Improved Boosting Algorithms Using Confidence-rated Predictions

Similar documents
solutions:, and it cannot be the case that a supersolution is always greater than or equal to a subsolution.

Extensional Equality in Intensional Type Theory

A Calculus for End-to-end Statistical Service Guarantees

Random Forests. Gradient Boosting. and. Bagging and Boosting

P(x) testing training. x Hi

ÈÖÓÚ Ò Ò ÁÑÔÐ Ø ÓÒ È É Ï Ö Ø ÐÓÓ Ø Û Ý ØÓ ÔÖÓÚ Ø Ø Ñ ÒØ Ó Ø ÓÖÑ Á È Ø Ò É ÓÖ È É Ì ÓÐÐÓÛ Ò ÔÖÓÓ ØÝÔ Ò Ð Ó Ù ØÓ ÔÖÓÚ Ø Ø Ñ ÒØ Ó Ø ÓÖÑ Ü È Üµ É Üµµ Ý ÔÔ

Two-Way Equational Tree Automata for AC-like Theories: Decidability and Closure Properties

Solutions of Implication Constraints yield Type Inference for More General Algebraic Data Types

38050 Povo (Trento), Italy Tel.: Fax: e mail: url:

Support Vector Machines

Event Based Sequential Program Development: Application to Constructing a Pointer Program

Refinement in Requirements Specification and Analysis: a Case Study

LET Õ Ò µ denote the maximum size of a Õ-ary code

Contact 3-Manifolds, Holomorphic Curves and Intersection Theory

Nominal Techniques in Isabelle/HOL

Supplementary Materials for Strategic Abstention in Proportional Representation Systems (Evidence from Multiple Countries)

ishares Core Composite Bond ETF

Ë ÁÌÇ ÌÓ Ó ÍÒ Ú Ö Øݵ Ç ¼ Ô Û Ö ÙÒÓ Ø Ò Ð Ä Ò ÙÖ ÖÝ ÓÒ ÒÓØ Ý ÛÓÖ Û Ø Ã ÞÙ ÖÓ Á Ö Ó ÒØ Ë Ò ÝÓ ÍÒ Ú Ö Øݵ Ç

Sequential Voting with Externalities: Herding in Social Networks

Infinite-Horizon Policy-Gradient Estimation

Regression. Linear least squares. Support vector regression. increasing the dimensionality fitting polynomials to data over fitting regularization

Estimating the Margin of Victory for Instant-Runoff Voting

Mathematics and Social Choice Theory. Topic 4 Voting methods with more than 2 alternatives. 4.1 Social choice procedures

½º»¾¼ º»¾¼ ¾º»¾¼ º»¾¼ º»¾¼ º»¾¼ º»¾¼ º»¾¼» ¼» ¼ ÌÓØ Ð»½ ¼

How hard is it to control sequential elections via the agenda?

Supporting Information Political Quid Pro Quo Agreements: An Experimental Study

A procedure to compute a probabilistic bound for the maximum tardiness using stochastic simulation

The Effectiveness of Receipt-Based Attacks on ThreeBallot

½ Ê Ú Û Ó ÆÒ ÕÙÓØ ÒØ ¾ ÇÖØ Ó ÓÒ Ð ÒÚ Ö ÒØ ÓÙ Ð Ö Ø ÓÒ Ý ÕÙÓØ ÒØ Ñ Ô ÇÖ Ø ÓÖÖ ÔÓÒ Ò Ü ÑÔÐ Ó ÓÖ Ø ÓÖÖ ÔÓÒ Ò Ü ÑÔÐ Ø Ò ÓÖ ÔÖÓ ÙØ Ü ÑÔÐ ÓÒØÖ Ø ÓÒ Ñ Ô ÇÔ Ò

ÙÒØ ÓÒ Ò Ø ÓÒ ÙÒØ ÓÒ ÖÓÑ ØÓ ÒÓØ Ö Ð Ø ÓÒ ÖÓÑ ØÓ Ù Ø Ø ÓÖ Ú ÖÝ Ü ¾ Ø Ö ÓÑ Ý ¾ Ù Ø Ø Ü Ýµ Ò Ø Ö Ð Ø ÓÒ Ò Ü Ýµ Ò Ü Þµ Ö Ò Ø Ö Ð Ø ÓÒ Ø Ò Ý Þº ÆÓØ Ø ÓÒ Á

Domain, Range, Inverse

ν = fraction of red marbles

Decomposition and Complexity of Hereditary History Preserving Bisimulation on BPP

Computational Inelasticity FHLN05. Assignment A non-linear elasto-plastic problem

Implementing Domain Specific Languages using Dependent Types and Partial Evaluation

ONLINE APPENDIX: Why Do Voters Dismantle Checks and Balances? Extensions and Robustness

ÔÖ Î µ ÛÖ Î Ø Ø Ó ÚÖØ ÖÔ Ø Ø Ó º ØØ Û Ö ÚÒ Ø Ò Ú ¼ ½ Ú ½ ¾ Ú ¾ Ú Ú ½ ÒÒ ÙÒØÓÒ Eº ÏÐ Ò Ø ÖÔ ÕÙÒ Ú ÛÖ Ú ¼ Ú ¾ Î ½ ¾ Ò E µ Ú ½ Ú º Ì ÛÐ ÐÓ Ø Ö Ø Ò Ð Ø ÚÖ

Overview. Ø Neural Networks are considered black-box models Ø They are complex and do not provide much insight into variable relationships

Lab 3: Logistic regression models

MATH4999 Capstone Projects in Mathematics and Economics Topic 3 Voting methods and social choice theory

É ÀÓÛ Ó Ý Ò ² Ö Ò ÁÒ Ö Ò «Ö ÓØ ÑÔ Ù ÔÖÓ Ð ØÝ ØÓ Ö ÙÒ ÖØ ÒØÝ ÙØ Ø Ý ÓÒ Ø ÓÒ ÓÒ «Ö ÒØ Ø Ò º Ü ÑÔÐ ÁÑ Ò Ð Ò Ð ØÖ Ð Û Ø Ò ½ Ñ Ø Ô Ö Ó Ù Ø º ÁÒ Ô Ö ÓÒ Ù Ø

Political Economics II Spring Lectures 4-5 Part II Partisan Politics and Political Agency. Torsten Persson, IIES

Social Rankings in Human-Computer Committees

Tengyu Ma Facebook AI Research. Based on joint work with Yuanzhi Li (Princeton) and Hongyang Zhang (Stanford)

Sampling Equilibrium, with an Application to Strategic Voting Martin J. Osborne 1 and Ariel Rubinstein 2 September 12th, 2002.

The Nominal Datatype Package in Isabelle/HOL

Chapter. Estimating the Value of a Parameter Using Confidence Intervals Pearson Prentice Hall. All rights reserved

Understanding factors that influence L1-visa outcomes in US

Æ ÛØÓÒ³ Å Ø Ó ÐÓ Ì ÓÖÝ Ò ËÓÑ Ø Ò ÓÙ ÈÖÓ ÐÝ Ò³Ø ÃÒÓÛ ÓÙØ Ú º ÓÜ Ñ Ö Ø ÓÐÐ

MSR, Access Control, and the Most Powerful Attacker

3 Electoral Competition

The Integer Arithmetic of Legislative Dynamics

ØÖ Ø Ì Î Ö ÈÖÓ Ö ÑÑ Ò Ä Ò Ù ÁÑÔ Ñ ÒØ Ø ÓÒ ÔÖÓ Ø Ú ÓÔ ÓÖÑ Ý Ú Ö ÑÔ Ñ ÒØ Ø ÓÒ Ó Ø Ë Ñ ÔÖÓ Ö ÑÑ Ò Ò Ù º Ì Ö ÔÓÖØ ÓÙÑ ÒØ Ø Ú Ô ÈÖ Ë Ñ Ò Ù Ù ØÓ ÔÖÓ Ö Ñ Ø Ú

Accept() Reject() Connect() Connect() Above Threshold. Threshold. Below Threshold. Connection A. Connection B. Time. Activity (cells/unit time) CAC

Coalitional Game Theory

Ø Ñ Ò Ò ÙØÙÑÒ ¾¼¼¾ Ò Ò Ö ÕÙ ÒØ ÐÓ µ Ø Û Ø ØÖ ØÖÙØÙÖ ½ ȹØÖ È¹ ÖÓÛØ ÄÇË Ì È¹ØÖ Ø ØÖÙØÙÖ È¹ ÖÓÛØ Ð ÓÖ Ø Ñ ÓÖ Ò Ò ÐÐ Ö ÕÙ ÒØ Ø ÄÇË Ì Ð ÓÖ Ø Ñ ÓÖ Ò Ò Ö ÕÙ

Batch binary Edwards. D. J. Bernstein University of Illinois at Chicago NSF ITR

Corruption and business procedures: an empirical investigation

CS269I: Incentives in Computer Science Lecture #4: Voting, Machine Learning, and Participatory Democracy

ÁÒØÖÓ ÙØ ÓÒ ØÓ ÓÑÔÙØ Ö ÈÖÓ Ö ÑÑ Ò Ò Ü Ñ ÂÙÒ ½ ¾¼¼ È ½ Ü Ö ½ ¾ ½ Å Ö µ µ ÓÒ Ö Ø ÓÓÛ Ò Ñ Ø Ó ÔÙ ÚÓ ÒØ ÒØ µ ß ¼ ¼µ ß Ö ØÙÖÒ ÒØ ¼µ ß ËÝ Ø ÑºÓÙغÔÖ ÒØÒ Ò Ø

Ò ÓÛ Æ ØÛÓÖ Ð ÓÖ Ø Ñ ÓÖ ¹ ÙÐ Ö ÓÒ

"Efficient and Durable Decision Rules with Incomplete Information", by Bengt Holmström and Roger B. Myerson

Tensor. Field. Vector 2D Length. SI BG cgs. Tensor. Units. Template. DOFs u v. Distribution Functions. Domain

Contents. Bibliography 121. Index 123

Subjectivity Classification

Ì ÄÈ Ë ÈÖÓ Ð Ñ Ì ÄÈ Ë ÐÓÒ Ø Ô Ö Ñ Ø Ö Þ ÓÑÑÓÒ Ù ÕÙ Ò µ ÔÖÓ Ð Ñ Ò Ö Ð Þ Ø ÓÒ Ó Û ÐÐ ÒÓÛÒ Ä Ë ÔÖÓ Ð Ñ ÓÒØ Ò Ò Ô¹ÓÒ ØÖ ÒØ º Ò Ø ÓÒ ÁÒ ÄÈ Ë(,, Ã ½, Ã ¾, )

Hoboken Public Schools. College Algebra Curriculum

È Ö Ø ² ÑÔ Ö Ø Ò ÓÖÑ Ø ÓÒ ÓÖ Ñ È Ö Ø Ò ÓÖÑ Ø ÓÒ ÈÐ Ý Ö ÒÓÛ ÓÙØ Ø ÔÖ Ú ÓÙ ÑÓÚ Ó ÓÔÔÓÒ ÒØ º º º Ð ¹ËØ Û ÖØ Ñ º ÁÑÔ Ö Ø Ò ÓÖÑ Ø ÓÒ ÈÐ Ý Ö Ó ÒÓØ ÒÓÛ ÓÙØ Û

Do two parties represent the US? Clustering analysis of US public ideology survey

Optimal Voting Rules for International Organizations, with an. Application to the UN

Chapter 11. Weighted Voting Systems. For All Practical Purposes: Effective Teaching

0.12. localization 0.9 L=11 L=12 L= inverse participation ratio Energy

Complexity of Manipulating Elections with Few Candidates

Expert Mining and Required Disclosure: Appendices

A constraint based dependancy parser for Sanskrit

ÇÙØÐ Ò Ó Ø Ð ÅÓØ Ú Ø ÓÒ ÔÓÐÝÒÓÑ Ð Ú ÓÒ ÒÓ Ò ÓÖ ÝÐ Ó ÙØÓÑÓÖÔ Ñ µ ÑÓ ÙÐ ÕÙ ¹ÝÐ µ ØÖÙ¹ ØÙÖ ÖĐÓ Ò Ö ÓÖ ÑÓ ÙÐ Ú ÐÙ Ø ÓÒ Ó ÖÓÑ ÓÖ Ö ÓÑ Ò Ò¹ ÐÙ Ò ÓÔÔ Ó µ Ü Ñ

ÚÓ Ù ØÖ Ó Ø Ö ÓÙÒØ Øµ ØÖÙØ Ø ÒÓ Ø Ñµ» Ø ÚÓ Ù ØÖ Ó Ø Ö ÓÙÒØ ÔÙص ØÖÙØ Ø ÒÓ Ø Ñµ» Ø ØÖÙØ Ù ØÖ Ó Ý Ö Ò Ñ ½¼ Ô ÒÓ Ø Ó» Ó Ý Ó» ØÖÙØ Ù ØÖ Ù Ø Ø ¾ Ñ Ü Þ» Ò Ø

½½ º º À Æ Æ º º Í Æ ÒÓØ ÔÓ Ø Ú Ñ ¹ Ò Ø ÙÒÐ Ø ÓÐÐÓÛ Ò ØÖÙ Ø Ö ÓÒ Ù ÔÖÓ Ð Ñ È ½ Û Ø Ò Ð ÐÐ ÓÒ ØÖ ÒØ Û Ó ÓÖÑ Ù Ø ØÓ Ñ Ò ¾Ê Ò µ ½ ¾ Ì Ì Ø Ì Ù ÔÖÓ Ð Ñ Ø Ð

¾ ÍÆ ÌÁÇÆ Ä ËÈ Á Á ÌÁÇÆ ÒÚ ÖÓÒÑ ÒØ ½ º½ ÓÖÑ Ø Ò º º º º º º º º º º º º º º º º º º º º º º º º ½ º½º½ Ö ØÓÖÝ ÒØÖ º º º º º º º º º º º º º º º º º º

Deadlock. deadlock analysis - primitive processes, parallel composition, avoidance

ÓÖ Ö ÛÓÖ Ò Ô Ö Ó ØÝ Ò Ø ÛÓÖ ÓÖ Ö Ø ÔÖÓÔ Ö ÔÖ Ü ÕÙ Ð ØÓ Ù Üº ÓÖ Ü ÑÔÐ ÓÖ Ö º Á ÛÓÖ ÒÓØ ÓÖ Ö Û Ý Ø ÙÒ ÓÖ Ö ÓÖ ÓÖ Ö¹ Ö º ÓÖ Ü ÑÔÐ ½¼ Ò = ½¼¼ ¼ Ö ÙÒ ÓÖ Ö

Approval Voting Theory with Multiple Levels of Approval

Comparison Sorts. EECS 2011 Prof. J. Elder - 1 -

Ä ÖÒ Ò ÖÓÑ Ø Ö Ëº Ù¹ÅÓ Ø Ð ÓÖÒ ÁÒ Ø ØÙØ Ó Ì ÒÓÐÓ Ý Ä ØÙÖ ½ Ì Ä ÖÒ Ò ÈÖÓ Ð Ñ ËÔÓÒ ÓÖ Ý ÐØ ³ ÈÖÓÚÓ Ø Ç ² Ë Ú ÓÒ Ò ÁËÌ ÌÙ Ý ÔÖ Ð ¾¼½¾

BIPOLAR MULTICANDIDATE ELECTIONS WITH CORRUPTION by Roger B. Myerson August 2005 revised August 2006

Ï Ó ØÖ Ù ÛÓÖÐ Ý Ù Ð Ø Ö Ø ÓÖ Ð Ö Ð Ø Ú ØÓ Û ÆÈ ËÈ ÊË Ó ÓØ Ú ÓÑÔÐ Ø Ø º Å Ö ÌÓÖ ÅÌ Ú Ö Ð Ø Ú Þ Ð ÔÖÓÓ Ø Ø ÓÔØ Ñ Ð ÔÖÓÓ Ý Ø Ñ Ü Ø Ø ÆÈ ËÈ ÊË Ó Ú ÓÑÔÐ Ø

Computational Social Choice: Spring 2007


Fair Division in Theory and Practice

An overview and comparison of voting methods for pattern recognition

function GENERAL-SEARCH( problem, strategy) returns a solution, or failure initialize the search tree using the initial state of problem loop do if

Experimental Computational Philosophy: shedding new lights on (old) philosophical debates

Density Data

Model of Voting. February 15, Abstract. This paper uses United States congressional district level data to identify how incumbency,

½º ÌÖ ÙØÓÑØ

Learning and Visualizing Political Issues from Voting Records Erik Goldman, Evan Cox, Mikhail Kerzhner. Abstract

MODELLING OF GAS-SOLID TURBULENT CHANNEL FLOW WITH NON-SPHERICAL PARTICLES WITH LARGE STOKES NUMBERS

ËØÖÙØÙÖ ½ Î Ö ÐÙ Ø Ö ¹ Ò ÒØÖÓ ÙØ ÓÒ ¾ Ì Ø Ì ÈÙÞÞÐ Ì Á ÓÒÐÙ ÓÒ ÈÖÓ Ð Ñ Å Ö ¹ÄÙ ÈÓÔÔ ÍÒ Ä ÔÞ µ È Ö Ø È ÖØ ÔÐ ¾¼º¼ º½ ¾» ¾

Essential Questions Content Skills Assessments Standards/PIs. Identify prime and composite numbers, GCF, and prime factorization.

Hoboken Public Schools. Algebra II Honors Curriculum

Transcription:

Improved Boosting Algorithms Using Confidence-rated Predictions ÊÇÊÌ º ËÀÈÁÊ schapire@research.att.com AT&T Labs, Shannon Laboratory, 18 Park Avenue, Room A279, Florham Park, NJ 7932-971 ÇÊÅ ËÁÆÊ singer@research.att.com AT&T Labs, Shannon Laboratory, 18 Park Avenue, Room A277, Florham Park, NJ 7932-971 Abstract. We describe several improvements to Freund and Schapire s AdaBoost boosting algorithm, particularly in a setting in which hypotheses may assign confidences to each of their predictions. We give a simplified analysis of AdaBoost in this setting, and we show how this analysis can be used to find improved parameter settings as well as a refined criterion for training weak hypotheses. We give a specific method for assigning confidences to the predictions of decision trees, a method closely related to one used by Quinlan. This method also suggests a technique for growing decision trees which turns out to be identical to one proposed by Kearns and Mansour. We focus next on how to apply the new boosting algorithms to multiclass classification problems, particularly to the multi-label case in which each example may belong to more than one class. We give two boosting methods for this problem, plus a third method based on output coding. One of these leads to a new method for handling the single-label case which is simpler but as effective as techniques suggested by Freund and Schapire. Finally, we give some experimental results comparing a few of the algorithms discussed in this paper. Keywords: Boosting algorithms, multiclass classification, output coding, decision trees 1. Introduction Boosting is a method of finding a highly accurate hypothesis (classification rule) by combining many weak hypotheses, each of which is only moderately accurate. Typically, each weak hypothesis is a simple rule which can be used to generate a predicted classification for any instance. In this paper, we study boosting in an extended framework in which each weak hypothesis generates not only predicted classifications, but also self-rated confidence scores which estimate the reliability of each of its predictions. There are two essential questions which arise in studying this problem in the boosting paradigm. First, how do we modify known boosting algorithms designed to handle only simple predictions to use confidence-rated predictions in the most effective manner possible? Second, how should we design weak learners whose predictions are confidence-rated in the manner described above? In this paper, we give answers to both of these questions. The result is a powerful set of boosting methods for handling more expressive weak hypotheses, as well as an advanced methodology for designing weak learners appropriate for use with boosting algorithms. We base our work on Freund and Schapire s (1997) AdaBoost algorithm which has received extensive empirical and theoretical study (Bauer & Kohavi, to appear; Breiman, 1998; Dietterich, to appear; Dietterich & Bakiri, 199; Drucker & Cortes, 1996; Freund & Schapire, 1996; Maclin & Opitz, 1997; Margineantu & Dietterich, 1997; Quinlan, 1996; Schapire, 1997; Schapire, Freund, Bartlett, & Lee, 1998; Schwenk & Bengio, 1998).

¾ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ To boost using confidence-rated predictions, we propose a generalization of AdaBoost in which the main parameters «Ø are tuned using one of a number of methods that we describe in detail. Intuitively, the «Ø s control the influence of each of the weak hypotheses. To determine the proper tuning of these parameters, we begin by presenting a streamlined version of Freund and Schapire s analysis which provides a clean upper bound on the training error of AdaBoost when the parameters «Ø are left unspecified. For the purposes of minimizing training error, this analysis provides an immediate clarification of the criterion that should be used in setting «Ø. As discussed below, this analysis also provides the criterion that should be used by the weak learner in formulating its weak hypotheses. Based on this analysis, we give a number of methods for choosing «Ø. We show that the optimal tuning (with respect to our criterion) of «Ø can be found numerically in general, and we give exact methods of setting «Ø in special cases. Freund and Schapire also considered the case in which the individual predictions of the weak hypotheses are allowed to carry a confidence. However, we show that their setting of «Ø is only an approximation of the optimal tuning which can be found using our techniques. We next discuss methods for designing weak learners with confidence-rated predictions using the criterion provided by our analysis. For weak hypotheses which partition the instance space into a small number of equivalent prediction regions, such as decision trees, we present and analyze a simple method for automatically assigning a level of confidence to the predictions which are made within each region. This method turns out to be closely related to a heuristic method proposed by Quinlan (1996) for boosting decision trees. Our analysis can be viewed as a partial theoretical justification for his experimentally successful method. Our technique also leads to a modified criterion for selecting such domain-partitioning weak hypotheses. In other words, rather than the weak learner simply choosing a weak hypothesis with low training error as has usually been done in the past, we show that, theoretically, our methods work best when combined with a weak learner which minimizes an alternative measure of badness. For growing decision trees, this measure turns out to be identical to one earlier proposed by Kearns and Mansour (1996). Although we primarily focus on minimizing training error, we also outline methods that can be used to analyze generalization error as well. Next, we show how to extend the methods described above for binary classification problems to the multiclass case, and, more generally, to the multi-label case in which each example may belong to more than one class. Such problems arise naturally, for instance, in text categorization problems where the same document (say, a news article) may easily be relevant to more than one topic (such as politics, sports, etc.). Freund and Schapire (1997) gave two algorithms for boosting multiclass problems, but neither was designed to handle the multi-label case. In this paper, we present two new extensions of AdaBoost for multi-label problems. In both cases, we show how to apply the results presented in the first half of the paper to these new extensions. In the first extension, the learned hypothesis is evaluated in terms of its ability to predict a good approximation of the set of labels associated with a given instance. As a special case, we obtain a novel boosting algorithm for multiclass problems in the more conventional single-label case. This algorithm is simpler but apparently as effective as the methods given by Freund and Schapire. In addition, we propose and analyze a modification of

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË this method which combines these techniques with Dietterich and Bakiri s (199) outputcoding method. (Another method of combining boosting and output coding was proposed by Schapire (1997). Although superficially similar, his method is in fact quite different from what is presented here.) In the second extension to multi-label problems, the learned hypothesis instead predicts, for a given instance, a ranking of the labels, and it is evaluated based on its ability to place the correct labels high in this ranking. Freund and Schapire s AdaBoost.M2 is a special case of this method for single-label problems. Although the primary focus of this paper is on theoretical issues, we give some experimental results comparing a few of the new algorithms. We obtain especially dramatic improvements in performance when a fairly large amount of data is available, such as large text categorization problems. 2. A Generalized Analysis of Adaboost Let Ë Ü ½ Ý ½ µ Ü Ñ Ý Ñ µ be a sequence of training examples where each instance Ü belongs to a domain or instance space, and each label Ý belongs to a finite label space. For now, we focus on binary classification problems in which ½ ½. We assume access to a weak or base learning algorithm which accepts as input a sequence of training examples Ë along with a distribution over ½ Ñ, i.e., over the indices of Ë. Given such input, the weak learner computes a weak (or base) hypothesis. In general, has the form Ê. We interpret the sign of ܵ as the predicted label ( ½ or ½) to be assigned to instance Ü, and the magnitude ܵ as the confidence in this prediction. Thus, if ܵ is close to or far from zero, it is interpreted as a low or high confidence prediction. Although the range of may generally include all real numbers, we will sometimes restrict this range. The idea of boosting is to use the weak learner to form a highly accurate prediction rule by calling the weak learner repeatedly on different distributions over the training examples. A slightly generalized version of Freund and Schapire s AdaBoost algorithm is shown in Figure 1. The main effect of AdaBoost s update rule, assuming «Ø ¼, is to decrease or increase the weight of training examples classified correctly or incorrectly by Ø (i.e., examples for which Ý and Ø Ü µ agree or disagree in sign). Our version differs from Freund and Schapire s in that (1) weak hypotheses can have range over all of Ê rather than the restricted range ½ ½ assumed by Freund and Schapire; and (2) whereas Freund and Schapire prescribe a specific choice of «Ø, we leave this choice unspecified and discuss various tunings below. Despite these differences, we continue to refer to the algorithm of Figure 1 as AdaBoost. As discussed below, when the range of each Ø is restricted to ½ ½, we can choose «Ø appropriately to obtain Freund and Schapire s original AdaBoost algorithm (ignoring superficial differences in notation). Here, we give a simplified analysis of the algorithm in which «Ø is left unspecified. This analysis yields an improved and more general method for choosing «Ø. Let

ʺ º ËÀÈÁÊ Æ º ËÁÆÊ Given: Ü ½ Ý ½ µ Ü Ñ Ý Ñ µ ; Ü ¾, Ý ¾ ½ ½ Initialize ½ µ ½Ñ. For Ø ½ Ì : Train weak learner using distribution Ø. Get weak hypothesis Ø Ê. Choose «Ø ¾ Ê. Update: Ø ½ µ Ø µ ÜÔ «Ø Ý Ø Ü µµ Ø where Ø is a normalization factor (chosen so that Ø ½ will be a distribution). Output the final hypothesis: À ܵ Ò Ì Ø½ «Ø Ø Üµ Figure 1. A generalized version of AdaBoost. ܵ Ì Ø½ «Ø Ø Üµ so that À ܵ Ò Üµµ. Also, for any predicate, let be ½ if holds and ¼ otherwise. We can prove the following bound on the training error of À. THEOREM 1 Assuming the notation of Figure 1, the following bound holds on the training error of À: ½ Ñ À Ü µ Ý Ì Ø½ Ø Proof: By unraveling the update rule, we have that Ì ½ µ ÜÔ È Ø «ØÝ Ø Ü µµ Ñ É Ø Ø ÜÔ É Ý Ü µµ Ñ Ø (1) Ø Moreover, if À Ü µ Ý then Ý Ü µ ¼ implying that ÜÔ Ý Ü µµ ½. Thus, À Ü µ Ý ÜÔ Ý Ü µµ (2)

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË Combining Eqs. (1) and (2) gives the stated bound on training error since ½ Ñ À Ü µ Ý ½ ÜÔ Ý Ü µµ Ñ Ø Ì ½ µ Ø Ø Ø The important consequence of Theorem 1 is that, in order to minimize training error, a reasonable approach might be to greedily minimize the bound given in the theorem by minimizing Ø on each round of boosting. We can apply this idea both in the choice of «Ø and as a general criterion for the choice of weak hypothesis Ø. Before proceeding with a discussion of how to apply this principle, however, we digress momentarily to give a slightly different view of AdaBoost. Let À ½ Æ be the space of all possible weak hypotheses, which, for simplicity, we assume for the moment to be finite. Then AdaBoost attempts to find a linear threshold of these weak hypotheses which gives good predictions, i.e., a function of the form À ܵ Ò ¼ Æ ½ ½ ܵ By the same argument used in Theorem 1, it can be seen that the number of training mistakes of À is at most Ñ ½ ÜÔ ¼ ½ Æ Ý Ü µ ½ (3) AdaBoost can be viewed as a method for minimizing the expression in Eq. (3) over the coefficients by a greedy coordinate-wise search: On each round Ø, a coordinate is chosen corresponding to Ø, that is, Ø. Next, the value of the coefficient is modified by adding «Ø to it; all other coefficient are left unchanged. It can be verified that the quantity Ø measures É exactly the ratio of the new to the old value of the exponential sum in Eq. (3) so that Ø Ø is the final value of this expression (assuming we start with all s set to zero). See Friedman, Hastie and Tibshirani (1998) for further discussion of the rationale for minimizing Eq. (3), including a connection to logistic regression. See also Appendix A for further comments on how to minimize expressions of this form. 3. Choosing «Ø To simplify notation, let us fix Ø and let Ù Ý Ø Ü µ, Ø, Ø, Ø and ««Ø. In the following discussion, we assume without loss of generality that µ ¼ for all. Our goal is to find «which minimizes or approximately minimizes as a function of «. We describe a number of methods for this purpose.

ʺ º ËÀÈÁÊ Æ º ËÁÆÊ 3.1. Deriving Freund and Schapire s choice of «Ø We begin by showing how Freund and Schapire s (1997) version of AdaBoost can be derived as a special case of our new version. For weak hypotheses with range ½ ½, their choice of «can be obtained by approximating as follows: µ «Ù µ ½ Ù ¾ «½ Ù «(4) ¾ This upper bound is valid since Ù ¾ ½ ½, and is in fact exact if has range ½ ½ (so that Ù ¾ ½ ½). (A proof of the bound follows immediately from the convexity of «Ü for any constant «¾ Ê.) Next, we can analytically choose «to minimize the right hand side of Eq. (4) giving «½ ¾ ÐÒ ½ Ö ½ Ö where Ö È µù. Plugging into Eq. (4), this choice gives the upper bound Ô ½ Ö ¾ We have thus proved the following corollary of Theorem 1 which is equivalent to Freund and Schapire s (1997) Theorem 6: COROLLARY 1 ((FREUND & SCHAPIRE, 1997)) Using the notation of Figure 1, assume each Ø has range ½ ½ and that we choose where ½ ÖØ «Ø ½ ¾ ÐÒ ½ Ö Ø Ö Ø Ø µý Ø Ü µ Ø Ý Ø Ü µ Then the training error of À is at most Ì Õ ½ ÖØ ¾ ؽ Thus, with this setting of «Ø, it is reasonable to try to find Ø that maximizes Ö Ø on each round of boosting. This quantity Ö Ø is a natural measure of the correlation of the predictions of Ø and the labels Ý with respect to the distribution Ø. It is closely related to ordinary error since, if Ø has range ½ ½ then

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ÈÖ Ø Ø Ü µ Ý ½ Ö Ø ¾ so maximizing Ö Ø is equivalent to minimizing error. More generally, if Ø has range ½ ½ then ½ Ö Ø µ¾ is equivalent to the definition of error used by Freund and Schapire ( Ø in their notation). The approximation used in Eq. (4) is essentially a linear upper bound of the function «Ü on the range Ü ¾ ½ ½. Clearly, other upper bounds which give a tighter approximation could be used instead, such as a quadratic or piecewise-linear approximation. 3.2. A numerical method for the general case We next give a general numerical method for exactly minimizing with respect to «. Recall that our goal is to find «which minimizes «µ µ «Ù The first derivative of is ¼ «µ «µù «Ù Ø ½ µù by definition of Ø ½. Thus, if Ø ½ is formed using the value of «Ø which minimizes Ø (so that ¼ «µ ¼), then we will have that Ø ½ µù Ø ½ Ý Ø Ü µ ¼ In words, this means that, with respect to distribution Ø ½, the weak hypothesis Ø will be exactly uncorrelated with the labels Ý. It can easily be verified that ¼¼ «µ ¾ «¾ is strictly positive for all «¾ Ê (ignoring the trivial case that Ù ¼ for all ). Therefore, ¼ «µ can have at most one zero. (See also Appendix A.) Moreover, if there exists such that Ù ¼ then ¼ «µ ½ as «½. Similarly, ¼ «µ ½ as «½ if Ù ¼ for some. This means that ¼ «µ has at least one root, except in the degenerate case that all non-zero Ù s are of the same sign. Furthermore, because ¼ «µ is strictly increasing, we can numerically find the unique minimum of «µ by a simple binary search, or more sophisticated numerical methods. Summarizing, we have argued the following: THEOREM 2 1. Assume the set Ý Ø Ü µ ½ Ñ includes both positive and negative values. Then there exists a unique choice of «Ø which minimizes Ø. 2. For this choice of «Ø, we have that Ø ½ Ý Ø Ü µ ¼

ʺ º ËÀÈÁÊ Æ º ËÁÆÊ 3.3. An analytic method for weak hypotheses that abstain We next consider a natural special case in which the choice of «Ø can be computed analytically rather than numerically. Suppose that the range of each weak hypothesis Ø is now restricted to ½ ¼ ½. In other words, a weak hypothesis can make a definitive prediction that the label is ½ or ½, or it can abstain by predicting ¼. No other levels of confidence are allowed. By allowing the weak hypothesis to effectively say I don t know, we introduce a model analogous to the specialist model of Blum (1997), studied further by Freund et al. (1997). For fixed Ø, let Ï ¼, Ï ½, Ï ½ be defined by Ï Ù µ for ¾ ½ ¼ ½, where, as before, Ù Ý Ø Ü µ, and where we continue to omit the subscript Ø when clear from context. Also, for readability of notation, we will often abbreviate subscripts ½ and ½ by the symbols and so that Ï ½ is written Ï, and Ï ½ is written Ï. We can calculate as: µ «Ù ¾ ½¼ ½ Ù Ï ¼ Ï µ ««Ï «It can easily be verified that is minimized when «½ Ï ¾ ÐÒ Ï For this setting of «, we have Ï ¼ ¾Ô Ï Ï () For this case, Freund and Schapire s original AdaBoost algorithm would instead have made the more conservative choice Ï «½ ¾ ÐÒ ½Ï ¾ ¼ Ï ½Ï ¾ ¼ giving a value of which is necessarily inferior to Eq. (), but which Freund and Schapire (1997) are able to upper bound by ¾ Õ Ï ½Ï ¾ ¼µ Ï ½Ï ¾ ¼µ (6) If Ï ¼ ¼ (so that has range ½ ½), then the choices of «and resulting values of are identical.

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË 4. A Criterion for Finding Weak Hypotheses So far, we have only discussed using Theorem 1 to choose «Ø. In general, however, this theorem can be applied more broadly to guide us in the design of weak learning algorithms which can be combined more powerfully with boosting. In the past, it has been assumed that the goal of the weak learning algorithm should be to find a weak hypothesis Ø with a small number of errors with respect to the given distribution Ø over training samples. The results above suggest, however, that a different criterion can be used. In particular, we can attempt to greedily minimize the upper bound on training error given in Theorem 1 by minimizing Ø on each round. Thus, the weak learner should attempt to find a weak hypothesis Ø which minimizes Ø Ø µ ÜÔ «Ø Ý Ø Ü µµ This expression can be simplified by folding «Ø into Ø, in other words, by assuming without loss of generality that the weak learner can freely scale any weak hypothesis by any constant factor «¾ Ê. Then (omitting Ø subscripts), the weak learner s goal now is to minimize µ ÜÔ Ý Ü µµ (7) For some algorithms, it may be possible to make appropriate modifications to handle such a loss function directly. For instance, gradient-based algorithms, such as backprop, can easily be modified to minimize Eq. (7) rather than the more traditional mean squared error. We show how decision-tree algorithms can be modified based on the new criterion for finding good weak hypotheses. 4.1. Domain-partitioning weak hypotheses We focus now on weak hypotheses which make their predictions based on a partitioning of the domain. To be more specific, each such weak hypothesis is associated with a partition of into disjoint blocks ½ Æ which cover all of and for which ܵ Ü ¼ µ for all Ü Ü ¼ ¾. In other words, s prediction depends only on which block a given instance falls into. A prime example of such a hypothesis is a decision tree whose leaves define a partition of the domain. Suppose that Ø and that we have already found a partition ½ Æ of the space. What predictions should be made for each block of the partition? In other words, how do we find a function Ê which respects the given partition and which minimizes Eq. (7)? Let ܵ for Ü ¾. Our goal is to find appropriate choices for. For each and for ¾ ½ ½, let Ï Ü ¾ Ý µ ÈÖ Ü ¾ Ý

½¼ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ be the weighted fraction of examples which fall in block with label. Then Eq. (7) can be rewritten µ ÜÔ Ý µ Ü ¾ Ï Ï (8) Using standard calculus, we see that this is minimized when Ï ½ ¾ ÐÒ Ï (9) Plugging into Eq. (8), this choice gives Õ ¾ Ï Ï () Note that the sign of is equal to the (weighted) majority class within block. Moreover, will be close to zero (a low confidence prediction) if there is a roughly equal split of positive and negative examples in block. Likewise, will be far from zero if one label strongly predominates. A similar scheme was previously proposed by Quinlan (1996) for assigning confidences to the predictions made at the leaves of a decision tree. Although his scheme differed in the details, we feel that our new theory provides some partial justification for his method. The criterion given by Eq. () can also be used as a splitting criterion in growing a decision tree, rather than the Gini index or an entropic function. In other words, the decision tree could be built by greedily choosing the split which causes the greatest drop in the value of the function given in Eq. (). In fact, exactly this splitting criterion was proposed by Kearns and Mansour (1996). Furthermore, if one wants to boost more than one decision tree then each tree can be built using the splitting criterion given by Eq. () while the predictions at the leaves of the boosted trees are given by Eq. (9). 4.2. Smoothing the predictions The scheme presented above requires that we predict as in Eq. (9) on block. It may well happen that Ï or Ï is very small or even zero, in which case will be very large or infinite in magnitude. In practice, such large predictions may cause numerical problems. In addition, there may be theoretical reasons to suspect that large, overly confident predictions will increase the tendency to overfit. To limit the magnitudes of the predictions, we suggest using instead the smoothed values Ï ½ ¾ ÐÒ Ï

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ½½ for some appropriately small positive value of. Because Ï and Ï are both bounded between ¼ and ½, this has the effect of bounding by ½ ½ ¾ ÐÒ ½ ¾ ÐÒ ½µ Moreover, this smoothing only slightly weakens the value of since, plugging into Eq. (8) gives ¼ ¾ ÚÙ Ù ÚÙ Ù ½ Ï Ø Ï Ï Ï Ø Ï Ï Õ Ï µï Õ Ï µï Õ Õ Ï ÕÏ ¾ Ï Ï Õ Ï Ô Ï ¾Æ (11) In the second inequality, we used the inequality Ô Ü Ý Ô Ü ÔÝ for nonnegative Ü and Ý. In the last inequality, we used the fact that which implies Ï Ï µ ½ Õ Ï Õ Ï Ô ¾Æ (Recall that Æ is the number of blocks in the partition.) Thus, comparing Eqs. (11) and (), we see that will not be greatly degraded by smoothing if we choose ½ ¾Æ µ. In our experiments, we have typically used on the order of ½Ñ where Ñ is the number of training examples.. Generalization Error So far, we have only focused on the training error, even though our primary objective is to achieve low generalization error. Two methods of analyzing the generalization error of AdaBoost have been proposed. The first, given by Freund and Schapire (1997), uses standard VC-theory to bound the generalization error of the final hypothesis in terms of its training error and an additional term which is a function of the VC-dimension of the final hypothesis class and the number of training examples. The VC-dimension of the final hypothesis class can be computed using the methods of Baum and Haussler (1989). Interpretting the derived upper bound as

½¾ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ a qualitative prediction of behavior, this analysis suggests that AdaBoost is more likely to overfit if run for too many rounds. Schapire et al. (1998) proposed an alternative analysis to explain AdaBoost s empirically observed resistance to overfitting. Following the work of Bartlett (1998), this method is based on the margins achieved by the final hypothesis on the training examples. The margin is a measure of the confidence of the prediction. Schapire et al. show that larger margins imply lower generalization error regardless of the number of rounds. Moreover, they show that AdaBoost tends to increase the margins of the training examples. To a large extent, their analysis can be carried over to the current context, which is the focus of this section. As a first step in applying their theory, we assume that each weak hypothesis Ø has bounded range. Recall that the final hypothesis has the form where À ܵ Ò Üµµ ܵ Ø «Ø Ø Üµ Since the Ø s are bounded and since we only care about the sign of, we can rescale the Ø s and normalize the «Ø s allowing us È to assume without loss of generality that each Ø ½ ½, each «Ø ¾ ¼ ½ and Ø «Ø ½. Let us also assume that each Ø belongs to a hypothesis space À. Schapire et al. define the margin of a labeled example Ü Ýµ to be Ý Üµ. The margin then is in ½ ½, and is positive if and only if À makes a correct prediction on this example. We further regard the magnitude of the margin as a measure of the confidence of À s prediction. Schapire et al. s results can be applied directly in the present context only in the special case that each ¾ À has range ½ ½. This case is not of much interest, however, since our focus is on weak hypotheses with real-valued predictions. To extend the margins theory, then, let us define to be the pseudodimension of À (for definitions, see, for instance, Haussler (1992)). Then using the method sketched in Section 2.4 of Schapire et al. together with Haussler and Long s (199) Lemma 13, we can prove the following upper bound on generalization error which holds with probability ½ Æ for all ¼ and for all of the form above: ÈÖ Ë Ý Üµ Ç ½ ÐÓ ¾ ѵ Ô ÐÓ ½Æµ Ñ ¾ ½¾ Here, ÈÖ Ë denotes probability with respect to choosing an example Ü Ýµ uniformly at random from the training set. Thus, the first term is the fraction of training examples with margin at most. A proof outline of this bound was communicated to us by Peter Bartlett and is provided in Appendix B. Note that, as mentioned in Section 4.2, this margin-based analysis suggests that it may be a bad idea to allow weak hypotheses which sometimes make predictions that are very large in magnitude. If Ø Üµ is very large for some Ü, then rescaling Ø leads to a very

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ½ large coefficient «Ø which, in turn, may overwhelm the other coefficients and so may dramatically reduce the margins of some of the training examples. This, in turn, according to our theory, can have a detrimental effect on the generalization error. It remains to be seen if this theoretical effect will be observed in practice, or, alternatively, if an improved theory can be developed. 6. Multiclass, Multi-label Classification Problems We next show how some of these methods can be extended to the multiclass case in which there may be more than two possible labels or classes. Moreover, we will consider the more general multi-label case in which a single example may belong to any number of classes. Formally, we let be a finite set of labels or classes, and let. In the traditional classification setting, each example Ü ¾ is assigned a single class Ý ¾ (possibly via a stochastic process) so that labeled examples are pairs Ü Ýµ. The goal then, typically, is to find a hypothesis À which minimizes the probability that Ý À ܵ on a newly observed example Ü Ýµ. In the multi-label case, each instance Ü ¾ may belong to multiple labels in. Thus, a labeled example is a pair Ü µ where is the set of labels assigned to Ü. The single-label case is clearly a special case in which ½ for all observations. It is unclear in this setting precisely how to formalize the goal of a learning algorithm, and, in general, the right formalization may well depend on the problem at hand. One possibility is to seek a hypothesis which attempts to predict just one of the labels assigned to an example. In other words, the goal is to find À which minimizes the probability that À ܵ ¾ on a new observation Ü µ. We call this measure the oneerror of hypothesis À since it measures the probability of not getting even one of the labels correct. We denote the one-error of a hypothesis with respect to a distribution over observations Ü µ by one-err Àµ. That is, one-err Àµ ÈÖ Ü µ À ܵ ¾ Note that, for single-label classification problems, the one-error is identical to ordinary error. In the following sections, we will introduce other loss measures that can be used in the multi-label setting, namely, Hamming loss and ranking loss. We also discuss modifications to AdaBoost appropriate to each case. 7. Using Hamming Loss for Multiclass Problems Suppose now that the goal is to predict all and only all of the correct labels. In other words, the learning algorithm generates a hypothesis which predicts sets of labels, and the loss depends on how this predicted set differs from the one that was observed. Thus, À ¾ and, with respect to a distribution, the loss is ½ Ü µ ܵ

½ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ Given: Ü ½ ½ µ Ü Ñ Ñ µ where Ü ¾, Initialize ½ µ ½ ѵ. For Ø ½ Ì : Train weak learner using distribution Ø. Get weak hypothesis Ø Ê. Choose «Ø ¾ Ê. Update: Ø ½ µ Ø µ ÜÔ «Ø Ø Ü µµ Ø where Ø is a normalization factor (chosen so that Ø ½ will be a distribution). Output the final hypothesis: À Ü µ Ò Ì Ø½ «Ø Ø Ü µ Figure 2. AdaBoost.MH: A multiclass, multi-label version of AdaBoost based on Hamming loss. where denotes symmetric difference. (The leading ½ is meant merely to ensure a value in ¼ ½.) We call this measure the Hamming loss of À, and we denote it by hloss Àµ. To minimize Hamming loss, we can, in a natural way, decompose the problem into orthogonal binary classification problems. That is, we can think of as specifying binary labels (depending on whether a label Ý is or is not included in ). Similarly, ܵ can be viewed as binary predictions. The Hamming loss then can be regarded as an average of the error rate of on these binary problems. For, let us define for ¾ to be ½ if ¾ ½ if ¾. To simplify notation, we also identify any function À ¾ with a corresponding two-argument function À ½ ½ defined by À Ü µ À ܵ. With the above reduction to binary classification in mind, it is rather straightforward to see how to use boosting to minimize Hamming loss. The main idea of the reduction is simply to replace each training example Ü µ by examples Ü µ µ for ¾. The result is a boosting algorithm called AdaBoost.MH (shown in Figure 2) which maintains a distribution over examples and labels. On round Ø, the weak learner accepts such a distribution Ø (as well as the training set), and generates a weak hypothesis Ø Ê. This reduction also leads to the choice of final hypothesis shown in the figure. The reduction used to derive this algorithm combined with Theorem 1 immediately implies a bound on the Hamming loss of the final hypothesis:

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ½ THEOREM 3 Assuming the notation of Figure 2, the following bound holds for the Hamming loss of À on the training data: hloss Àµ Ì Ø½ Ø We now can apply the ideas in the preceding sections to this binary classification problem. As before, our goal is to minimize Ø Ø µ ÜÔ «Ø Ø Ü µµ (12) on each round. (Here, it is understood that the sum is over all examples indexed by and all labels ¾.) As in Section 3.1, if we require that each Ø have range ½ ½ then we should choose where ½ ÖØ «Ø ½ ¾ ÐÒ ½ Ö Ø Ö Ø This gives Ø (13) Ø µ Ø Ü µ (14) Õ ½ Ö ¾ Ø and the goal of the weak learner becomes maximization of Ö Ø. Note that ½ Ö Ø µ¾ is equal to ÈÖ µø Ø Ü µ which can be thought of as a weighted Hamming loss with respect to Ø. Example. As an example of how to maximize Ö Ø, suppose our goal is to find an oblivious weak hypothesis Ø which ignores the instance Ü and predicts only on the basis of the label. Thus we can omit the Ü argument and write Ø Ü µ Ø µ. Let us also omit Ø subscripts. By symmetry, maximizing Ö is equivalent to maximizing Ö. So, we only need to find which maximizes Ö µ µ µ µ Clearly, this is maximized by setting µ Ò µ

½ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ 7.1. Domain-partitioning weak hypotheses We also can combine these ideas with those in Section 4.1 on domain-partitioning weak hypotheses. As in Section 4.1, suppose that is associated with a partition ½ Æ of the space. It is natural then to create partitions of the form consisting of all sets for ½ Æ and ¾. An appropriate hypothesis can then be formed which predicts Ü µ for Ü ¾. According to the results of Section 4.1, we should choose Ï ½ ¾ ÐÒ Ï È where Ï µ Ü ¾. This gives Õ ¾ Ï Ï (16) 7.2. Relation to one-error and single-label classification We can use these algorithms even when the goal is to minimize one-error. The most natural way to do this is to set À ½ ܵ Ö ÑÜ Ý Ø () «Ø Ø Ü Ýµ (17) i.e., to predict the label Ý most predicted by the weak hypotheses. The next simple theorem relates the one-error of À ½ and the Hamming loss of À. THEOREM 4 With respect to any distribution over observations Ü µ where, one-err À ½ µ hloss Àµ Proof: Assume and suppose À ½ ܵ ¾. We argue that this implies À ܵ. If the maximum in Eq. (17) is positive, then À ½ ܵ ¾ À ܵ. Otherwise, if the maximum is nonpositive, then À ܵ. In either case, À ܵ, i.e., À ܵ ½. Thus, À ½ ܵ ¾ À ܵ which, taking expectations, implies the theorem. In particular, this means that AdaBoost.MH can be applied to single-label multiclass classification problems. The resulting bound on the training error of the final hypothesis À ½ is at most Ø Ø (18)

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ½ where Ø is as in Eq. (12). In fact, the results of Section 8 will imply a better bound of ¾ Ø Ø (19) Moreover, the leading constant ¾ can be improved somewhat by assuming without loss of generality that, prior to examining any of the data, a ¼th weak hypothesis is chosen, namely ¼ ½. For this weak hypothesis, Ö ¼ ¾µ and ¼ is minimized by setting «¼ ½ ¾ ÐÒ ½µ which gives ¼ ¾ Ô ½. Plugging into the bound of Eq. (19), we therefore get an improved bound of ¾ Ì Ø¼ Ô Ì Ø ½ Ø Ø½ This hack is equivalent to modifying the algorithm of Figure 2 only in the manner in which ½ is initialized. Specifically, ½ should be chosen so that ½ Ý µ ½ ¾Ñµ (where Ý is the correct label for Ü ) and ½ µ ½ ¾Ñ ½µµ for Ý. Note that À ½ is unaffected. 8. Using Output Coding for Multiclass Problems The method above maps a single-label problem into a multi-label problem in the simplest and most obvious way, namely, by mapping each single-label observation Ü Ýµ to a multilabel observation Ü Ýµ. However, it may be more effective to use a more sophisticated mapping. In general, we can define a one-to-one mapping ¾ ¼ which we can use to map each observation Ü Ýµ to Ü Ýµµ. Note that maps to subsets of an unspecified label set ¼ which need not be the same as. Let ¼ ¼. It is desirable to choose to be a function which maps different labels to sets which are far from one another, say, in terms of their symmetric difference. This is essentially the approach advocated by Dietterich and Bakiri (199) in a somewhat different setting.they suggested using error correcting codes which are designed to have exactly this property. Alternatively, when ¼ is not too small, we can expect to get a similar effect by choosing entirely at random (so that, for Ý ¾ and ¾ ¼, we include or do not include in ݵ with equal probability). Once a function has been chosen we can apply AdaBoost.MH directly on the transformed training data Ü Ý µµ. How then do we classify a new instance Ü? The most direct use of Dietterich and Bakiri s approach is to evaluate À on Ü to obtain a set À ܵ ¼. We then choose the label Ý ¾ for which the mapped output code ݵ has the shortest Hamming distance to À ܵ. That is, we choose Ö ÑÒ Ýµ À ܵ ݾ A weakness of this approach is that it ignores the confidence with which each label was included or not included in À ܵ. An alternative approach is to predict that label Ý which, if it had been paired with Ü in the training set, would have caused Ü Ýµ to be given the smallest weight under the final distribution. In other words, we suggest predicting the label

½ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ Given: Ü ½ Ý ½ µ Ü Ñ Ý Ñ µ where Ü ¾, Ý ¾ a mapping ¾ ¼ Run AdaBoost.MH on relabeled data: Ü ½ Ý ½ µµ Ü Ñ Ý Ñ µµ Get back final hypothesis À of form À Ü Ý ¼ µ Ò Ü Ý ¼ µµ where Ü Ý ¼ µ «Ø Ø Ü Ý ¼ µ Ø Output modified final hypothesis: (Variant 1) À ½ ܵ Ö ÑÒ Ýµ À ܵ ݾ (Variant 2) À ¾ ܵ Ö ÑÒ Ý¾ Ý ¼ ¾ ¼ ÜÔ ÝµÝ ¼ Ü Ý ¼ µµ Figure 3. AdaBoost.MO: A multiclass version of AdaBoost based on output codes. Ö ÑÒ Ý¾ Ý ¼ ¾ ¼ ÜÔ ÝµÝ ¼ Ü Ý ¼ µµ where, as before, Ü Ý ¼ µ È Ø «Ø Ø Ü Ý ¼ µ. We call this version of boosting using output codes AdaBoost.MO. Pseudocode is given in Figure 3. The next theorem formalizes the intuitions above, giving a bound on training error in terms of the quality of the code as measured by the minimum distance between any pair of code words. THEOREM Assuming the notation of Figure 3 and Figure 2 (viewed as a subroutine), let ÑÒ ½ µ ¾ µ ½ ¾¾ ½ ¾ When run with this choice of, the training error of AdaBoost.MO is upper bounded by ¾ ¼ Ì Ø½ Ø for Variant 1, and by ¼ Ì Ø½ for Variant 2. Ø Proof: We start with Variant 1. Suppose the modified output hypothesis À ½ for Variant 1 makes a mistake on some example Ü Ýµ. This means that for some Ý, À ܵ µ À ܵ ݵ

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ½ which implies that ¾À ܵ ݵ À ܵ ݵ À ܵ µ À ܵ ݵµ À ܵ µµ ݵ µ where the second inequality uses the fact that for any sets and. Thus, in case of an error, À ܵ ݵ ¾. On the other hand, the Hamming error of AdaBoost.MH on the training set is, by definition, ½ Ñ ¼ Ñ ½ À Ü µ Ý µ which is at most É Ø Ø by Theorem 3. Thus, if Å is the number of training mistakes, then Å ¾ Ñ ½ À Ü µ Ý µ Ñ ¼ Ø Ø which implies the stated bound. For Variant 2, suppose that À ¾ makes an error on some example Ü Ýµ. Then for some Ý Ý ¼ ¾ ¼ ÜÔ µý ¼ Ü Ý ¼ µµ Ý ¼ ¾ ¼ ÜÔ ÝµÝ ¼ Ü Ý ¼ µµ () Fixing Ü, Ý and, let us define Û Ý ¼ µ ÜÔ ÝµÝ ¼ Ü Ý ¼ µµ. Note that Û Ý ÜÔ µý ¼ Ü Ý ¼ ¼ µ if ÝµÝ ¼ µý ¼ µµ ½Û Ý ¼ µ otherwise. Thus, Eq. () implies that Ý ¼ ¾Ë Û Ý ¼ µ Ý ¼ ¾Ë ½Û Ý ¼ µ where Ë Ýµ µ. This implies that Ý ¼ ¾ ¼ Û Ý ¼ µ Ý ¼ ¾Ë Û Ý ¼ µ ½ ¾ Ý ¼ ¾Ë Û Ý ¼ µ ½Û Ý ¼ µµ Ë The third inequality uses the fact that Ü ½Ü ¾ for all Ü ¼. Thus, we have shown that if a mistake occurs on Ü Ýµ then Ý ¼ ¾ ¼ ÜÔ ÝµÝ ¼ Ü Ý ¼ µµ If Å is the number of training errors under Variant 2, this means that

¾¼ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ Å Ñ ½ Ý ¼ ¾ ¼ ÜÔ Ý µý ¼ Ü Ý ¼ µµ Ñ ¼ Ø Ø where the equality uses the main argument of the proof of Theorem 1 combined with the reduction to binary classification described just prior to Theorem 3. This immediately implies the stated bound. If the code is chosen at random (uniformly among all possible codes), then, for large ¼, we expect to approach ½¾ Ó ½µµ ¼. In this case, the leading coefficients in the bounds of Theorem approach 4 for Variant 1 and 2 for Variant 2, independent of the number of classes in the original label set. We can use Theorem to improve the bound in Eq. (18) for AdaBoost.MH to that in Eq. (19). We apply Theorem to the code defined by ݵ Ý for all Ý ¾. Clearly, ¾ in this case. Moreover, we claim that À ½ as defined in Eq. (17) produces identical predictions to those generated by Variant 2 in AdaBoost.MO since Ý ¼ ¾ ÜÔ ÝµÝ ¼ Ü Ý ¼ µµ Üݵ Üݵ Ý ¼ ¾ Üݼµ (21) Clearly, the minimum of Eq. (21) over Ý is attained when Ü Ýµ is maximized. Applying Theorem now gives the bound in Eq. (19). 9. Using Ranking Loss for Multiclass Problems In Section 7, we looked at the problem of finding a hypothesis that exactly identifies the labels associated with an instance. In this section, we consider a different variation of this problem in which the goal is to find a hypothesis which ranks the labels with the hope that the correct labels will receive the highest ranks. The approach described here is closely related to one used by Freund et al. (1998) for using boosting for more general ranking problems. To be formal, we now seek a hypothesis of the form Ê with the interpretation that, for a given instance Ü, the labels in should be ordered according to Ü µ. That is, a label ½ is considered to be ranked higher than ¾ if Ü ½ µ Ü ¾ µ. With respect to an observation Ü µ, we only care about the relative ordering of the crucial pairs ¼ ½ for which ¼ ¾ and ½ ¾. We say that misorders a crucial pair ¼ ½ if Ü ½ µ Ü ¼ µ so that fails to rank ½ above ¼. Our goal is to find a function with a small number of misorderings so that the labels in are ranked above the labels not in. Our goal then is to minimize the expected fraction of crucial pairs which are misordered. This quantity is called the ranking loss, and, with respect to a distribution over observations, it is defined to be Ü µ ¼ ½ µ ¾ µ Ü ½ µ Ü ¼ µ We denote this measure rloss µ. Note that we assume that is never empty nor equal to all of for any observation since there is no ranking problem to be solved in this case.

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ¾½ Given: Ü ½ ½ µ Ü Ñ Ñ µ where Ü ¾, Initialize ½ ¼ ½ µ ½ Ñ µ if ¼ ¾ and ½ ¾ ¼ else. For Ø ½ Ì : Train weak learner using distribution Ø. Get weak hypothesis Ø Ê. Choose «Ø ¾ Ê. Update: Ø ½ ¼ ½ µ Ø ¼ ½ µ ÜÔ ½ ¾ «Ø Ø Ü ¼ µ Ø Ü ½ µµ where Ø is a normalization factor (chosen so that Ø ½ will be a distribution). Output the final hypothesis: Ü µ Ì Ø½ «Ø Ø Ü µ Ø Figure 4. AdaBoost.MR: A multiclass, multi-label version of AdaBoost based on ranking loss. A version of AdaBoost for ranking loss called AdaBoost.MR is shown in Figure 4. We now maintain a distribution Ø over ½ Ñ. This distribution is zero, however, except on the relevant triples ¼ ½ µ for which ¼ ½ is a crucial pair relative to Ü µ. Weak hypotheses have the form Ø Ê. We think of these as providing a ranking of labels as described above. The update rule is a bit new. Let ¼ ½ be a crucial pair relative to Ü µ (recall that Ø is zero in all other cases). Assuming momentarily that «Ø ¼, this rule decreases the weight Ø ¼ ½ µ if Ø gives a correct ranking ( Ø Ü ½ µ Ø Ü ¼ µ), and increases this weight otherwise. We can prove a theorem analogous to Theorem 1 for ranking loss: THEOREM 6 Assuming the notation of Figure 4, the following bound holds for the ranking loss of on the training data: rloss µ Ì Ø½ Ø Proof: The proof is very similar to that of Theorem 1. Unraveling the update rule, we have that Ì ½ ¼ ½ ¼ ½ µ ÜÔ ½ Ü ¾ ¼ µ Ü ½ µµ ½ µ É Ø Ø

¾¾ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ The ranking loss on the training set is ½ ¼ ½ µ Ü ¼ µ Ü ½ µ ¼ ½ ¼ ½ ½ ¼ ½ µ ÜÔ ½ ¾ Ü ¼ µ Ü ½ µµ ¼ ½ Ì ½ ¼ ½ µ Ø Ø (Here, each of the sums is over all example indices and all pairs of labels in.) This completes the theorem. So, as before, our goal on each round is to try to minimize ¼ ½ µ ÜÔ ½«Ü ¾ ¼ µ Ü ½ µµ ¼ ½ where, as usual, we omit Ø subscripts. We can apply all of the methods described in previous sections. Starting with the exact methods for finding «, suppose we are given a hypothesis. Then we can make the appropriate modifications to the method of Section 3.2 to find «numerically. Alternatively, in the special case that has range ½ ½, we have that ½ Ü ¾ ¼ µ Ü ½ µµ ¾ ½ ¼ ½ Ø Ø Therefore, we can use the method of Section 3.3 to choose «exactly: «½ ¾ ÐÒ Ï Ï (22) where Ï As before, ¼ ½ ¼ ½ µ Ü ¼ µ Ü ½ µ ¾ (23) Ï ¼ ¾Ô Ï Ï (24) in this case. How can we find a weak hypothesis to minimize this expression? A simplest first case is to try to find the best oblivious weak hypothesis. An interesting open problem then is, given a distribution, to find an oblivious hypothesis ½ ½ which minimizes when defined as in Eqs. (23) and (24). We suspect that this problem may be NP-complete when the size of is not fixed. We also do not know how to analytically find the best oblivious hypothesis when we do not restrict the range of, although numerical methods may be reasonable. Note that finding the best oblivious hypothesis is the simplest case of the natural extension of the

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ¾ technique of Section 4.1 to ranking loss. Folding «¾ into as in Section 4, the problem is to find Ê to minimize ¼ ½ This can be rewritten as ¼ ½ µ ÜÔ ¼ µ ½ µµ ¼ ½ Û ¼ ½ µ ÜÔ ¼ µ ½ µµ (2) È where Û ¼ ½ µ ¼ ½ µ. In Appendix A we show that expressions of the form given by Eq. (2) are convex, and we discuss how to minimize such expressions. (To see that the expression in Eq. (2) has the general form of Eq. (A.1), identify the Û ¼ ½ µ s with the Û s in Eq. (A.1), and the µ s with the s.) Since exact analytic solutions seem hard to come by for ranking loss, we next consider approximations such as those in Section 3.1. Assuming weak hypotheses with range in ½ ½, we can use the same approximation of Eq. (4) which yields where ½ Ö ¾ «½ Ö ¾ «(26) Ö ½ ¾ ¼ ½ µ Ü ½ µ Ü ¼ µµ (27) ¼ ½ As before, the right hand side of Eq. (26) is minimized when which gives «½ ¾ ÐÒ ½ Ö ½ Ö Ô ½ Ö ¾ Thus, a reasonable and more tractable goal for the weak learner is to try to maximize Ö. Example. To find the oblivious weak hypothesis ½ ½ which maximizes Ö, note that by rearranging sums, where Ö µ ½ ¾ µ µ ¼ ¼ µ ¼ µµ Clearly, Ö is maximized if we set µ Ò µµ. Note that, although we use this approximation to find the weak hypothesis, once the weak hypothesis has been computed by the weak learner, we can use other methods to choose «such as those outlined above. (28)

¾ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ Given: Ü ½ ½ µ Ü Ñ Ñ µ where Ü ¾, Initialize Ú ½ µ Ñ µ ½¾ For Ø ½ Ì : Train weak learner using distribution Ø (as defined by Eq. (29)) Get weak hypothesis Ø Ê. Choose «Ø ¾ Ê. Update: where Ø Ú Ø ½ µ Ú Ø µ ÜÔ ½ ¾ «Ø Ø Ü µ Ô Ø ¾¼ ½ Ú Ø µ ÜÔ ½«¾ Ø Ø Ü µ ¾ ¾ ½ Ú Ø µ ÜÔ «¾ Ø Ø Ü µ Output the final hypothesis: Ü µ Ì Ø½ «Ø Ø Ü µ Figure. A more efficient version of AdaBoost.MR (Figure 4). 9.1. A more efficient implementation The method described above may be time and space inefficient when there are many labels. In particular, we naively need to maintain weights for each training example Ü µ, and each weight must be updated on each round. Thus, the space complexity and time-per-round complexity can be as bad as Ñ ¾ µ. In fact, the same algorithm can be implemented using only Ç Ñµ space and time per round. By the nature of the updates, we will show that we only need to maintain weights Ú Ø over ½ Ñ. We will maintain the condition that if ¼ ½ is a crucial pair relative to Ü µ, then Ø ¼ ½ µ Ú Ø ¼ µ Ú Ø ½ µ (29) at all times. (Recall that Ø is zero for all other triples ¼ ½ µ.) The pseudocode for this implementation is shown in Figure. Eq. (29) can be proved by induction. It clearly holds initially. Using our inductive hypothesis, it is straightforward to expand the computation of Ø in Figure to see that it is equivalent to the computation of Ø in Figure 4. To show that Eq. (29) holds on round Ø ½, we have, for crucial pair ¼ ½ : Ø ½ ¼ ½ µ Ø ¼ ½ µ ÜÔ ½ ¾ «Ø Ø Ü ¼ µ Ø Ü ½ µµ Ø

ÁÅÈÊÇÎ ÇÇËÌÁÆ ÄÇÊÁÌÀÅË ¾ Ú Ø ¼ µ ÜÔ ½«¾ Ø Ø Ü ¼ µ Ô Ú ½ Ø ½ µ ÜÔ «¾ Ø Ø Ü ½ µ Ø Ú Ø ½ ¼ µ Ú Ø ½ ½ µ Finally, note that all space requirements and all per-round computations are Ç Ñµ, with the possible exception of the call to the weak learner. However, if we want the weak learner to maximize Ö as in Eq. (27), then we also only need to pass Ñ weights to the weak learner, all of which can be computed in Ç Ñµ time. Omitting Ø subscripts, we can rewrite Ö as where Ö ½ ¾ ½ ¾ ½ ¾ ¼ ½ ¼ ½ µ Ü ½ µ Ü ¼ µµ ¾ Ô Ø ¼¾ ½¾ Ú ¼ µú ½ µ Ü ½ µ ½ Ü ¼ µ ¼ µ ¼¾ ½¾ µ ½ Ú µ ¾ Ú ¼ µ ¼ Ú ½ µ ½¾ Ú ½ µ ½ ¼ Ü ¼ µ Ú ¼ µ ½ Ü ½ µ ¼¾ µ Ü µ () ¼ ¼ Ú ¼ µ All of the weights µ can be computed in Ç Ñµ time by first computing the sums which appear in this equation for the two possible cases that is ½ or ½. Thus, we only need to pass Ç Ñµ weights to the weak learner in this case rather than the full distribution Ø of size Ç Ñ ¾ µ. Moreover, note that Eq. () has exactly the same form as Eq. (14) which means that, in this setting, the same weak learner can be used for either Hamming loss or ranking loss. 9.2. Relation to one-error As in Section 7.2, we can use the ranking loss method for minimizing one-error, and therefore also for single-label problems. Indeed, Freund and Schapire s (1997) pseudoloss - based algorithm AdaBoost.M2 is a special case of the use of ranking loss in which all data are single-labeled, the weak learner attempts to maximize Ö Ø as in Eq. (27), and «Ø is set as in Eq. (28). As before, the natural prediction rule is À ½ ܵ Ö ÑÜ Ý Ø Ü Ýµ

¾ ʺ º ËÀÈÁÊ Æ º ËÁÆÊ in other words, to choose the highest ranked label for instance Ü. We can show: THEOREM 7 With respect to any distribution over observations Ü µ where is neither empty nor equal to, one-err À ½ µ ½µ rloss µ Proof: Suppose À ½ ܵ ¾. Then, with respect to and observation Ü µ, misorderings occur for all pairs ½ ¾ and ¼ À ½ ܵ. Thus, ¼ ½ µ ¾ µ Ü ½ µ Ü ¼ µ ½ ½ ½ Taking expectations gives ½ ½ Ü µ À ½ ܵ ¾ rloss µ which proves the theorem.. Experiments In this section, we describe a few experiments that we ran on some of the boosting algorithms described in this paper. The first set of experiments compares the algorithms on a set of learning benchmark problems from the UCI repository. The second experiment does a comparison on a large text categorization task. More details of our text-categorization experiments appear in a companion paper (Schapire & Singer, to appear). For multiclass problems, we compared three of the boosting algorithms: Discrete AdaBoost.MH: In this version of AdaBoost.MH, we require that weak hypotheses have range ½ ½. As described in Section 7, we set «Ø as in Eq. (13). The goal of the weak learner in this case is to maximize Ö Ø as defined in Eq. (14). Real AdaBoost.MH: In this version of AdaBoost.MH, we do not restrict the range of the weak hypotheses. Since all our experiments involve domain-partitioning weak hypotheses, we can set the confidence-ratings as in Section 7.1 (thereby eliminating the need to choose «Ø ). The goal of the weak learner in this case is to minimize Ø as defined in Eq. (16). We also smoothed the predictions as in Sec. 4.2 using ½ ¾Ñµ. Discrete AdaBoost.MR: In this version of AdaBoost.MR, we require that weak hypotheses have range ½ ½. We use the approximation of Ø given in Eq. (26) and therefore set «Ø as in Eq. (28) with a corresponding goal for the weak learner of maximizing Ö Ø as defined in Eq. (27). Note that, in the single-label case, this algorithm is identical to Freund and Schapire s (1997) AdaBoost.M2 algorithm.