM Press, 2008.

[27] C. H. Papadimitriou and M. Yannakakis. On the approximability of trade- offs and optimal access of web sources. In 41st Annual Symposium on Foundations of Computer Science (FOCS ’00), pages 86–92. IEEE Press, 2000.

[28] H. Yıldız and S. Suri. On Klee’s measure problem for grounded boxes. In ACM Symposium on Computational Geometry (SoCG ’12), pages 111–120, 2012.

[29] E. Zitzler. Hypervolume metric calculation, 2001. Computer Engineering and Networks Laboratory (TIK), ETH Z¨urich, Switzerland, see ftp.tik.ee.ethz.ch/pub/people/zitzler/hypervol.c.

[30] E. Zitzler and S. K¨unzli. Indicator-based selection in multiobjective search. In 8th International Conference on Parallel Problem Solving from Nature (PPSN VIII), volume 3242 of LNCS, pages 832–842. Springer, 2004.

[31] E. Zitzler and L. Thiele. An evolutionary approach for multiobjective opti- mization: The strength Pareto approach. Tik report, Computer Engineering and Networks Laboratory (TIK), ETH Zurich, May 1998.

[32] E. Zitzler and L. Thiele. Multiobjective evolutionary algorithms: A com- parative case study and the strength Pareto approach. IEEE Trans. Evolutionary Computation, 3:257–271, 1999.

[33] E. Zitzler, L. Thiele, M. Laumanns, C. M. Fonseca, and V. Grunert da Fon- seca. Performance assessment of multiobjective optimizers: An analysis and review. IEEE Trans. Evolutionary Computation, 7:117–132, 2003.

[34] E. Zitzler, D. Brockhoff, and L. Thiele. The hypervolume indicator revisited: On the design of Pareto-compliant indicators via weighted integration. In Fourth International Conference on Evolutionary Multi-Criterion Optimization (EMO ’07), volume 4403 of LNCS, pages 862–876. Springer, 2007.


Designed for Accessibility and to further Open Science