Bayes Optimal Multilabel Classification via Probabilistic Classifier Chains Krzysztof Dembczy´ ski1,2 n dembczynski@informatik.uni-marburg.de Weiwei Cheng1 cheng@informatik.uni-marburg.de Eyke H¨ llermeier1 u eyke@informatik.uni-marburg.de 1 Mathematics and Computer Science, Marburg University, Hans-Meerwein-Str., 35032 Marburg, Germany 2 Institute of Computing Science, Pozna´ University of Technology, Piotrowo 2, 60-965 Pozna´, Poland n n Abstract In the realm of multilabel classification (MLC), it has become an opinio communis that optimal predictive performance can only be achieved by learners that explicitly take label dependence into account. The goal of this paper is to elaborate on this postulate in a critical way. To this end, we formalize and analyze MLC within a probabilistic setting. Thus, it becomes possible to look at the problem from the point of view of risk minimization and Bayes optimal prediction. Moreover, inspired by our probabilistic setting, we propose a new method for MLC that generalizes and outperforms another approach, called classifier chains, that was recently introduced in the literature. class labels has to be predicted simultaneously, it is arguably important to exploit these dependencies. Today, it seems to be an opinio communis that optimal predictive performance can only be achieved by methods that explicitly take label correlations into account. Many papers provide empirical evidence for this conjecture: A new method is proposed that exploits label correlations in one way or the other. Using a set of benchmark data sets, this method is then shown to outperform others in terms of different loss functions. Without questioning the value of these contributions, one may argue that this is not sufficient to gain a deeper understanding of the MLC problem. There are several reasons for this, notably the following. First, the notion of "label correlation" is often used in a purely intuitive manner, referring to a kind of non-independence, but without giving a precise formal definition. Likewise, MLC methods are often ad-hoc extensions of existing methods. Second, many studies report improvements on average, but without carefully investigating under which conditions label correlations are crucial, and when they are perhaps less important. Third, the reasons for improvements are often not carefully distinguished. As the performance of a method depends on many factors, which are hard to isolate, it is not always clear that the improvements can be fully credited to the consideration of label correlations. The goal of this paper is to provide a formal setting that allows for a more thorough analysis of MLC in general and label dependence in particular (Section 2). To this end, we distinguish two types of label dependence, conditional and unconditional, and focus our analysis on the former. We propose a probabilistic framework that suggests to look at the problem from the point of view of risk minimization and Bayes optimal prediction. Concretely, we analyze three types of 1. Introduction In contrast to conventional (single-label) classification, the setting of multilabel classification (MLC) allows an instance to belong to several classes simultaneously. At first sight, MLC problems can be solved in a quite straightforward way, namely through decomposition into several binary classification problems: One binary classifier is trained for each label and used to predict whether, for a given query instance, this label is present (relevant) or not. This approach is known as binary relevance (BR) learning. However, BR has been criticized for ignoring important information hidden in the label space, namely information about the interdependencies between the labels: Since the presence or absence of the different Appearing in Proceedings of the 27 th International Conference on Machine Learning, Haifa, Israel, 2010. Copyright 2010 by the author(s)/owner(s). Bayes-Optimal Multilabel Classification via Probabilistic Classifier Chains loss functions and, based on the results, raise the following conjecture: While considering conditional label dependence can indeed be useful for certain loss functions, there are others that are less likely to benefit (Section 3). A second important contribution of this paper is a new method for MLC, called probabilistic classifier chains (Section 4). It estimates the entire joint distribution of labels and, therefore, allows us to experimentally confirm our theoretical claims (Section 5). This method generalizes and provides a proper interpretation of the recently introduced classifier chains (Read et al., 2009). It also outperforms this algorithm, however, at the cost of an increased computational complexity. such that the labels i are simply sorted in decreasing order according to their scores fi (x). The problem of MLC can be stated as follows: Given training data in the form of a finite set of observations (x, y) X × Y, drawn independently from P(X, Y), the goal is to learn a classifier h : X Y that generalizes well beyond these observations in the sense of minimizing the expected risk with respect to a specific loss function. Before having a more detailed look at this problem in Section 3 below, let us note that the posterior probability distributions Px (Y) provide a convenient means for analyzing label dependence. A distribution of this kind informs about the probability of each label combination as well as the marginals, like in this simple example of the case m = 2: Px (Y) 0 1 (2) Px (1) 0 0.4 0.3 0.7 1 0.0 0.3 0.3 Px (1) 0.4 0.6 1 (1) 2. Multilabel Classification In this section, we describe the MLC problem in more detail and formalize it within a probabilistic setting. Along the way, we introduce the notation used throughout the paper. Let X denote an instance space, and let L = {1 , 2 , . . . , m } be a finite set of class labels. We assume that an instance x X is (non-deterministically) associated with a subset of labels L 2L ; this subset is often called the set of relevant labels, while the complement L \ L is considered as irrelevant for x. We identify a set L of relevant labels with a binary vector y = (y1 , y2 , . . . , ym ), in which yi = 1 i L. By Y = {0, 1}m we denote the set of possible labelings. We assume observations to be generated independently and randomly according to a probability distribution P(X, Y) on X × Y, i.e., an observation y = (y1 , . . . , ym ) is the realization of a corresponding random vector Y = (Y1 , Y2 , . . . , Ym ). We denote by Px (Y) = P(Y | x) the conditional distribution of (i) Y given X = x, and by Px (Yi ) = P(i) (Yi | x) the corresponding marginal distribution of Yi : P(i) (b) = x yY:yi =b In a stochastic sense, the labels are not independent if the joint conditional distribution is not the product of the marginals (like in the above example): m Px (Y) = i=1 P(i) (Yi ), x (2) and the degree of dependence could in principle be quantified in terms of measures like cross entropy or KL divergence. More specifically, we shall speak of conditional dependence in the case of (2), since the probability of Y is conditioned on the instance x. This type of (in)dependence can be distinguished from unconditional (in)dependence, which looks at the unconditional probabilities of labels, i.e., P(Y) and P(i) (Yi ) obtained, respectively by integrating Px (Y) (i) and Px (Yi ) over all x. One readily verifies that conditional does not imply unconditional dependence nor the other way around. Upon closer examination, it becomes clear that several MLC methods, including those based on the idea of stacking (Godbole & Sarawagi, 2004; Cheng & H¨llermeier, 2009), seek to exploit unconu ditional label dependence, and so do related methods from other fields, like multivariate regression in statistics (Breiman & Friedman, 1997). As will become clear later on, however, Bayes optimal prediction may require the consideration of conditional dependence. In this paper, we shall therefore focus on this type of dependence in the first place. Px (y). A multilabel classifier h is an X Y mapping that assigns a (predicted) label subset to each instance x X . Thus, the output of a classifier h is a vector h(x) = (h1 (x), h2 (x), . . . , hm (x)). Often, MLC is treated as a ranking problem, in which the labels are sorted according to the degree of relevance. Then, the prediction takes the form of the ranking or scoring function: f (x) = (f1 (x), f2 (x), . . . , fm (x)) (1) Bayes-Optimal Multilabel Classification via Probabilistic Classifier Chains 3. Risk Minimization In this section, we address the issue of optimal decision making in light of different loss functions. In general, we would like to find a risk-minimizing model h , i.e., a model that minimizes the expected loss over the joint distribution P(X, Y): R(h) = EXY L(Y, h(X)), where L(·) is a loss function on multilabel predictions. This model is given by h = arg min EXY L(Y, h(X)) h all relevant labels ideally precede all irrelevant ones, and the rank loss simply counts the number of label pairs violating this condition: Lr (y, f (x)) = (i,j):yi >yj fi < fj + 1 fi = fj 2 . (7) (3) In passing, we note that there is also a normalized variant of the rank loss, in which this number is divided by the maximum number of possible mistakes on y, i.e., by the number of summands in (7); this number is given by r(m - r)/2, with r = y1 + . . . + ym the number of relevant labels. To minimize (7), it is enough to sort the labels by their probability of relevance. Formally, we can show the following result. Theorem 3.1. A ranking function that sorts the labels according to their probability of relevance, i.e., using the scoring function f (·) with fi (x) = P(i) (1) , x minimizes the expected rank loss (7). Proof. The risk of a scoring vector f = f (x) can be written as EY | X Lr (Y, f ) = yY and determined in a pointwise way by the Bayes optimal decisions h (x) = arg min EY|X L(Y, y). y (4) In the following, we consider the problem of risk minimization for three different types of loss functions that are frequently used in the context of MLC and to some extent representative of losses used in this field. 3.1. Hamming Loss The performance in MLC is perhaps most frequently reported in terms of the Hamming loss, which is defined as the number (or, in its normalized version, the fraction) of labels whose relevance is incorrectly predicted:1 m (8) Px (y)Lr (y, f ) = fi < fj + yi >yj = yY Px (y) 1 fi = fj 2 . LH (y, h(x)) = i=1 yi = hi (x) . (5) The two sums can be swapped, and doing so yields the expression Px (y) yi > yj fi < fj + 1 fi = fj 2 For the Hamming loss (5), it is easy to see that the risk minimizer (4) is obtained by h (x) i = arg max b{0,1} 1i,jm yY which in turn can be written as g(i, j) + g(j, i) 1i yj ) In fact, the Hamming loss is a sum of the conventional 0/1 loss functions widely used in binary classification, so the form of the risk minimizer is not surprising. 3.2. Rank Loss Another loss function that is commonly used in MLC is the rank loss. Instead of comparing two label subsets, this loss function compares the true label subset with a predicted ranking (total order) of labels, as represented by the ranking function (1). In this ranking, 1 For a predicate P , the expression P evaluates to 1 if P is true and to 0 if P is false. fi < fj + 1 fi = fj 2 . For each pair of labels yi , yj , the sum g(i, j) + g(j, i) is obviously minimized by choosing the scores fi , fj such that fi fj if and only if Px (yi > yj ) Px (yj > yi ), (i) (j) and since Px (yi > yj )-Px (yj > yi ) = Px (1)-Px (1), the condition on the right-hand side is equivalent to (i) (j) Px (1) Px (1). Consequently, the scores (8) minimize the sums g(i, j) + g(j, i) simultaneously for all label pairs and, therefore, minimize risk. Bayes-Optimal Multilabel Classification via Probabilistic Classifier Chains 3.3. Subset Zero-One Loss Finally, it is of course also possible to generalize the well-known 0/1 loss from the conventional to the multilabel setting: Ls (y, h(x)) = y = h(x) . (9) 4. Probabilistic Classifier Chains Given a query instance x, the (conditional) probability of each label combination y = (y1 , . . . , ym ) Y can be computed using the product rule of probability: m Px (y) = Px (y1 ) · i=2 Px (yi | y1 , . . . , yi-1 ) (11) This loss function is referred to as subset zero-one loss. Admittedly, it may appear overly stringent, especially in the case of many labels. Moreover, since making a mistake on a single label is punished as hardly as a mistake on all labels, it does not discriminate well between "almost correct" and completely wrong predictions. Still, this measure is obviously interesting with regard to dependence between labels. The Bayes prediction for (9) is rather straight-forward. As for any other 0/1 loss, it simply consists of predicting the mode of the distribution: h (x) = arg max Px (y) . yY Thus, to estimate the joint distribution of labels, one possibility is to learn m functions fi (·) on an augmented input space X × {0, 1}i-1 , taking y1 , . . . , yi-1 as additional attributes: fi : X × {0, 1}i-1 (x, y1 , . . . , yi-1 ) [0, 1] P(yi = 1 | x, y1 , . . . , yi-1 ) We assume here that the function fi (·) can be interpreted as a probabilistic classifier whose prediction is the probability that yi = 1, or at least a reasonable approximation thereof. Thus, (11) becomes m (10) Px (y) = f1 (x) · fi (x, y1 , . . . , yi-1 ) i=2 (12) 3.4. Consequences and Conjectures As one of the most important consequences of the above results we note that, according to (6) and (7), a risk-minimizing prediction for the Hamming and the rank loss can be obtained from the marginal distribu(i) tions Px (Yi ) (i = 1, . . . , m) alone. In other words, it is not necessary to know the joint label distribution Px (Y) on Y. As opposed to this, (10) shows that the entire distribution of Y given x (or at least the mode of this distribution) is needed to minimize the subset zero-one loss. Let us remark, however, that for conditionally independent labels, the risk minimizers for the Hamming and the subset 0/1 loss are exactly the same. The same holds if the probability of the mode is greater or equal 0.5. Now, since marginal distributions can in principle be estimated independently of each other, without taking conditional dependencies into consideration, we claim that, while measures like (9) may indeed benefit from MLC methods that are able to exploit conditional dependencies, the gain, if any, will be much smaller for measures like Hamming and rank loss. In Section 5, we shall test this conjecture in an empirical way. Before we can do so, we need a method that is able to estimate Bayes optimal predictions. A method of this kind, that is, a learning algorithm producing a model that takes an instance x as input and produces the distribution Px (Y) as output, is proposed in the next section. Given Px (and a loss function L(·) to be minimized), an optimal prediction (4) can then be derived in an explicit way. This approach will subsequently be referred to as the probabilistic classifier chain (PCC). 4.1. The Original Classifier Chain Our method is inspired by the classifier chain (CC) that was recently proposed by Read et al. (2009) as a meta-technique for MLC. More specifically, what they propose is the above idea of "chaining" classifiers, albeit without any connection to probability theory. Their approach works as follows: One classifier hi is trained for each label similarly to our scoring function fi . Given a new instance x to be classified, the model h1 predicts y1 , i.e., the relevance of 1 for x, as usual. Then, h2 predicts the relevance of 2 , taking x plus the predicted value y1 {0, 1} as an input. Proceeding in this way, hi predicts yi using y1 , . . . , yi-1 as additional input information. Interestingly, the original classifier chain can be seen as a deterministic approximation of (12), in the sense of using {0, 1}-valued probabilities. In fact, CC is recovered from (12) in the special case where the fi (·) output either 0 or 1. This results in the estimation Px (y) = y = y CC , (13) where y CC is the label combination predicted by the classifier chain, which pretends certainty about the estimation y CC . Bayes-Optimal Multilabel Classification via Probabilistic Classifier Chains Needless to say, (13) will normally be a poor estimation of the true distribution Px (Y). In fact, considering the idea of chaining classifiers as searching a path in a binary tree whose leaf nodes are associated with a labeling y Y, and with 0/1 branches for yi on level i, CC just follows a single path in this tree in a greedy manner. It can be seen as a "mode seeker", hoping to find the labeling y with highest probability. Due to its greedy nature, however, the mode will not always be reached. For example, suppose that the base classifiers produce exact probability estimates, and that CC turns a probability estimate p into the prediction p > 0.5 . Then, it is easy to show that y CC = y if Px (y ) > 0.5. If the probability of the mode is smaller than 1/2, however, CC may fail. As a small illustration, consider the case m = 3 and suppose that Px (0, 0, 1) = 0.4, Px (1, 0, 1) = 0.25 and Px (1, 1, 0) = 0.35. In this case, CC will already start incorrectly, namely with y1 = 1, and eventually produce the suboptimal prediction y = (1, 1, 0). 4.2. Complexity The hope that PCC will produce better estimates is clearly justified in light of these observations. Of course, the price to pay is a much higher complexity. In fact, while CC searches only a single path in the aforementioned binary tree, PCC has to look at each of the 2m paths. This limits the applicability of the method to data sets with a small to moderate number of labels, say, not more than about 15. First, however, apart from the fact that complexity is not our main concern here, one may argue that several other methods suffer from the same problem, including the label power-set approach (Tsoumakas & Katakis, 2007), as also graphical models estimating the joint distribution (Ghamrawi & McCallum, 2005). Second, there are possibilities to develop approximate inference schemes that trade off accuracy against efficiency in a reasonable way, lying somehow in-between the exact inference (12) and the extremely crude approximation (13). This can be done in different ways, for example by pruning single labels (with provably low probability of relevance), or by ignoring label combinations with low probability (to minimize the subset zero-one loss, only the most probable label combination is needed). We can also try to factorize the high-dimensional joint distribution into several lower-dimensional distributions, exploiting label independence whenever possible. 4.3. Chain Ensembles Theoretically, the result of the product rule does not depend on the order of the variables. Practically, two different classifier chains will produce different results, simply because they involve different classifiers learned on different training sets. To reduce the influence of the label order, Read et al. (2009) propose to average the multilabel predictions of CC over a (randomly chosen) set of permutations. Thus, the labels 1 , . . . , m are first re-ordered by a permutation of {1, . . . , m}, which moves the label i from position i to position (i), and CC is then applied as usual. This extension is called the ensembled classifier chain (ECC). Of course, the same idea can be applied to our probabilistic variant. In this case, it is natural to average over the predicted distributions Px (Y) directly. We call this approach the ensembled probabilistic classifier chain (EPCC). 5. Experimental Results In this section, we describe experimental studies on two artificial data sets and twelve benchmark data sets that we performed in order to verify our theoretical conjectures. The experiments are specifically designed to show that the performance of a given classifier strongly depends on the measure used. We have considered the three measures discussed above: Hamming loss, rank loss, and the subset 0/1 loss. We have performed 3-fold cross-validation for each data set except the large benchmark data sets (more than 10000 instances) for which we used the 66% split. The results are reported as an average of these measures over test instances. Moreover, for each data set we provide a ranking of the algorithms. For the benchmark data, the two-step procedure suggested by Demar (2006) was used to test for statistically signifs icant differences between the algorithms. We have considered five classifiers: binary relevance (BR), classifier chains (CC), probabilistic classifier chains (PCC), ensembled classifier chains (ECC), ensembled probabilistic classifier chains (EPCC). BR is the simplest approach that treats labels independently. According to our theoretical considerations, this approach is well-suited for the Hamming and the rank loss. Let us remind that CC is well-suited for the subset 0/1 loss, since it approximates the mode of the joint distribution. ECC averages over several CC predictions, however, it is rather difficult to say, what this approach tends to estimate. The probabilistic versions of these algorithms, PCC and EPCC, are well-suited for all measures, since corresponding risk minimizers Bayes-Optimal Multilabel Classification via Probabilistic Classifier Chains can be computed from the joint distribution estimated by these methods. We tried to eliminate as many as possible additional effects that may bias the results. Since all algorithms are meta-learners, we used logistic regression without regularization as a base learner for all of them. This choice is also motivated by the fact that logistic regression provides (conditional) probability estimates as predictions. CC and ECC are implemented according to (Read et al., 2009) with minor exceptions. We have not performed additional sampling from the data set for training the ensemble members in ECC, and we did not tune the threshold for predicting class labels; we simply set it to 0.5 (as in all algorithms). Permutations of labels in all chain-based algorithms have been generated at random. Let us underline that learning for (E)CC and (E)PCC is the same, and the difference concerns the prediction phase only. The ensemble size in ECC and EPCC was set to 10. Let us also notice that the results highly depend on data sets used. For data sets without dependencies among the labels or with high probability of the joint modes ( 0.5), the risk minimizers for the Hamming and the subset 0/1 loss coincide. Of course, it is hard to estimate the nature of a data set in advance, since even a high correlation between labels in the data set does not mean that the joint distribution for given x shares these dependencies and vice versa. 5.1. Artificial Data We have used two artificial data sets with three labels. The first one is a collection of independent problems. In the second one, the labels are strongly dependent. The data models used for generating these data sets are supposed to be as simple as possible. For each data set, we generated 10000 instances. The independent data set was generated by uniformly drawing instances from the square x [-0.5, 0.5]2 . The label distribution is given by the product of the marginal distributions defined by Px (yi ) = 1 (1 + exp(-fi (x)), where the fi are linear functions: f1 (x) = x1 + x2 , f2 (x) = -x1 + x2 , f3 (x) = x1 - x2 . The cardinality of labels (the average number of relevant labels for an instance) is 1.503. The dependent data set was generated by drawing the instances from a univariate uniform distribution x [-0.5, 0.5]. The label distribution is given by the product rule2 : Px (Y) = Px (y1 )Px (y2 | y1 )Px (y3 | y1 , y2 ), 2 We have experimented with several models of dependence and obtained similar results; the product rule model is presented, since it is prominently discussed in the paper. Table 1. Results on two artificial data sets: independent model (top) and dependent model (down). classifier BR CC PCC ECC EPCC B-O BR CC PCC ECC EPCC B-O Hamming loss 0.4178(2) 0.4189(4.5) 0.4178(2) 0.4189(4.5) 0.4178(2) 0.4179 0.3921(3) 0.4308(4) 0.3920(1.5) 0.4320(5) 0.3920(1.5) 0.3920 rank loss 0.5527(1) 0.5934(5) 0.5528(2.5) 0.5907(4) 0.5528(2.5) 0.5532 0.5675(2) 0.6930(4) 0.5676(3) 0.6954(5) 0.5674(1) 0.5671 subset 0/1 loss 0.8108(3) 0.8124(5) 0.8088(1.5) 0.8120(4) 0.8088(1.5) 0.8088 0.7374(5) 0.6100(3) 0.6052(2) 0.6112(4) 0.6051(1) 0.6057 where the probabilities are modeled by linear functions in a similar way as before: f1 (x) = x, f2 (y1 , x) = -x - 2y1 + 1, f3 (y2 , y1 , x) = x + 12y1 - 2y2 - 11. The cardinality of labels for this data set is 1.314. The results are shown in Table 5.1. In the case of the independent data we can see that, in agreement with our theoretical considerations, there is almost no difference between the algorithms with respect to the Hamming and the subset 0/1 loss. The poor performance of CC and ECC on rank loss follows from the fact that these methods do not estimate the marginals. For the dependent data, we see that PCC and EPCC adapt to the loss function. Moreover, as expected, BR performs well for the Hamming and the rank loss. CC and ECC are better than BR with respect to the subset 0/1 loss, which is in agreement with our theoretical results. However, we observe that both algorithms approximate the optimal prediction in a suboptimal way and are outperformed by PCC and EPCC. It seems that ensembling the CC classifiers does not improve performance. Since both models are known, we have also computed the Bayes-optimal predictions (denoted by B-O). By comparing the performance to B-O, one can measure the regret, that is, the drop in performance caused by not being well-adapted to the loss (since B-O predictions are computed on a finite sample, some classifiers may obtain slightly better results). 5.2. Benchmark Data The second part of the experiment was performed on a relatively large collection of 12 multilabel classification data sets3 . A summary of the data sets and 3 Taken from mlkd.csd.auth.gr/multilabel.html and www.cs.waikato.ac.nz/~jmr30/#datasets Bayes-Optimal Multilabel Classification via Probabilistic Classifier Chains their properties are given in Table 2. Since PCC and EPCC are computationally complex, we have limited the number of labels to 10. For each data set, we kept the most frequent labels (and subsequently removed all instances having only relevant or only irrelevant labels). The reuters data set has been preprocessed as u in (Cheng & H¨llermeier, 2009). The results are summarized in Table 3. We conducted the Friedman test based on the average ranks in order to verify whether the difference between algorithms are statistically significant. For each loss function, the null hypothesis was rejected at a significance level of 5%. According to the post-hoc analysis based on Nemenyi statistics, the significant difference in average ranks of the algorithms is 1.84. The results are in agreement with our theoretical conjectures and the previous experiment on artificial data. The EPCC algorithm performs best, as it can be tailored for all loss functions and benefits from averaging over the ensemble. It seems that for those benchmark data sets that are much larger in the feature and label space, ensembling produces a significant improvement: Both EPCC and ECC outperform their respective regular variant (for the Hamming and the rank loss, the difference is significant). We note, however, that the comparison between ensemble and non-ensemble methods is not completely fair. This is why we claim that BR performs well with respect to the Hamming and the rank loss. Using a more powerful classifier in BR, we even expect the difference to EPCC to decrease further. The worst in this case is the CC algorithm, as it is not tailored for these losses. In the case of the subset 0/1 loss, BR is obviously the worst, and CC obtains better results. It is interesting that ECC has the second lowest average rank for the Hamming and the rank loss, which may suggest that the averaging used in this classifier brings the prediction closer to the marginals. However, this method has also obtained the second lowest average rank for the subset 0/1 loss. The interpretation of this result is difficult, since we do not know the real dependencies between labels. It seems that some of the data sets do not contain strong dependencies or the probability of joint modes is high ( 0.5), which could partly explain the overall good performance of ECC. important insights into the nature of MLC. In particular, they show that the main concern of recent contributions to MLC, namely the exploitation of label dependence, should be considered with diligence: First, one has to distinguish two types of dependence, conditional and unconditional, and second, the latter is strongly related to the loss function to be minimized. A second contribution is a new MLC method, probabilistic classifier chains, that extends the recently introduced CC classifier. This algorithm is able to estimate the entire joint distribution of the labels and, therefore, can be tailored to any loss function. Its drawback is the computational complexity at prediction time. We therefore plan to develop approximate inference schemes that trade off complexity against accuracy of probability estimation in an optimal way. Our analysis has also shed light on the CC classifier itself that was lacking a sound theoretical motivation so far. Here, we have shown that it can be seen as a greedy approximation of the most probable label combination. This interpretation enables a formal analysis of the algorithm. Acknowledgments We thank Willem Waegeman for interesting discussions and useful suggestions. This work has been supported by the Germany Research Foundation (DFG). References Breiman, L. and Friedman, J. Predicting multivariate responses in multiple linear regression. J R Stat Soc B, 69:3­54, 1997. Cheng, W. and H¨llermeier, E. Combining instanceu based learning and logistic regression for multilabel classification. Machine Learning, 76:211­225, 2009. Demar, J. Statistical comparisons of classifiers over s multiple data sets. JMRL, 7:1­30, 2006. Ghamrawi, N. and McCallum, A. Collective multilabel classification. In CIKM '05, pp. 195­200, 2005. Godbole, S. and Sarawagi, S. Discriminative methods for multi-labeled classification. In PAKDD 2004, pp. 22­30, 2004. Read, J., Pfahringer, B., Holmes, G., and Frank, E. Classifier chains for multi-label classification. In ECML/PKDD 2009, pp. 254­269, 2009. Tsoumakas, G. and Katakis, I. Multi label classification: An overview. Int J Data Warehouse and Mining, 3:1­13, 2007. 6. Conclusions We proposed a probabilistic framework of multilabel classification and analyzed the risk minimizers of three common loss functions. Even though most of the theoretical results are quite obvious, they provide some Bayes-Optimal Multilabel Classification via Probabilistic Classifier Chains data set Table 2. Data sets used in the experiment. # inst. # attr. # labels cardinality data set # inst. # attr. # labels cardinality 135 294 72 243 103 120 5 6 6 7 10 10 1.236 1.074 1.868 1.241 3.971 3.175 enron-10 1677 medical-10 785 slashdot-10 3279 ohsumed-10 11934 tmc2007-500-10 27409 imdb-10 86290 1001 1449 1079 1002 500 1001 10 10 10 10 10 10 2.513 1.195 1.134 1.461 1.979 1.623 image 2000 scene 2407 emotions 593 reuters 7119 yeast-10 2389 mediamill-10 41583 BR image scene emotions reuters yeast-10 mediamill-10 enron-10 medical-10 slashdot-10 ohsumed-10 tmc-500-10 imdb-10 Ave. Rank image scene emotions reuters yeast-10 mediamill-10 enron-10 medical-10 slashdot-10 ohsumed-10 tmc-500-10 imdb-10 Ave. Rank image scene emotions reuters yeast-10 mediamill-10 enron-10 medical-10 slashdot-10 ohsumed-10 tmc-500-10 imdb-10 Ave. Rank 0.2063 0.1644 0.2395 0.0546 0.2602 0.1628 0.3157 0.0648 0.2230 0.1386 0.1038 0.1632 (2) (3) (2) (3) (3) (3) (3) (5) (5) (3) (4) (3) 3.25 (2) (3) (3) (1) (1) (3) (3) (5) (4) (3) (4) (1) 2.75 Table 3. Results on 12 benchmark data sets. CC PCC ECC Hamming loss 0.2279 0.1780 0.2448 0.0593 0.2736 0.1695 0.3196 0.0641 0.1888 0.1398 0.1059 0.1916 (5) (4.5) (5) (5) (5) (5) (5) (4) (4) (5) (5) (5) 4.792 (5) (5) (5) (5) (5) (5) (4) (4) (5) (5) (5) (5) 4.833 (4) (4) (2) (4) (3) (3) (4.5) (4) (4) (3) (4) (3) 3.542 0.2098 0.1780 0.2417 0.0583 0.2593 0.1621 0.3193 0.0640 0.1887 0.1387 0.1029 0.1629 Rank loss (3.5) (4.5) (3) (4) (2) (2) (4) (3) (3) (4) (2) (1) 3 (4) (4) (4) (4) (3) (2) (5) (3) (3) (4) (3) (3) 3.5 (2.5) (3) (1) (3) (2) (2) (4.5) (3) (3) (4) (2) (1) 2.583 0.2098 0.1503 0.2428 0.0520 0.2651 0.1646 0.3110 0.0550 0.1683 0.1356 0.1035 0.1806 (3.5) (2) (4) (2) (4) (4) (1) (2) (1.5) (2) (3) (4) 2.75 (3) (2) (1) (3) (4) (4) (2) (2) (2) (2) (2) (4) 2.583 (2.5) (2) (4) (2) (4) (4) (2) (1) (1) (1) (3) (4) 2.542 EPCC 0.2033 0.1498 0.2372 0.0514 0.2583 0.1619 0.3111 0.0547 0.1683 0.1344 0.1026 0.1630 (1) (1) (1) (1) (1) (1) (2) (1) (1.5) (1) (1) (2) 1.208 (1) (1) (2) (2) (2) (1) (1) (1) (1) (1) (1) (2) 1.333 (1) (1) (3) (1) (1) (1) (1) (2) (2) (2) (1) (2) 1.5 0.9510 1.0262 1.4367 0.2598 4.4562 2.1063 6.0247 0.7638 3.0933 2.1059 0.9518 3.4139 1.0910 1.1554 1.5816 0.3494 5.3347 2.4565 6.1410 0.7584 3.1232 2.1489 0.9742 3.7588 0.9869 1.0841 1.5445 0.3144 4.5115 2.0755 6.1655 0.7240 3.0461 2.1067 0.9471 3.4424 0.9525 0.8884 1.4332 0.2748 4.6346 2.1631 5.9213 0.5096 2.7585 1.9556 0.9462 3.4514 0.9439 0.8855 1.4349 0.2698 4.4617 2.0720 5.8915 0.4982 2.7219 1.9323 0.9366 3.4260 Subset 0/1 loss 0.6900 (5) 0.6448 (5) 0.8178 (5) 0.2889 (5) 0.8393 (5) 0.7965 (5) 0.9350 (3) 0.4497 (5) 0.8866 (5) 0.7650 (5) 0.6429 (5) 0.9463 (5) 4.833 0.6260 0.5954 0.7757 0.2860 0.7861 0.7690 0.9416 0.4370 0.8588 0.7568 0.6364 0.8391 0.6095 0.5949 0.7723 0.2843 0.7761 0.7517 0.9416 0.4357 0.8582 0.7575 0.6268 0.8239 0.6095 0.5505 0.7976 0.2653 0.8049 0.7698 0.9320 0.3885 0.8112 0.7459 0.6294 0.8411 0.5890 0.5123 0.7790 0.2583 0.7710 0.7504 0.9302 0.3898 0.8286 0.7489 0.6242 0.8259