multiclass classification with potential function rules: margin distribution and generalization

12
Multiclass classification with potential function rules: Margin distribution and generalization Fei Teng a , Yixin Chen a, , Xin Dang b a Department of Computer and Information Science, The University of Mississippi, University, MS 38677, USA b Department of Mathematics, The University of Mississippi, University, MS 38677, USA article info Article history: Received 12 January 2011 Received in revised form 13 April 2011 Accepted 14 May 2011 Available online 26 May 2011 Keywords: Multiclass classification Consistent classification rules Potential function rules Kernel rules Margin distribution Large margin classifiers Generalization bounds Model selection abstract Motivated by the potential field of static electricity, a binary potential function classifier views each training sample as an electrical charge, positive or negative according to its class label. The resulting potential field divides the feature space into two decision regions based on the polarity of the potential. In this paper, we revisit potential function classifiers in their original form and reveal their connections with other well-known results in the literature. We derive a bound on the generalization performance of multiclass potential function classifiers based on the observed margin distribution of the training data. A new model selection criterion using a normalized margin distribution is then proposed to learn ‘‘good’’ potential function classifiers in practice. & 2011 Elsevier Ltd. All rights reserved. 1. Introduction For thousands of years, various civilizations have observed ‘‘static electricity’’ where pieces of small objects with the same kind of electricity repelled each other and pieces with the opposite kind attracted each other. In pattern recognition and machine learning, potential function rules were motivated from the underlying property of static electricity to predict the unknown binary nature of an observation, a problem commonly known as binary classification. Potential function rules were originally studied by Aizerman et al. and several other researchers in the 1960’s [14,10,17,18]. In its simplest form, a potential function rule puts a unit of positive electrical charge at every positive observation and a unit of negative electrical charge at every negative observation. The resulting potential field defines an intuitively appealing classifier: a new observation is predicted positive if the potential at that location is positive, and negative if its potential is negative. In this paper, we revisit potential function rules (PFRs) in their original form and reveal their connections with other well-known results in the literature. We derive a bound on the generalization performance of potential function classifiers based on the observed margin distribution of the training data. A new model selection criterion using a normalized margin distribution is then proposed to learn ‘‘good’’ potential function classifiers in practice. There is an abundance of prior work in the field of pattern recognition and machine learning. It is beyond the scope of this article to supply a complete review of the area (for more comprehensive surveys on various subjects, the reader is referred to [30,35,14] for pattern recognition, to [58,59] for kernel meth- ods, to [5,45] for computational learning theory, and to [53,42,65] for machine/statistical learning). Nevertheless, a brief synopsis of some of the main findings most related to this article will serve to provide a rationale for the use of PFRs in pattern recognition and machine learning applications. 1.1. The Bayesian decision theory and plug-in decisions A multiclass classification problem aims at foretelling the unknown nature of an observation. More formally, an observation is a d-dimensional vector of numerical measurements denoted as x A R d . The unknown nature of the observation, z, takes values in a finite set K ¼f1, 2, ... , Kg, the set of class labels. A mapping f : R d -K, which is named a classifier, predicts the class label of an observation. Does there exist an ‘‘optimal’’ classifier for a given classifica- tion task? Under a probabilistic setting, the Bayesian decision theory [11,13] gives an affirmative answerthe Bayes decision Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/pr Pattern Recognition 0031-3203/$ - see front matter & 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.patcog.2011.05.009 Corresponding author. Tel.: þ1 662 915 7438; fax: þ1 662 915 5623. E-mail address: [email protected] (Y. Chen). Pattern Recognition 45 (2012) 540–551

Upload: fei-teng

Post on 11-Sep-2016

216 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Multiclass classification with potential function rules: Margin distribution and generalization

Pattern Recognition 45 (2012) 540–551

Contents lists available at ScienceDirect

Pattern Recognition

0031-32

doi:10.1

� Corr

E-m

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

Multiclass classification with potential function rules:Margin distribution and generalization

Fei Teng a, Yixin Chen a,�, Xin Dang b

a Department of Computer and Information Science, The University of Mississippi, University, MS 38677, USAb Department of Mathematics, The University of Mississippi, University, MS 38677, USA

a r t i c l e i n f o

Article history:

Received 12 January 2011

Received in revised form

13 April 2011

Accepted 14 May 2011Available online 26 May 2011

Keywords:

Multiclass classification

Consistent classification rules

Potential function rules

Kernel rules

Margin distribution

Large margin classifiers

Generalization bounds

Model selection

03/$ - see front matter & 2011 Elsevier Ltd. A

016/j.patcog.2011.05.009

esponding author. Tel.: þ1 662 915 7438; fa

ail address: [email protected] (Y. Chen).

a b s t r a c t

Motivated by the potential field of static electricity, a binary potential function classifier views each

training sample as an electrical charge, positive or negative according to its class label. The resulting

potential field divides the feature space into two decision regions based on the polarity of the potential.

In this paper, we revisit potential function classifiers in their original form and reveal their connections

with other well-known results in the literature. We derive a bound on the generalization performance

of multiclass potential function classifiers based on the observed margin distribution of the training

data. A new model selection criterion using a normalized margin distribution is then proposed to learn

‘‘good’’ potential function classifiers in practice.

& 2011 Elsevier Ltd. All rights reserved.

1. Introduction

For thousands of years, various civilizations have observed‘‘static electricity’’ where pieces of small objects with the samekind of electricity repelled each other and pieces with theopposite kind attracted each other. In pattern recognition andmachine learning, potential function rules were motivated fromthe underlying property of static electricity to predict theunknown binary nature of an observation, a problem commonlyknown as binary classification. Potential function rules wereoriginally studied by Aizerman et al. and several other researchersin the 1960’s [1–4,10,17,18]. In its simplest form, a potentialfunction rule puts a unit of positive electrical charge at everypositive observation and a unit of negative electrical charge atevery negative observation. The resulting potential field definesan intuitively appealing classifier: a new observation is predictedpositive if the potential at that location is positive, and negative ifits potential is negative.

In this paper, we revisit potential function rules (PFRs) in theiroriginal form and reveal their connections with other well-knownresults in the literature. We derive a bound on the generalizationperformance of potential function classifiers based on the

ll rights reserved.

x: þ1 662 915 5623.

observed margin distribution of the training data. A new modelselection criterion using a normalized margin distribution is thenproposed to learn ‘‘good’’ potential function classifiers in practice.There is an abundance of prior work in the field of patternrecognition and machine learning. It is beyond the scope of thisarticle to supply a complete review of the area (for morecomprehensive surveys on various subjects, the reader is referredto [30,35,14] for pattern recognition, to [58,59] for kernel meth-ods, to [5,45] for computational learning theory, and to [53,42,65]for machine/statistical learning). Nevertheless, a brief synopsis ofsome of the main findings most related to this article will serve toprovide a rationale for the use of PFRs in pattern recognition andmachine learning applications.

1.1. The Bayesian decision theory and plug-in decisions

A multiclass classification problem aims at foretelling theunknown nature of an observation. More formally, an observationis a d-dimensional vector of numerical measurements denoted asxARd. The unknown nature of the observation, z, takes values in afinite set K¼ f1,2, . . . ,Kg, the set of class labels. A mappingf : Rd-K, which is named a classifier, predicts the class label ofan observation.

Does there exist an ‘‘optimal’’ classifier for a given classifica-tion task? Under a probabilistic setting, the Bayesian decisiontheory [11,13] gives an affirmative answer—the Bayes decision

Page 2: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551 541

rule (or called the Bayes classifier). If the pair of observation andits nature, (x, z), is a random variable with a joint probabilitydistribution p(x, z), the Bayes classifier, fn, selects the class labelfor an observation x as f nðxÞ ¼ argmaxzAKPrðzjxÞ ¼ argmaxzAK

pðx,zÞ. The optimality of fn is defined by the minimum probabilityof error, i.e., Pr½f nðxÞaz�rPr½f ðxÞaz� for any f : Rd-K, which iswell known as the Bayesian probability of error. This probabilitymeasures the ‘‘hardness’’ of a classification problem. It cantheoretically be evaluated if the joint distribution is known, butthe calculation may be (and usually is) intractable in practice dueto the min operator inside the integral. Several tight bounds wereproposed in the literature for computational approximations ofthe Bayesian probability of error [25,41,6].

The crux of the Bayesian approach is the difficulty of deter-mining the joint distribution. Plug-in decision [30] is a naturalway of applying the Bayesian classification in practice, where anapproximated Bayes classifier is constructed using an estimatedjoint distribution. Depending upon the way in which the jointdistribution is estimated, plug-in decision rules fall roughly intothe parametric approaches and the nonparametric approaches.

In a parametric approach, the unknown joint distribution isdescribed by a set of parameters based on certain structuralassumptions, e.g., conditional independence of attributes withineach class [49,34,15,50], mixture of Gaussians [46,61,19], andmixture of Bernoullis [61]. The values of the parameters areobtained by optimizing a loss function, e.g., a likelihood function.In many applications, a parametric approach presents an efficientmeans of incorporating prior knowledge about the data.For example, Hofmann et al. [43] used a latent variable model(aspect model) to remove the statistical dependence among wordsin a document for textual data. Barnard et al. [7] explored severalgenerative models to describe statistical relevance betweenimages regions and associated texts. Veeramachaneni and Nagy[66] studied the interpattern dependence, named style context, foroptical character recognition. Intraclass style (statistical depen-dence between patterns of the same class in a field) and interclassstyle (statistical dependence between patterns of different classesin the same field) were formalized to derive style-constrainedBayesian classification.

The performance of a plug-in decision rule is determined by thequality of the estimated joint distribution. Ben-Bassat et al. [12]analyzed the sensitivity of Bayesian classification under multiplica-tive perturbation on the joint distribution. Devroye [26] presented amore general result showing that if the estimated posterior prob-ability is close to the true posterior probability in L1-sense, the errorprobability of the plug-in decision rule is near the Bayesian prob-ability of error. Nevertheless, does the error probability converge tothe Bayesian probability of error if more training samples areobtained to approximate an arbitrary joint distribution? This is aquestion regarding the universal consistency of a classification rule.Loosely speaking, a universally consistent rule [30] guarantees us thattaking more samples suffices to roughly reconstruct an arbitrary,fixed, but unknown distribution, hence to asymptotically achievethe optimality. While parametric approaches are efficient, in generalthey are not universally consistent.

In 1977, Stone proved the existence of a universally consistentrule [60]. He showed that any k-nearest neighbor classifier isuniversally consistent if k is allowed to grow with n, the samplesize, at a speed slower than that of n. Since then several rules havebeen shown to be universally consistent including histogramrules [37] and kernel rules [30]. We put these approaches underthe category of nonparametric plug-in decisions because of theunderlying nonparametric estimation of joint distributions.Representing all the data with a nonparametric model is some-times preferred over summarizing it with a parametric modelbecause of the rich detail held by very large data sources [40].

1.2. Classifier selection and complexity regularization

Universal consistency describes the asymptotic behavior of aclassifier, i.e., the number of training samples goes to infinity.For real-life problems, however, the size of a training set is finiteand, usually, fixed. This leads to a basic question in classifierdesign: how do we select a classifier, which performs well onfuture examples, from a given set of classifiers based on a givenfinite training set? Two basic principles were investigated in theliterature for classifier selection: empirical risk minimization [63]and complexity regularization [51].

In order to achieve good generalization performance, theempirical risk minimization principle seeks for a classifier thatminimizes the training error (empirical risk). Vapnik and Chervo-nenkis presented a theoretical ground for empirical risk mini-mization [63]. It was shown that if the ‘‘capacity’’ of C, the set ofclassifiers to choose from, is sufficiently restricted, minimizingthe empirical risk guarantees a classifier whose performance isclose to that of the best classifier in C. Here the capacity of C isdefined by the VC-dimension of C. The above result reveals twocompeting factors in classifier selection. On one hand, a lowcapacity model set may not contain any classifier that generalizeswell. On the other hand, too much freedom may over fit the dataresulting a model behaving like a refined look-up-table: perfectfor the training data but poor on generalization.

This suggests that a classifier, built on a finite training set,generalizes the best if the right tradeoff is found between thetraining accuracy and the capacity of the model set. Complexityregularization applies the above idea to search for a classifier thatminimizes the sum of empirical risk and a term penalizing thecomplexity [64,8,51,9,16]. Amongst various definitions of thepenalty term, margin-based approaches received broad attentionin the literature. A series of results were obtained that exhibit theintrinsic connection between generalization and different mea-sures of margin distribution (e.g., maximal margin, marginpercentile, soft margin) [65,48,59,55,39]. These theoretical resultsled to the discovery of new learning algorithms (e.g., supportvector machines [65], margin distribution optimization [36], largemargin multiple-instance learning [21], margin trees [62], largemargin semi-supervised learning [67], dissimilarity-based learn-ing [54], similarity-based learning [52,22], large margin nearestneighbor classification [69,44]), large margin conditional randomfields [47], and new interpretations of known learning algorithms(e.g., boosting [57,56], additive fuzzy systems [20]).

Classifiers derived from complexity regularization are notnecessarily consistent. Lugosi and Zeger [51] presented a suffi-cient condition for universal consistency of a particular method ofcomplexity regularization, structural risk minimization, usingVapnik-Chervonenkis complexity classes [65]. As discussed inSection 1.1, several nonparametric plug-in decision rules areuniversally consistent. A question arises naturally: Can wecombine complexity regularization with a universally consistentnonparametric plug-in decision rule to improve the generaliza-tion performance thereof? This is the question that we attempt toaddress in this article. In particular, we focus on the regularizationof multiclass PFRs based on a margin distribution.

1.3. A historical timeline of potential function method

Potential function method was invented by Russian research-ers in the 1960’s to model a general function reconstructionproblem [1–4,10,17,18]. Various algorithms were proposed totackle the reconstruction problem for specialized purposes. In aseries of three articles, Aizerman et al. proposed algorithms usinga potential function for binary classification [1], estimation of the

Page 3: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551542

posterior probability of binary classes [2], and approximatinga functional relationship [3].

Under the specific problem settings, various convergence resultswere obtained [1,17,18,4]. In [1], a learning method for potentialfunction classifier was shown to converge in finite number ofiterations when the two classes are separable by the given familyof potential functions. Under the same assumption, a probabilisticbound was derived in [4] that can be used as a terminationcondition for the training process. The bound is essentially a samplecomplexity bound in computational learning theory [45]. Bravermanand Pyatnitskii [17] established criteria for the choice of potentialfunction used in convergent learning algorithms. In a later work,they estimated the rate of convergence [18].

In the simplest form, a binary PFR views each training sampleas an electrical charge, positive or negative according to its classlabel. The resulting potential field divides the feature space intotwo decision regions based on the polarity of the potential. In theoriginal formulation of PFR [1,17], the electrical charge of eachtraining sample is assigned a weight coefficient. In this article, weconsider a simpler form with uniform weights. The basic idea ofbinary PFRs can be generalized to the multiclass scenario,in which a potential function is defined for each class using thetraining observations within that class. A new observation is thenassigned a label corresponding to the class of the highestpotential value. In addition to the intuitive appealingness, goodscalability is a main advantage of PFRs in applications of a‘‘dynamic’’ nature, i.e., the structure of the problem may varyover time. Adding new classes does not affect the existingpotential functions. Removing or merging classes influence onlythe potential functions of the classes involved in the operation.

1.4. An overview of the article

The contributions of this article are given as follows:

Connections of PFRs with the Bayes decision theory. Given chargedensity functions a priori, we present conditions under whicha PFR is essentially optimal under the framework of theBayesian decision theory. We then look into a more practicalscenario where a PFR is built from a given set of trainingobservations with unknown but fixed charge density func-tions. We show that a PFR is, in this case, equivalent to a plug-in decision rule using kernel density estimation, hence uni-versally consistent. � A new generalization bound for PFRs. We discuss the classifier

selection for PFRs using complexity regularization. An upperbound on the generalization performance for PFRs is derivedusing a margin distribution.

� A simple classifier selection method for PFRs. Motivated by the

above generalization bound, we propose a simple kernelselection method using a normalized margin distribution.Extensive experimental results on real applications demon-strate the competitive performance of the proposedframework.

The remainder of the paper is organized as follows: Section 2discusses PFRs from the perspective of the Bayesian decisiontheory. Conditions under which PFRs are equivalent to Bayesdecision rules are presented. Section 3 shows a connectionbetween a PFR and a plug-in decision rule using kernel densityestimation. Section 4 presents a generalization bound for PFRsbased on a margin distribution. Motivated by this bound on thegeneralization performance of PFRs, we propose in Section 5 amodel selection method using a normalized margin distribution.In Section 6, we explain the experimental studies conducted anddemonstrate the results. We conclude in Section 7.

2. Potential function rules and the Bayes decision theory

We start with a brief review of electrostatic potential functions[38]. We then introduce the general form of binary potentialfunction classifiers. Finally, we demonstrate connections betweenPFRs and the Bayes classifiers.

2.1. Potential function rules

Given a positive point charge at location y, the electrostaticpotential at location x is proportional to 1=Jx�yJ, which is calledthe electrostatic point potential function. For a ‘‘cloud’’ of positivecharges with density rþ over a space X, the electrostaticpotential function F is, modular a constant scale factor,

FðxÞ ¼ZX

rþ ðyÞJx�yJ

dy:

Therefore, if rþ and r� are, respectively, the charge density ofpositive and negative charges over X, the electrostatic potentialfunction F is, modular a constant scale factor, defined as

FðxÞ ¼ZX

rþ ðyÞJx�yJ

dy�ZX

r�ðyÞJx�yJ

dy:

The above electrostatic potential function can be generalizedby replacing the electrostatic point potential function with ageneral point potential function c : X�X-R:

FðxÞ ¼ZXrþ ðyÞcðx,yÞ dy�

ZXr�ðyÞcðx,yÞ dy: ð1Þ

Note that the electrostatic potential at a location x is not welldefined if x falls in the support of rþ or r� due to the fact that1=Jx�yJ is1when x¼y. This limitation, however, can be avoidedby a general potential function (1) with a proper choice of thepoint potential function c.

Given rþ and r�, let Qþ and Q� be the total positive chargeand negative charge, respectively:

Qþ ¼

ZXrþ ðxÞ dx, Q� ¼

ZXr�ðxÞ dx:

We normalize the potential function (1) by the sum of the totalpositive and total negative charges:

FðxÞQþ þQ�

¼Qþ

Qþ þQ�

ZX

rþ ðyÞQþ

cðx,yÞ dy�Q�

Qþ þQ�

ZX

r�ðyÞQ�

cðx,yÞ dy:

It is not difficult to check that rþ ðxÞ=Qþ and r�ðxÞ=Q� can beviewed as probability density functions because they are non-negative over X and

RXðrþ ðxÞ=Qþ Þdx¼

RXðr�ðxÞ=Q�Þdx¼ 1, i.e.,

normalized charge densities are probability density functions.Therefore, we define conditional probability density functions as

pðxjþÞ ¼rþ ðxÞ

Qþ, pðxj�Þ ¼

r�ðxÞQ�

ð2Þ

and prior probability as

PrðþÞ ¼Qþ

Qþ þQ�, Prð�Þ ¼

Q�Qþ þQ�

: ð3Þ

Consequently, the above normalized potential function is rewrit-ten in terms of (2) and (3) as

FðxÞQþ þQ�

¼ PrðþÞ

ZX

pðyjþÞcðx,yÞ dy�Prð�Þ

ZX

pðyj�Þcðx,yÞ dy:

Hence a binary potential function classifier is defined as

f ðxÞ ¼ signðFðxÞÞ

¼ sign PrðþÞ

ZX

pðyjþÞcðx,yÞ dy�Prð�Þ

ZX

pðyj�Þcðx,yÞ dy

� �,

ð4Þ

i.e., the polarity of the potential determines the class label.

Page 4: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551 543

2.2. PFRs and Bayes classifiers

Next, we present a Bayesian interpretation of the above potentialfunction classifier. In particular, we show that with a proper choiceof c, the decision boundary of (4) is identical to that of the optimalBayes classifier. Our first choice of c is the Dirac delta functionwhich is zero everywhere except at the origin, where it is infinite,

dðxÞ ¼þ1, x¼ 0,

0, xa0

(

and which also satisfies the identityZ 1�1

dðxÞ dx¼ 1:

Theorem 1. Let rþ and r� be the charge densities; pðxjþÞ, pðxj�Þ,PrðþÞ, and Prð�Þ be defined by (2) and (3). If we choose

cðx,yÞ ¼ dðx�yÞ, the decision boundary of the potential function

classifier (4) is equivalent to that of the Bayes classifier for condi-

tional probability distributions pðxjþÞ and pðxj�Þ, and class prior

probabilities PrðþÞ and Prð�Þ.

A proof of Theorem 1 is given in the Appendix. We may interpret theabove theorem from the perspective of Fourier analysis. Specifically,for a translation invariant point potential function, i.e., cðx,yÞ ¼cðx�yÞ, the evaluation of

RXpðyjþÞcðx�yÞdy is essentially the

convolution of pðxjþÞ and cðxÞ, which is equivalent to computingthe inverse Fourier transform of the product of the Fourier transformsof pðxjþÞ and cðxÞ. When c is the Dirac delta function, potentialfunction classifiers are equivalent to Bayes classifiers because theFourier transform of the Dirac delta function is the constant 1.

Theorem 1 holds independent of the specific forms of thecharge densities, i.e., it is distribution free. Nevertheless, theunboundedness of the Dirac delta function makes it a poor choicein numerical implementations. Next, by assuming that the Fouriertransform of the charge densities have finite support, we extendthe conclusion of Theorem 1 to a wider class of translationinvariant point potential functions.

Theorem 2. Let rþ and r� be the charge densities; pðxjþÞ, pðxj�Þ,PrðþÞ, and Prð�Þ be defined by (2) and (3). Let pþ ðxÞ and p�ðxÞ be

the Fourier transform of pðxjþÞ and pðxj�Þ, respectively, i.e.,

pþ ðxÞ ¼ZX

pðxjþÞe�2pixT x dx,

p�ðxÞ ¼ZX

pðxj�Þe�2pixT x dx,

where i is the complex numberffiffiffiffiffiffiffi�1p

. We assume that pþ and p� have

finite support, namely, there exist constants sþ and s� such that

pþ ðxÞ ¼ 0 for JxJZsþ and p�ðxÞ ¼ 0 for JxJZs�. For any

translation invariant point potential function cðx,yÞ ¼cðx�yÞ, if its

Fourier transform satisfies that CðxÞ ¼ 1 for JxJos¼maxðsþ ,s�Þ,the decision boundary of the potential function classifier (4) is identical

to that of the Bayes classifier using conditional probability distribu-

tions pðxjþÞ and pðxj�Þ, and class prior probabilities PrðþÞ and Prð�Þ.

A proof of Theorem 2 is given in the Appendix. The above theoremstates that if the charge densities are ‘‘band limited’’ (i.e., itsFourier transform is zero everywhere outside a hyperball of finiteradius s) and the point potential function has value 1 over thesupport of charge densities in the frequency domain, the potentialfunction conveys the same information as the class conditional-density. In the one dimensional case, a possible choice of c is asinc function,

cðx,yÞ ¼sin½2psðx�yÞ�

pðx�yÞ¼ 2s � sinc½2sðx�yÞ�,

whose Fourier transform is a rectangular window function

CðoÞ ¼1, jojrs

0, joj4s

(¼ rect

o2s

� �:

This choice of C can be generalized to higher dimensional spaces:for a hyper-rectangular window function in a d-dimensionalfrequency domain,

CðxÞ ¼1, joijrs,8iA ½1,d�

0, joij4s,(iA ½1,d�

(¼Yd

i ¼ 1

rectoi

2s

� �

the corresponding point potential function is

cðx,yÞ ¼ ð2sÞdYd

i ¼ 1

sinc½2sðxi�yiÞ�: ð5Þ

Theorem 2 has implications on the practical design of potentialfunction classifiers using a finite training set. This will bediscussed in the next section.

In statistics, an idea similar to Theorem 2 had been explored innonparametric density estimation. Watson and Leadbetter [68]discussed the L2 error of kernel density estimates and related it tothe spectral property of the density function (i.e., characteristicfunction). They concluded that the form of the optimal estimatedepends critically on the tail of the characteristic function. Davis[23,24] showed that using a sinc kernel, with a carefully chosenscale factor, the Fourier integral error estimate is asymptoticallyoptimal within a constant factor for all densities. The analyses in[68,23,24] were performed under the L2 measure. Devroye [27–29]developed a series of asymptotic performance bounds for kernelestimates using the L1 measure. Compared with the L2 error, the L1

error has a clearer physical interpretation. As consistent densityestimates yield consistent classifiers, a plug-in decision rule usingthese kernel density estimates is naturally consistent.

3. Potential function rules as plug-in decision rules

The main difficulty of using the potential function classifier (4) inpractice is that charge densities are usually unknown. An approx-imation method is therefore presented in this section. Next, we firstgeneralize the above binary potential function classifier to multipleclasses. All the results discussed in Section 2 can be extended to themulticlass scenario. We then present an approximation on PFR and adiscussion on its connection with plug-in decision rules.

3.1. An approximation on multiclass potential function classifiers

Let zAK¼ f1, . . . ,Kg be the class label of observation xAX. Theobservation–label pair (x,z) is generated by a distribution F, whichis a mixture of K unknown distributions F1, . . . ,FK ,

F ¼XK

k ¼ 1

PkFk,

where Pk is the marginal probability of label k, i.e., Pk ¼ Prðz¼ kÞ;Fk is the cumulative distribution function of x conditioned onz¼k. Analogous to (1), (2), and (3), we define Fk as a classpotential function—the potential with respect to PkFk:

FkðxÞ ¼ Pk

ZXcðx,yÞdFkðyÞ: ð6Þ

A multiclass potential classifier is defined as

f ðxÞ ¼ argmaxk

FkðxÞ: ð7Þ

Page 5: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551544

Note that the class potential (6) is the product of Pk and theexpectation of the point potential function c with respect to Fk, i.e.,

FkðxÞ ¼ PkEy � Fk½cðx,yÞ�:

Although F is unknown in most applications, a training set isusually given. Therefore, we approximate the above expectation bythe sample mean. Let S ¼ fðx1,z1Þ, . . . ,ðx‘ ,z‘Þg �X�K be the train-ing set, a random i.i.d. sample from F.

Definition 1 (Sample class potential function). Given a pointpotential function c : X�X-R, we define the sample classpotential of an observation x with respect to class k and sample S as

fkðx,SÞ ¼ 1

jSjX

zi ¼ k

cðx,xiÞ: ð8Þ

A multiclass sample potential classifier is then defined usingsample class potential functions as follows.

Definition 2 (A multiclass sample potential function classifier).Given S, a set of i.i.d. training samples generated by an unknown

distribution F on X�K, we define a potential classifier fS : X-K as

fSðxÞ ¼ argmaxk

fkðx,SÞ: ð9Þ

Clearly, the sample class potential (8) can be written as

fkðx,SÞ ¼ jSkj

jSj1

jSkj

Xzi ¼ k

cðx,xiÞ ,

where Sk ¼ fðx,zÞAS : z¼ kg. It is not difficult to observe thatjSkj=jSj is an estimate of the marginal probability Pk. Furthermore,

if we restrict c to be a nonnegative translation invariant function

andRXcðxÞdx¼ co1, it is straightforward to show that

ð1=cÞjSkjP

zi ¼ kcðx,xiÞ is an estimate of the probability density of

Fk at location x using the kernel density estimation ðc is the

kernel function). Hence, for any given x, fkðx,SÞ is proportional to

an estimation of the posterior probability Prðz¼ kjxÞ.This implies that the family of multiclass potential function

classifiers (9) includes those plug-in decision Bayes classifiers thatuse kernel density estimation. Therefore, if c is chosen fromregular kernels, the universal consistency of PFRs follows from theuniversal consistency of kernel rules [30]. Universal consistencycharacterizes an asymptotic property of a decision rule—a deci-sion rule converges to the optimal solution as the number oftraining sample is sufficiently large. For kernel rules, universalconsistency requires the ‘‘width’’ of the kernel to decrease to 0 asthe sample size increases to infinity. Next, we show that underthe conditions of Theorem 2, for a fixed width of c (i.e., 1=s in (5)is fixed), with high probability the prediction of a sample PFRconverges to that of the Bayes classifier for any given input.

It is worth noting that when the point potential function1=Jx�yJ is used, the above PFR is called Hilbert rule. Hilbert ruleswere investigated by Devroye et al. for density estimation,regression and classification [31–33]. An interesting property ofHilbert rules is that it does not have a smoothing parameter.Consistency theory of Hilbert rules was developed in [31–33].

3.2. The potential gap and the generalization performance

For a set of numbers a1, . . . ,aK , the k-th smallest number isdenoted by aðkÞ, i.e., að1Þrað2Þr � � �raðKÞ. We define the potential

gap of a multiclass classifier f given in (7) on an observation x by

GðxÞ ¼Ff ðxÞðxÞ�FðK�1ÞðxÞ, ð10Þ

which is the difference between the largest class potential and thesecond largest class potential at x. It should be clear that GðxÞZ0.The following theorem demonstrates that under the conditions of

Theorem 2 (class conditional densities are band limited), theperformance of a sample potential classifier (9) is closely relatedto the potential gap.

Theorem 3. Let S ¼ fðx1,z1Þ, . . . ,ðx‘ ,z‘Þg �Rd�K be a random i.i.d.

sample from F, a mixture of K distributions F1, . . . ,FK : F ¼PKk ¼ 1 PkFk, where Pk is the marginal probability of class k; Fk,

defined by a density function pkðxÞ, is the distribution of x for class

k. The conditional density functions are band limited, i.e., there exists

s40 such that pkðxÞ ¼ 0 when JxJZs for all k¼ 1, . . . ,K , where

pkðxÞ is the Fourier transform of pkðxÞ. For any xARd the following

inequality holds:

Pr½fSðxÞa f nðxÞ�r2Ke�‘GðxÞ2=2ð2sÞ2d

, ð11Þ

where fSðxÞ is the sample potential function classifier given in

Definition 2 with (5) being the point potential function, f nðxÞ the

Bayes classifier, and GðxÞ the potential gap.

A proof of Theorem 3 is given in the Appendix.Theorem 3 suggests that for any given x and a band limited

joint probability density function, the probability that the samplepotential function classifier behaves differently from the Bayesclassifier depends on two parameters: the potential gap GðxÞ andthe sample size ‘. The larger the potential gap and the samplesize, the more likely that the sample potential function classifiermakes the optimal prediction. In this sense, the generalizationperformance of fS depends on the potential gap. Nevertheless,Theorem 3 does not tell us how to pick a sample size ‘, neithercould we compute the right hand side of the inequality (11),because the potential gap is unknown in practice. Motivated bythe potential gap, we present, in the next section, a probabilisticbound on the generalization performance of a sample potentialfunction classifier based on the margin of fS , which is closelyrelated to the sample version of the potential gap.

4. A generalization bound for potential function classifiers

As indicated in Definition 1, the sample class potential fkðx,SÞis an estimate of the class potential FkðxÞ. Analogous to thepotential gap, we define the margin of fS on an observationðx,zÞARd

�K as

gðx,z,SÞ ¼fzðx,SÞ�fðK�1Þðx,SÞ: ð12Þ

Given a classifier fS and a desired margin a40, we denote by xthe bounded amount by which fS fails to achieve the desiredmargin a on sample (x, z),

x¼minfa,½a�gðx,z,SÞ�þ g,

where ½x�þ ¼ x if xZ0 and 0 otherwise. For an observationðxi,ziÞAS, we define its margin shortage, xi, as

xi ¼minfa,½a�gðxi,zi,SðiÞÞ�þ g, ð13Þ

where SðiÞ ¼ S�fðxi,ziÞg. Note that both x and xiA ½0,a�.We illustrate the concepts of margin and x in Fig. 1 under a

3-class scenario. The solid curve in Fig. 1(a) shows the variationsof the margin for an observation, (x,3), as a function of its sampleclass potential f3ðx,SÞ. The sample class potentials of x withrespect to class 1 and class 2, i.e., f1ðx,SÞ and f2ðx,SÞ, are fixed.For a desired margin a¼ 0:3, the dashed curve represents thevalue of x: the bounded amount by which the margin is lessthan a. Fig. 1(b) shows three sample class potential functionsconstructed from 12 training observations. Each class is asso-ciated with a distinct marker: circle, triangle, or square. The pointpotential function defined in (5) with s¼0.1 is used in theevaluation of the sample class potential functions. We visualizethe margin for each training observation using an arrow where

Page 6: Multiclass classification with potential function rules: Margin distribution and generalization

−0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8−0.3

−0.2

−0.1

0

0.1

0.2

0.3

0.4

0.5

φ3 (x, S)

Mar

gin

and

ξ (α

= 0

.3)

margin γ (x, 3, S)ξ

φ2(x, S)φ1(x, S)

α

α

−5 −4 −3 −2 −1 0 1 2 3 4 5−0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

x

Sam

ple

Cla

ss P

oten

tion,

φk (

x)

Class 1Class 2Class 3

0.0296

−0.0142

0.0024

0.05270.0183

0.0206

0.05010.0452

0.012

−0.0181

0.0

0.0

Fig. 1. Sample class potential functions and margins under a 3-class scenario. (a) The solid curve describes the variation of margin gðx,3,SÞ with respect to the sample class

potential f3ðx,SÞ when the sample class potential f1ðx,SÞ and f2ðx,SÞ are fixed. The dashed curve represents x, the bounded amount by which the margin is less than

a¼ 0:3. (b) The three curves represent sample class potential functions built upon 12 training observations (denoted by the markers on the horizontal axis) using a 1-d sinc

point potential function with s¼0.1. Each arrow corresponds to a margin, which is computed as the difference between the vertical coordinate of the tip of the arrow and

that of the end of the arrow. The numeric value of the margin is given along with the arrow. The arrow is absent if the margin is 0.

F. Teng et al. / Pattern Recognition 45 (2012) 540–551 545

the margin is computed as the difference between the verticalcoordinate of the tip of the arrow ðfzðx,SÞÞ and that of the end ofthe arrow ðfð2Þðx,SÞÞ. The numerical value of a margin is also listedalong with the arrow. For observations (�0.5,2) and (4,3), thearrows are absent because their margins are 0.

It is not difficult to relate margins to classification errors.Positive margins suggest correct classifications. Negative marginsimply mis-classifications. There are only two scenarios that resultin the 0 margin: fzðx,SÞ ¼fðKÞðx,SÞ ¼fðK�1Þðx,SÞ or fðKÞðx,SÞ4fzðx,SÞ ¼ fðK�1Þðx,SÞ. In the former case, which is rare in practice,the correctness of the classification depends on the tie breakingstrategy, which is usually random. The second case is morecommon, for example the two 0 margins in Fig. 1(b). It leads tomis-classifications. If we introduce the following indicator func-tion:

Iðx,z,SÞ ¼1 gðx,z,SÞr0,

0 otherwise,

(ð14Þ

P‘i ¼ 1 Iðxi,zi,SðiÞÞ is an upper bound on the number of mis-

classified observations in a leave-one-out evaluation.The connection between the bounded margin shortage xi,

which is defined in (13), and a classification error is more subtle.If we divide xi by a, we have

xi

a ¼

1, gðxi,zi,SðiÞÞr0,

1�gðxi,zi,SðiÞÞ

a, 0ogðxi,zi,SðiÞÞra,

0, 0ogðxi,zi,SðiÞÞ:

8>>><>>>:

ð15Þ

Fig. 2(a) compares xi=a with Iðxi,zi,SðiÞÞ as a function ofgðxi,zi,SðiÞÞ. It is clear that xi=a is always greater than or equalto Iðxi,zi,SðiÞÞ. Therefore,

P‘i ¼ 1 xi=a is an upper bound on the

number of mis-classified observations in a leave-one-out evalua-tion. Next, we present a generalization bound based on thedesired margin a and the bounded margin shortage xi for anygiven bounded point potential function c. Without loss of gen-erality, we assume that c : X�X-½0,1�.

Theorem 4. Let S ¼ fðx1,z1Þ, . . . ,ðx‘ ,z‘Þg �X�K be a random i.i.d.

sample from an unknown distribution F, and fS : X-K a sample

potential function classifier defined according to (9) using a given

point potential function c : X�X-½0,1�. For a fixed dAð0,1Þ, a

desired margin a40, and a new random sample (x,z) generated from

F, the following bound holds with probability at least 1�d over S:

PrF ½za fSðxÞjS�r1

X‘i ¼ 1

xi

a þ2

‘a þ 1þ4

a

� � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffilnð2=dÞ

2‘

r, ð16Þ

where xi is defined in (13).

A proof of Theorem 4 is given in the Appendix. It is worthwhileto note that there are two sources of randomness in the aboveinequality: the random sample S and the random observation(x,z). For a specific S, the above bound is either true of false, i.e., itis not random. For a random sample S, the probability that thebound is true is at least 1�d. The inequality shows that the errorprobability, PrF ½za fSðxÞjS�, of a sample potential function classi-fier depends on three terms. The first term, ð1=‘Þ

P‘i ¼ 1 xi=a, is an

upper bound on the leave-one-out training error. The second andthe third terms are determined by the training sample size ‘, thedesired margin a, and the confidence parameter d. In general, forfixed ‘ and d, the generalization performance of fS is a tradeoffbetween training error and the desired margin a. On one hand, asmaller a produces a tighter bound on the training error, butlarger values for the second and the third terms. On the otherhand, a larger a can reduce the values of the second and the thirdterms, but makes the first term a looser bound on the trainingerror. This is illustrated in Fig. 2(b) using margins generated froma uniform distribution on [�0.1, 1]. The values of the upperbound are shown as a function of the desired margin a. In thenext section, we discuss classifier selection methods motivated bythe above bound on the generalization performance.

5. Margin distributions and classifier selection

The learning of a potential function classifier is essentially theselection of a point potential function (or its parameters).Fig. 2(b) shows that given ‘ and d, the upper bound on theprobability of error has a minimum. Hence it is tempting tochoose a classifier that minimizes the upper bound in (16).Unfortunately, this is not an effective approach in practice

Page 7: Multiclass classification with potential function rules: Margin distribution and generalization

−0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

0

0.2

0.4

0.6

0.8

1

γ (xi , zi , S (i))

ξ i /α

I(xi, zi , S(i))

α = 0.2α = 0.5

slope = 1/α

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

α

Upp

er B

ound

on

the

Prob

abili

ty o

f Err

or

Sample Size 1,000Sample Size 10,000Sample Size 100,000

Fig. 2. (a) Plots of xi , the bounded margin shortage, as a function of the margin gðxi ,zi ,SðiÞÞ. When a approaches 0, xi=a converges to the indicator function Iðxi ,zi,SðiÞÞ.(b) Plots of the upper bound on the probability of error in (16) as a function of the desired margin a.

−0.05 0 0.05 0.1 0.15 0.2 0.250

1

2

3

4

5

6

7

Margin

log

(Pro

babi

lity

Den

sity

+ 1

)

σ=0.9487, support region = [−0.0020, 0.0368]σ=1.5811, support region = [−0.0118, 0.1221]σ=2.2136, support region = [−0.0310, 0.2068]

−0.4 −0.2 0 0.2 0.4 0.6 0.8 10

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

Normalized Margin

log

(Pro

babi

lity

Den

sity

+ 1

)

σ=0.9487σ=1.5811σ=2.2136

Fig. 3. Distributions of margin and normalized margin under a Gaussian point potential function e�Jx�yJ2=s2with different values of s. (a) Probability density of margin.

(b) Probability density of normalization margin.

F. Teng et al. / Pattern Recognition 45 (2012) 540–551546

because the bound is usually loose even for large training setswith 50,000–100,000 observations.

As we discussed in Section 4, the desired margin a plays a keyrole in estimating the generalization performance. If we define

in ¼ argmini ¼ 1,...,‘,gðxi ,zi ,SðiÞÞ40

gðxi,zi,SðiÞÞ

it is clear from Fig. 2(a) that ð1=‘ÞP‘

i ¼ 1 xi=a achieves the mini-mum (which is equal to the training error) when0oargðxin ,zin ,SðinÞÞ.1 Although a larger value of a decreases thevalues of the last two terms in (16), it also increases the value ofð1=‘Þ

P‘i ¼ 1 xi=a. However, for a fixed value of a, the bound is

tighter if the margins are concentrated more towards the positive

1 This is because there will be no observations whose margin falls into the

sloped region. Hence ð1=‘ÞP‘

i ¼ 1 xi=a¼ ð1=‘ÞP‘

i ¼ 1 Iðxi ,zi ,SðiÞÞ, which is the leave-

one-out training error.

end than towards the negative end. This suggest that we mayselect classifiers based on the distribution of margins.

However, a direct comparison of margin distributions may notbe meaningful because the support region of a margin distribu-tion largely depends on the selected point potential function cand its parameters. For example, Fig. 3(a) shows the probabilitydistributions of margins under a Gaussian point potential func-tion (i.e., cðx,yÞ ¼ e�Jx�yJ2=s2

Þ using the MAGIC dataset from UCIMachine Learning Repository (details of the dataset are given inSection 6). The support region of the margin distribution variessignificantly with the values of s.

To make margins comparable under different point potentialfunctions or different parameter values, we propose the followingnormalization procedure. For any given xARd, we define anormalized sample class potential, fkðx,SÞ, as

fkðx,SÞ ¼ fkðx,SÞPKi ¼ 1 jfiðx,SÞj

:

Page 8: Multiclass classification with potential function rules: Margin distribution and generalization

Table 1The comparison results of model selection using leave-one-out error and the

margin distribution metric defined in (17). ‘train: the size of training set; ‘test: the

size of test set; d: feature dimension; K: the number of classes; nB: the number of

experiments in which the proposed method performs better than leave-one-out

model selection; nE: the number of experiments in which the proposed method

and leave-one-out model selection have same test error; nW: the number of

experiments in which the proposed method performs worse than leave-one-out

model selection.

Dataset ‘train ‘test d K nB nE nW

Balancescale 570 55 4 2 7 37 6Bloodtransfusion 600 148 4 2 29 13 8Breastcancer 600 83 9 2 10 39 1Ecoli 200 136 7 8 15 22 13Glass 150 64 9 6 22 21 7Imageseg 2100 210 19 7 15 22 13Ionosphere 320 31 34 2 5 29 16

Letter 18,000 2000 16 26 32 7 11Liver 300 45 6 2 23 14 13Magic 10,000 9020 10 2 44 4 2Multi-Feature1 1800 200 216 10 9 34 7Multi-Feature2 1800 200 64 10 7 43 0Multi-Feature3 1800 200 240 10 9 39 2Satimage 5835 600 36 6 16 23 11Sonar 150 58 60 2 15 20 15

Spectfheart 200 67 44 2 3 37 0Survival 206 100 3 2 22 8 20Vehicle 800 46 18 4 8 31 11

Vowel 890 100 10 11 3 33 14

F. Teng et al. / Pattern Recognition 45 (2012) 540–551 547

Clearly, the above normalization does not change the order ofsample class potentials, hence the classification decisions. Thenormalized margin of fS on an observation ðx,zÞARd

�K is thendefined as

gðx,z,SÞ ¼ fzðx,SÞ�fðK�1Þðx,SÞ:

Fig. 3(b) shows the probability densities of the margins afternormalization. In both figures, the densities are shown under alog transformation. As we discussed in Section 3.1, if c is anonnegative translation invariant function that is integrable overX, fkðx,SÞ is proportional to an estimation of the posteriorprobability Prðz¼ kjxÞ. The normalized class potential, fkðx,SÞ, isan estimate of the posterior probability Prðz¼ kjxÞ. Hence thenormalized margin g can be viewed as an estimation on theposterior probability gap.

In classifier selection, we would like to choose a classifierwhose margins concentrate towards the positive end. In terms ofnormalized margin, this suggests that g should concentratetowards 1. We propose the following metric:

hðfSÞ ¼ varg�meang , ð17Þ

where meang ¼ ð1=‘ÞP‘

i ¼ 1 gðxi,zi,SðiÞÞ and varg ¼ ð1=ð‘�1ÞÞP‘

i ¼ 1

½gðxi,zi,SðiÞÞ�meang �2. Clearly, the desired normalized margins

should have large mean and small variance, i.e., we select aclassifier that minimizes h.

Winequality 6000 497 11 7 17 0 33

6. Experimental results

We compare the proposed model selection method usingnormalized margin distribution with a traditional approach usingthe leave-one-out training error. The experiments wereconducted on 20 datasets from UCI Machine Learning Repository.Each dataset is randomly divided into a training set and a test set.We built a potential function classifier with Gaussian pointpotential function for each dataset. The bandwidth parameter sof the point potential function is determined from 20 differentvalues (0.03, 0.04, 0.05, 0.06, 0.07, 0.08, 0.09, 0.1, 0.2, 0.3, 0.4, 0.5,0.6, 0.7, 0.8, 0.9, 1.0, 2.0, 3.0, and 4.0), using two strategies:(1) minimizing the leave-one-out training error; (2) minimizingthe margin distribution metric defined in (17). The above proce-dure was repeated for 50 runs. In each run, test errors wererecorded. In Table 1, we list the names of the datasets, the sizes oftraining and test sets, the dimension of the feature space, thenumber of categories, and the number of runs in which theproposed model selection method outperformed (nB), tied with(nE), and underperformed (nW) the leave-one-out approach.Among the 20 datasets, the proposed method outperformed theleave-one-out model selection on 15 datasets (i.e., nB4nW Þ,which are highlighted in Table 1. The two approaches are tiedon 1 dataset (Sonar). This suggests a very competitive perfor-mance of the proposed method.

7. Conclusions

In this paper, we revisited potential function rules (PFRs) intheir original form and reveal their connections with other well-known results in the literature. We derive a bound on thegeneralization performance of potential function classifiers basedon the observed margin distribution of the training data. A newmodel selection criterion using a normalized margin distributionis then proposed to learn ‘‘good’’ potential function classifiers inpractice. We evaluated the proposed model selection method over20 UCI datasets. In comparison with the traditional model

selection using leave-one-out training error, the margin distribu-tion based metric demonstrates very competitive performance.

Acknowledgments

This work was supported by the US National Science Founda-tion under award numbers MCB-1027989 and EPS-0903787.

Appendix A

Proof of Theorem 1. Because dð�Þ is a Dirac delta function, itfollows thatZX

pðyjþÞdðx�yÞdy¼ pðxjþÞ,

ZX

pðyj�Þdðx�yÞdy¼ pðxj�Þ:

Therefore,

PrðþÞ

ZX

pðyjþÞdðx�yÞ dypPrðþjxÞ,

Prð�Þ

ZX

pðyj�Þdðx�yÞ dypPrð�jxÞ,

i.e., the potential of the positive (negative) class is proportional tothe posterior probability of the positive (negative) class. Hencethe decision boundary of (4) is identical to that of the Bayesclassifier. &

Proof of Theorem 2. For a translation invariant c,ZX

pðyjþÞcðx,yÞ dy¼ZX

pðyjþÞcðx�yÞ dy¼F�1½pþ ðxÞCðxÞ�,

where F�1 is the inverse Fourier transform. Because pþ ðxÞ ¼ 0for JxJZs and CðxÞ ¼ 1 for JxJrs, we have pþ ðxÞCðxÞ ¼pþ ðxÞ. It follows that

PrðþÞ

ZX

pðyjþÞcðx,yÞ dy¼ PrðþÞpðxjþÞpPrðþjxÞ:

Page 9: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551548

Similarly,

Prð�Þ

ZX

pðyj�Þcðx,yÞ dy¼ Prð�Þpðxj�ÞpPrð�jxÞ:

The potential of the positive (negative) class hence is proportionalto the posterior probability of the positive (negative) class.Therefore, the decision boundary of (4) is identical to that of theBayes classifier. &

We need the following lemma to prove Theorem 3.

Lemma 1. For any a1,a2, . . . ,aK AR and b1,b2, . . . ,bK AR, if

jak�bkjre for all kAK, we have jaðjÞ�bðjÞjre for all jAK.

Proof. For any jAK,

aðjÞ�eraðjþ1Þ�er � � �raðKÞ�e:

Because bkZak�e for all kAK, the number of bk’s that are greaterthan or equal to aðjÞ�e is at least K�jþ1. Therefore, bðjÞZaðjÞ�e.Similarly, for any jAK,

að1Þ þerað2Þ þer � � �raðjÞ þe:

Because bkrakþe for all kAK, the number of bk’s that are lessthan or equal to aðjÞ þe is at least j. Therefore, bðjÞraðjÞ þe. Thiscompletes the proof. &

Proof of Theorem 3. We introduce a new random variableSk ¼ jfðx,zÞAS : z¼ kgj. For any given x,

EF ½fkðx,SÞ� ¼ ESkEFjSk

Sk

1

Sk

Xzi ¼ k

cðx,xiÞ

24

35

8<:

9=;

¼ ESk

Sk

‘Ey � Fk

½cðx,yÞ�

� �¼ PkEy � Fk

½cðx,yÞ� ¼FkðxÞ:

We rewrite fkðx,SÞ as fkðx,SÞ ¼ ð1=‘ÞP‘

i ¼ 1 Iðzi ¼ kÞcðx,xiÞ wherethe indicator function Iðzi ¼ kÞ ¼ 1 if zi ¼ k, Iðzi ¼ kÞ ¼ 0 otherwise.Because ðxi,ziÞ’s are i.i.d., so are Iðzi ¼ kÞcðx,xiÞ. In addition, from(5) it is clear that jIðzi ¼ kÞcðx,xiÞjr ð2sÞd. It follows from Hoeffd-ing’s inequality that for any given x, e40, and k¼ 1, . . . ,K ,

Pr½jfkðx,SÞ�FkðxÞjZe�r2e�2‘e2=ð2sÞ2d

: ðA:1Þ

Because the conditional densities are band limited, it follows

from the proof of Theorem 2 that

FkðxÞ ¼ PkEy � Fk½cðx,yÞ�pPrðz¼ kjxÞ:

Hence we have f nðxÞ ¼ argmaxkFkðxÞ. From Lemma 1, we know

that if jfkðx,SÞ�FkðxÞjrGðxÞ=2 for k¼ 1, . . . ,K , jfðKÞðx,SÞ�FðKÞðxÞjr GðxÞ=2. Combining this with the facts that

ffS ðxÞðx,SÞ ¼fðKÞðx,SÞ and Ff nðxÞðxÞ ¼FðKÞðxÞ

it is straightforward to derive that fSðxÞ ¼ f nðxÞ. Therefore,

Pr jfkðx,SÞ�FkðxÞjoGðxÞ

2,8k¼ 1, . . . ,K

rPr½fSðxÞ ¼ f nðxÞ�: ðA:2Þ

Let e¼GðxÞ=2. Using (A.1), (A.2), and the union bound, we have

Pr½fSðxÞa f nðxÞ�rPr (k,jfkðx,SÞ�FkðxÞjZGðxÞ

2

r2Ke�‘GðxÞ

2=2ð2sÞ2d

:

This completes the proof. &

In order to prove the upper bound on generalization of samplepotential function classifiers in Theorem 4, we need the followinglemma and an inequality attributed to McDiarmid.

Lemma 2. Let SðiÞ ¼ S�fðxi,ziÞg. For a change of one ðxt ,ztÞ to ðxt ,ztÞ,denote

S t ¼ fðx1,z1Þ, . . . ,ðxt�1,zt�1Þ,ðxt ,ztÞ,ðxtþ1,ztþ1Þ, . . . ,ðx‘ ,z‘Þg:

We define S tðiÞ ¼ S t�fðxi,ziÞg, hence S tðtÞ ¼ SðtÞ. Let xAX be any

observation in X and zAK a class label. The following inequalities

hold for any point potential function c : X�X-½0,1�:

jgðx,z,SÞ�gðx,z,SðiÞÞjr 2

‘, ðA:3Þ

jgðx,z,SðiÞÞ�gðx,z,S tðiÞÞjr2

‘�1, ðA:4Þ

jgðxi,zi,SðiÞÞ�gðxi,zi,S tðiÞÞjr2

‘�1if iat: ðA:5Þ

Proof. It is readily checked that for any zAK,

jfzðx,SÞ�fzðx,SðiÞÞj ¼

1

Pzj ¼ z

cðx,xjÞ�1

‘�1

Pzj ¼ z,ja i

cðx,xjÞ

���������� if z¼ zi,

1

Pzj ¼ z

cðx,xjÞ�1

‘�1

Pzj ¼ z

cðx,xjÞ

���������� if zazi,

8>>>>><>>>>>:

¼

1

‘cðx,xiÞ�

1

‘ð‘�1Þ

Pzj ¼ z,ja i

cðx,xjÞ

����������r 1

‘ if z¼ zi,

1

‘ð‘�1Þ

Pzj ¼ z

cðx,xjÞ

����������r1

‘if zazi:

8>>>>><>>>>>:

From (12) we have

jgðx,z,SÞ�gðx,z,SðiÞÞj¼ jfzðx,SÞ�fðK�1Þðx,SÞ�fzðx,SðiÞÞþfðK�1Þðx,SðiÞÞj

r1

‘þjfðK�1Þðx,SÞ�fðK�1Þðx,SðiÞÞjr 2

‘,

where the last step is based on Lemma 1.

It is not difficult to show that for any zAK, jfzðx,SðiÞÞ�fzðx,S tðiÞÞj ¼ 0 when i¼t, otherwise,

jfzðx,SðiÞÞ�fzðx,S tðiÞÞj

¼

0 if zt az, zt az

1

‘�1cðx,xtÞ

�������� if zt az, zt ¼ z

1

‘�1cðx,xtÞ�

1

‘�1cðx,xtÞ

�������� if zt ¼ z, zt ¼ z

1

‘�1cðx,xtÞ

�������� if zt ¼ z, zt az

8>>>>>>>>>><>>>>>>>>>>:

r1

‘�1:

Therefore,

jgðx,z,SðiÞÞ�gðx,z,S tðiÞÞj

r jfzðx,SðiÞÞ�fzðx,S tðiÞÞjþjfðK�1Þðx,SðiÞÞ�fðK�1Þðx,S tðiÞÞj

r2

‘�1:

Finally, for iat and any ziAK,

jfziðxi,SðiÞÞ�fzi

ðxi,S tðiÞÞj

¼

0 if zt azi, zt azi

1

‘�1cðxi,xtÞ

�������� if zt azi, zt ¼ zi

1

‘�1cðxi,xtÞ�

1

‘�1cðxi,xtÞ

�������� if zt ¼ zi, zt ¼ zi

1

‘�1cðxi,xtÞ

�������� if zt ¼ zi, zt azi

8>>>>>>>>>><>>>>>>>>>>:

r1

‘�1:

Page 10: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551 549

Therefore,

jgðxi,zi,SðiÞÞ�gðxi,zi,S tðiÞÞj

r jfziðxi,SðiÞÞ�fzi

ðxi,S tðiÞÞjþjfðK�1Þðxi,SðiÞÞ�fðK�1Þðxi,S tðiÞÞj

r2

‘�1:

This completes the proof. &

Lemma 3 (McDiarmid’s inequality). Let X1,X2, . . . ,Xn be indepen-

dent random variables taking values in a set X. Suppose that f :Xn-R satisfies

supx1 ,...,xn ,x j AX

jf ðx1, . . . ,xnÞ�f ðx1, . . . ,xj, . . . ,xnÞjrcj

for constants cj,1r jrn. Then for every e40,

Pr½f ðX1, . . . ,XnÞ�Ef Ze�rexp�2e2Pnj ¼ 1 c2

j

!:

Proof of Theorem 4. Consider the loss function

gðx,z,SÞ ¼

1 if gðx,z,SÞr0,

a�gðx,z,SÞa if 0ogðx,z,SÞra,

0 otherwise:

8>><>>:

It is not difficult to show that

PrF ½za fSðxÞjS�rEFjS ½gðx,z,SÞ�,

where the equality holds when a¼ 0. Hence it suffices to showthat EFjS ½gðx,z,SÞ� is bounded by the right side of (16).

We break EFjS ½gðx,z,SÞ��ð1=‘aÞP‘

i ¼ 1 xi ¼ EFjS ½gðx,z,SÞ��ð1=‘ÞP‘i ¼ 1 gðxi,zi,SðiÞÞ into AþBþC:

A¼ EFjS ½gðx,z,SÞ��EFjS1

X‘i ¼ 1

gðx,z,SðiÞÞ" #

,

B¼ EFjS1

X‘i ¼ 1

gðx,z,SðiÞÞ" #

�EF ½gðxj,zj,SðjÞÞ�,

C ¼ EF ½gðxj,zj,SðjÞÞ��1

X‘i ¼ 1

gðxi,zi,SðiÞÞ,

where ðxj,zjÞ is any fixed sample in S.

We first look at A. It is straightforward to show that

jgðx,z,SÞ�gðx,z,SðiÞÞjr 1

a jgðx,z,SÞ�gðx,z,SðiÞÞjr 2

‘a ,

where the last inequality is based on (A.3). Therefore,

A¼ EFjS1

X‘i ¼ 1

½gðx,z,SÞ�gðx,z,SðiÞÞ�( )

rEFjS1

X‘i ¼ 1

½gðx,z,SÞ�gðx,z,SðiÞÞ������

�����r 2

‘a: ðA:6Þ

Next, we look at B. It is not difficult to verify that

EF EFjS1

X‘i ¼ 1

gðx,z,SðiÞÞ" #( )

¼ EF ½gðxj,zj,SðjÞÞ�:

For a change of one ðxt ,ztÞ to ðxt ,ztÞ, we denote

S t ¼ fðx1,z1Þ, . . . ,ðxt�1,zt�1Þ,ðxt ,ztÞ,ðxtþ1,ztþ1Þ, . . . ,ðx‘ ,z‘Þg:

From (A.4) we have for any zAK

jgðx,z,SðiÞÞ�gðx,z,S tðiÞÞjr1

ajgðx,z,SðiÞÞ�gðx,z,S tðiÞÞjr

2

að‘�1Þ:

Therefore,

supðx1 ,z1Þ,...,ðx‘ ,z‘Þ,ðx t ,z tÞ

EFjS1

X‘i ¼ 1

gðx,z,SðiÞÞ" #

�EFjS t

1

X‘i ¼ 1

gðx,z,S tðiÞÞ

" #����������

¼ supðx1 ,z1Þ,...,ðx‘ ,z‘ Þ,ðx t ,z t Þ

1

X‘i ¼ 1

EFjS,S t½gðx,z,SðiÞÞ�gðx,z,S tðiÞÞ�

����������

r supðx1 ,z1Þ,...,ðx‘ ,z‘ Þ,ðx t ,z t Þ

1

X‘i ¼ 1

EFjS,S tjgðx,z,SðiÞÞ�gðx,z,S tðiÞÞj

r2

a‘: ðA:7Þ

By (A.7), we apply the McDiamid’s inequality to get

PrðB4e1Þrexp�a2‘e2

1

2

� �: ðA:8Þ

Next, we look at C. It is clear that

EF1

X‘i ¼ 1

gðxi,zi,SðiÞÞ" #

¼ EF ½gðxj,zj,SðjÞÞ�:

Let gðSÞ ¼ ð1=‘ÞP‘

i ¼ 1 gðxi,zi,SðiÞÞ. For a change of one ðxt ,ztÞ to

ðxt ,ztÞ, denote

S t ¼ fðx1,z1Þ, . . . ,ðxt�1,zt�1Þ,ðxt ,ztÞ,ðxtþ1,ztþ1Þ, . . . ,ðx‘ ,z‘Þg:

For any iat, it follows from (A.5) that for any ziAK

jgðxi,zi,SðiÞÞ�gðxi,zi,S tðiÞÞjr1

a jgðxi,zi,SðiÞÞ�gðxi,zi,S tðiÞÞjr2

að‘�1Þ:

Therefore,

supðx1 ,z1Þ,...,ðx‘ ,z‘Þ,ðx t ,z tÞ

jgðSÞ�gðSt Þj

¼ supðx1 ,z1Þ,...,ðx‘ ,z‘ Þ,ðx t ,z t Þ

1

X‘i ¼ 1

gðxi,zi,SðiÞÞ� gðxt ,zt ,S tðtÞÞh�����

þX‘

i ¼ 1,ia t

gðxi,zi,S tðiÞÞ

#�����r

1

‘þ supðx1 ,z1Þ,...,ðx‘ ,z‘ Þ,ðx t ,z t Þ

1

X‘i ¼ 1,ia t

½gðxi,zi,SðiÞÞ�gðxi,zi,S tðiÞÞ�

����������

r1

‘þ

2

a‘ : ðA:9Þ

By (A.9), we apply the McDiarmid’s inequality to get

PrðC4e2Þrexp�2‘e2

2

1þ2

a

� �2

0BBB@

1CCCA: ðA:10Þ

Finally, setting

exp�g2‘e2

1

2

� �¼ exp

�2‘e22

1þ2

a

� �2

0BBB@

1CCCA¼ d

2

and solving for e1 and e2, we obtain

e1 ¼1

a

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2lnð2=dÞ

r, e2 ¼

1

1

a

� � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2lnð2=dÞ

r:

Because BþC4e1þe2 implies B4e1 or C4e2,

PrðBþC4e1þe2ÞrPrðB4e1 or C4e2ÞrPrðB4e1ÞþPrðC4e2Þrd:

So, with probability at least 1�d, BþCre1þe2. Because Ar2=‘a,

BþCre1þe2 implies that AþBþCre1þe2þ2=‘a. Therefore,

Page 11: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551550

with probability at least 1�d, AþBþCre1þe2þ2=‘a, i.e.,

EFjS ½gðx,z,SÞ�r 1

X‘i ¼ 1

gðxi,zi,SðiÞÞþ2

‘a þ 1þ4

a

� � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffilnð2=dÞ

2‘

r:

It is easy to verify that ð1=‘ÞP‘

i ¼ 1 gðxi,zi,SðiÞÞ ¼ ð1=‘ÞP‘

i ¼ 1 xi=a.

Therefore, with probability at least 1�d,

EFjS ½gðx,z,SÞ�r 1

X‘i ¼ 1

xi

a þ2

‘a þ 1þ4

a

� � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffilnð2=dÞ

2‘

r:

This completes the proof. &

References

[1] M.A. Aizerman, E.M. Braverman, L.I. Rozonoer, Theoretical fundations of thepotential function method in pattern recognition learning, Automation andRemote Control 25 (6) (1964) 917–936.

[2] M.A. Aizerman, E.M. Braverman, L.I. Rozonoer, The probability problem ofpattern recognition learning and the method of potential functions, Automa-tion and Remote Control 25 (9) (1964) 1307–1323.

[3] M.A. Aizerman, E.M. Braverman, L.I. Rozonoer, The method of potentialfunctions for the problem of restoring the characteristic of a functionconverter from randomly observed points, Automation and Remote Control25 (12) (1964) 1705–1714.

[4] M.A. Aizerman, E.M. Braverman, L.I. Rozonoer, Extrapolative problems inautomatic control and the method of potential functions, American Mathe-matical Society Translations 87 (2) (1970) 281–303.

[5] M. Anthony, N. Biggs, Computational Learning Theory, Cambridge UniversityPress, 1992.

[6] H. Avi-Itzhak, T. Diep, Arbitrarily tight upper and lower bounds on theBayesian probability of error, IEEE Transactions on Pattern Analysis andMachine Intelligence 18 (1) (1996) 89–91.

[7] K. Barnard, P. Duygulu, D. Forsyth, N. de Freitas, D.M. Blei, M.I. Jordan,Matching words and pictures, Journal of Machine Learning Research 3 (2003)1107–1135.

[8] A. Barron, Complexity regularization with application to artificial neuralnetworks, in: Nonparametric Functional Estimation and Related Topics,Kluwer Academic Publisher, 1991, pp. 561–576.

[9] P.L. Bartlett, For valid generalization, the size of the weights is moreimportant than the size of the network, Advances in Neural InformationProcessing Systems 9 (1997) 134–140.

[10] O.A. Bashkirov, E.M. Braverman, I.B. Muchnik, Potential function algorithmsfor pattern recognition learning machines, Automation and Remote Control25 (5) (1964) 692–695.

[11] T. Bayes, An essay towards solving a problem in the doctrine of chances, ThePhilosophical Transactions 53 (1763) 370–418.

[12] M. Ben-Bassat, K.L. Klove, M.H. Weil, Sensitivity analysis in Bayesianclassification models: multiplicative deviations, IEEE Transactions on PatternAnalysis and Machine Intelligence 2 (3) (1980) 261–266.

[13] J.O. Berger, Statistical Decision Theory and Bayesian Analysis, second ed.,Springer, 1985.

[14] C.M. Bishop, Pattern Recognition and Machine Learning, Springer, 2006.[15] M. Boulle, Compression-based averaging of selective naive Bayes classifiers,

Journal of Machine Learning Research 8 (2007) 1659–1685.[16] O. Bousquet, A. Elisseeff, Stability and generalization, Journal of Machine

Learning Research 2 (2002) 499–526.[17] E.M. Braverman, On the method of potential functions, Automation and

Remote Control 26 (12) (1965) 2205–2213.[18] E.M. Braverman, E.S. Pyatnitskii, Estimation of the rate of convergence of

algorithms based on the potential functions method, Automation andRemote Control 27 (1) (1966) 95–112.

[19] P. Bruneau, M. Gelgon, F. Picarougne, Parsimonious reduction of Gaussianmixture models with a variational-Bayes approach, Pattern Recognition 43(3) (2010) 850–858.

[20] Y. Chen, J.Z. Wang, Support vector learning for fuzzy rule-based classificationsystems, IEEE Transactions on Fuzzy Systems 11 (6) (2003) 716–728.

[21] Y. Chen, J. Bi, J.Z. Wang, MILES: multiple-instance learning via embeddedinstance selection, IEEE Transactions on Pattern Analysis and MachineIntelligence 28 (12) (2006) 1931–1947.

[22] Y. Chen, E.K. Garcia, M.R. Gupta, A. Rahimi, L. Cazzanti, Similarity-basedclassification: concepts and algorithms, Journal of Machine LearningResearch 10 (2009) 747–776.

[23] K.B. Davis, Mean square error properties of density estimates, The Annals ofStatistics 3 (4) (1975) 1025–1030.

[24] K.B. Davis, Mean integrated square error properties of density estimates, TheAnnals of Statistics 5 (3) (1977) 530–535.

[25] P.A. Devijver, On a new class of bounds on Bayes risk in multi-hypothesispattern recognition, IEEE Transactions on Computers 23 (1) (1974) 70–80.

[26] L. Devroye, On the asymptotic probability of error in nonparametric dis-crimination, The Annals of Statistics 9 (6) (1981) 1320–1327.

[27] L. Devroye, The equivalence of weak, strong and complete convergencein L1 for kernel density estimates, The Annals of Statics 11 (3) (1983)896–904.

[28] L. Devroye, Asymptotic performance bounds for the kernel estimate, TheAnnals of Statistics 16 (3) (1988) 1162–1179.

[29] L. Devroye, A universal lower bound for the kernel estimate, Statistics andProbability Letters 8 (1989) 419–423.

[30] L. Devroye, L. Gyorfi, G. Lugosi, A Probabilistic Theory of Pattern Recognition,Springer-Verlag, New York, 1996.

[31] L. Devroye, L. Gyorfi, A. Krzyzak, The Hilbert kernel regression estimate,Journal of Multivariate Analysis 65 (1998) 209–227.

[32] L. Devroye, A. Krzyzak, On the Hilbert kernel density estimate, Statistics andProbability Letters 44 (1999) 299–308.

[33] L. Devroye, A. Krzyzak, New multivariate product density estimator, Journalof Multivariate Analysis 82 (2002) 88–110.

[34] P. Domingos, M.J. Pazzani, On the optimality of the simple Bayesian classifierunder zero–one loss, Machine Learning 29 (2–3) (1997) 103–130.

[35] R.O. Duda, P.E. Hart, D.G. Stork, Pattern Classification, second ed., John Wiley& Sons Inc., 2001.

[36] A. Garg, D. Roth, Margin distribution and learning algorithms, in: Proceedingsof the Twentieth International Conference on Machine Learning, 2003,pp. 210–217.

[37] L. Gordon, R.A. Olshen, Asymptotically efficient solutions to the classificationproblem, The Annals of Statistics 6 (3) (1978) 515–533.

[38] D.J. Griffiths, Introduction to Electrodynamics, third ed., Prentice Hall,1998.

[39] Y. Guermeur, VC theory of large margin multi-category classifiers, Journal ofMachine Learning Research 8 (2007) 2551–2594.

[40] A. Halevy, P. Norvig, F. Pereira, The unreasonable effectiveness of data, IEEEIntelligent Systems 24 (2) (2009) 8–12.

[41] W.A. Hashlamoun, P.K. Varshney, V.N.S. Samarasooriya, A tight upper boundon the Bayesian probability of error, IEEE Transactions on Pattern Analysisand Machine Intelligence 16 (2) (1994) 220–224.

[42] T. Hastie, R. Tibshirani, J. Friedman, The Elements of Statistical Learning: DataMining, Inference, and Prediction, Springer, 2001.

[43] T. Hofmann, J. Puzicha, M.I. Jordan, Unsupervised learning from dyadic data,Advances in Neural Information Processing Systems 11 (1999) 466–472.

[44] X.-B. Jin, C.-L. Liu, X. Hou, Regularized margin-based conditional log-like-lihood loss for prototype learning, Pattern Recognition 43 (7) (2010)428–2438.

[45] M.J. Kearns, U.V. Vazirani, An Introduction to Computational Learning Theory,The MIT Press, 1994.

[46] H.-C. Kim, Z. Ghahramani, Bayesian Gaussian process classification with theEM–EP algorithm, IEEE Transactions on Pattern Analysis and MachineIntelligence 28 (12) (2006) 1948–1959.

[47] M. Kim, Large margin cost-sensitive learning of conditional random fields,Pattern Recognition 43 (10) (2010) 3683–3692.

[48] J. Langford, J. Shawe-Taylor, PAC-Bayes and margins, Advances in NeuralInformation Processing Systems 15 (2002) 439–446.

[49] P. Langley, W. Iba, K. Thompson, An analysis of Bayesian classifiers, in:Proceedings of the Tenth National Conference on Artificial Intelligence, 1992,pp. 223–228.

[50] H. Langseth, T.D. Nielsen, Latent classification models for binary data, PatternRecognition 42 (11) (2009) 2724–2736.

[51] G. Lugosi, K. Zeger, Concept learning using complexity regularization, IEEETransactions on Information Theory 42 (1) (1996) 48–54.

[52] A. Maurer, Learning similarity with operator-valued large-margin classifiers,Journal of Machine Learning Research 9 (2008) 1049–1082.

[53] T.M. Mitchell, Machine Learning, McGraw-Hill Companies Inc., 1997.[54] E. Pekalska, P. Paclik, R.P.W. Duin, A generalized kernel approach to dissim-

ilarity-based classification, Journal of Machine Learning Research 2 (2001)175–211.

[55] G. Ratsch, M.K. Warmuth, Efficient margin maximizing with boosting, Journalof Machine Learning Research 6 (2005) 2131–2152.

[56] S. Rosset, J. Zhu, T. Hastie, Boosting as a regularized path to a maximummargin classifier, Journal of Machine Learning Research 5 (2004) 941–973.

[57] R.E. Schapire, Y. Freund, P. Bartlett, W.S. Lee, Boosting the margin: a newexplanation for the effectiveness of voting methods, The Annals of Statistics26 (5) (1998) 1651–1686.

[58] B. Scholkopf, A.J. Smola, Learning with Kernels: Support Vector Machines,Regularization, Optimization, and Beyond, The MIT Press, 2002.

[59] J. Shawe-Taylor, N. Cristianini, Kernel Methods for Pattern Analysis, Cam-bridge University Press, 2004.

[60] C.J. Stone, Consistent nonparametric regression, The Annals of Statistics 5 (4)(1977) 595–620.

[61] J. Sung, Z. Ghahramani, S.-Y. Bang, Latent-space variational Bayes, IEEETransactions on Pattern Analysis and Machine Intelligence 30 (12) (2008)2236–2242.

[62] R. Tibshirani, T. Hastie, Margin trees for high-dimensional classification,Journal of Machine Learning Research 8 (2007) 637–652.

[63] V.N. Vapnik, A.Ya. Chervonenkis, On the uniform convergence of relativefrequencies of events to their probabilities, Theory of Probabilities and ItsApplications 16 (2) (1971) 264–280.

[64] V.N. Vapnik, Estimation of Dependencies Based on Empirical Data, Springer-Verlag, 1982.

[65] V.N. Vapnik, Statistical Learning Theory, John Wiley & Sons Inc., 1998.

Page 12: Multiclass classification with potential function rules: Margin distribution and generalization

F. Teng et al. / Pattern Recognition 45 (2012) 540–551 551

[66] S. Veeramachaneni, G. Nagy, Analytical results on style-constrained Bayesianclassification of pattern fields, IEEE Transactions on Pattern Analysis andMachine Intelligence 29 (7) (2007) 1280–1285.

[67] J. Wang, X. Shen, Large margin semi-supervised learning, Journal of MachineLearning Research 8 (2007) 1867–1891.

[68] G.S. Watson, M.R. Leadbetter, On the estimation of the probability density I,The Annals of Mathematical Statistics 34 (2) (1963) 480–491.

[69] K.Q. Weinberger, L.K. Saul, Distance metric learning for large margin nearestneighbor classification, Journal of Machine Learning Research 10 (2009)207–244.

Fei Teng received B.S. degree in computer science from the Beijing University of Posts and Telecommunications and M.S. degree in computer science from the University ofNew Orleans. He is a Ph.D. candidate in computer and information science, The University of Mississippi. His research interests include pattern recognition and brian–computer interface.

Yixin Chen received B.S. degree (1995) from the Department of Automation, Beijing Polytechnic University, the M.S. degree (1998) in control theory and application fromTsinghua University, and the M.S. (1999) and Ph.D. (2001) degrees in electrical engineering from the University of Wyoming. In 2003, he received the Ph.D. degree incomputer science from The Pennsylvania State University. He had been an Assistant Professor of computer science at the University of New Orleans. He is now an AssociateProfessor at the Department of Computer and Information Science, The University of Mississippi. His research interests include machine learning, data mining, computervision, bioinformatics, and robotics and control. Dr. Chen is a member of the ACM, the IEEE, the IEEE Computer Society, the IEEE Neural Networks Society, and the IEEERobotics and Automation Society.

Xin Dang received Ph.D. degree in statistics from the University of Texas at Dallas, working with Robert Serfling on statistical depth functions. Currently she is an AssistantProfessor of the Department of Mathematics at the University of Mississippi, joined as the faculty in 2005. Her research interests include robust and nonparametricstatistics, statistical and numerical computing, and multivariate data analysis. In particular, she has focused on data depth and application, machine learning, and robustprocedure computation. Dr. Dang is a member of the American Statistical Association and Institute of Mathematical Statistics.