learning feature representations for music...

125
LEARNING FEATURE REPRESENTATIONS FOR MUSIC CLASSIFICATION A DISSERTATION SUBMITTED TO THE DEPARTMENT OF MUSIC AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Juhan Nam December 2012

Upload: doannhan

Post on 01-Aug-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

LEARNING FEATURE REPRESENTATIONS FOR MUSIC

CLASSIFICATION

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF MUSIC

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Juhan Nam

December 2012

http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/jn972gn0355

© 2012 by Juhan Nam. All Rights Reserved.

Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-Noncommercial 3.0 United States License.

ii

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Julius Smith, III, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Jonathan Berger

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Malcolm Slaney

Approved for the Stanford University Committee on Graduate Studies.

Patricia J. Gumport, Vice Provost Graduate Education

This signature page was generated electronically upon submission of this dissertation in electronic format. An original signed hard copy of the signature page is on file inUniversity Archives.

iii

Abstract

In the recent past music has become ubiquitous as digital data. The scale of music

collections in some online music services surpasses ten million tracks. This significant

growth and resulting changes in the music content industry pose challenges in terms

of efficient and effective content search, retrieval and organization. The most common

approach to these needs involves the use of text-based metadata or user data. How-

ever, limitations of these methods, such as popularity bias, have prompted research

in content-based methods that use audio data directly.

The content-based methods are generally composed of two processing modules–

extracting features from audio and training a system using the features and ground

truth. The audio features, the main interest of this thesis, are conventionally designed

in a highly engineered manner based on acoustic knowledge, such as in mel-frequency

cepstral coefficients (MFCCs) or chroma. As an alternative approach, there is increas-

ing interest in learning features automatically from data without relying on domain

knowledge or manual refinement. This feature representation approach has been

studied primarily in the areas of computer vision or speech recognition.

In this thesis, we investigate the learning-based feature representation with appli-

cations to content-based music information retrieval. Specifically, we suggest a data

processing pipeline to effectively learn short-term acoustic dependencies from musical

signals and build a song-level feature for music genre classification and music anno-

tation/retrieval. While visualizing the learned acoustics patterns, we will attempt to

interpret how they are associated with high-level musical semantics such as genre,

emotion or song quality. Through a detailed analysis, we will show the effect of in-

dividual processing units in the pipeline and meta parameters of learning algorithms

iv

on performance. In addition to these tasks, we also examine the feature learning

approach for classification-based piano transcriptions. Throughout experiments on

popularly used datasets, we will show that the learned feature representations achieve

results comparable to state-of-the-art algorithms or outperform them.

v

Acknowledgements

Most of all, I would like to thank my advisor Julius O. Smith, who constantly sup-

ported me and provided a great deal of freedom to explore diverse research areas.

My PhD study was a journey of understanding his in-depth knowledge and wisdom

in DSP and acoustics, and exploring a possibility motivated from future prospects in

his book.1

I also would like to thank Malcolm Slaney for being my mentor throughout this

thesis research. His advice and insight were indispensable in directing experiments

and reasoning. He always supported me in a generous and intimate manner. This

was a huge encouragement to me.

In addition, I would like to thank Jonathan S. Abel. I really enjoyed the white-

board discussions with him. He provided me with interesting ideas and led me to step

up to the next level all the time. I appreciate his friendship and support (particularly

in the last year of my PhD study).

I am also grateful to other CCRMA professors, Jonathan Berger, Chris Chafe,

John Chowning and Ge Wang for guiding me to various aspects of music and giving

inspirations through courses, works and concerts. Especially I give thanks to John

for warm encouragement (I cannot forget the nice dinner with Korean friends at his

home).

The CCRMA community was an excellent environment for my thesis research.

First of all, I significantly benefited from recently upgraded computers “to learn fea-

tures of musical signals”. Thanks to Fernando Lopez-Lezcano and Carr Wilkerson

1https://ccrma.stanford.edu/~jos/sasp/Future_Prospects.html

vi

for their efforts and helps. While spending six years at CCRMA, I have met won-

derful friends: Ed Berdahl, Nick Bryan, Juan Pablo Caceres, Luke Dahl, Rob Hamil-

ton, Jorge Herrera, Blair Bohanan Kaneshiro, Miriam Kolar, Nelson Lee, Gautham

Mysore, Jack Perng, Mauricio Rodriguez and David Yeh. Especially I give thanks to

Nick and Gautham for having fruitful discussions on DSP and machine learning in

addition to being great companions. My thanks extend to Jorge who was my excel-

lent research partner during the last year and incredibly elaborated my thesis work,

especially credited to the real-time visualizer. Also I would like to thank the Korean

community at CCRMA: Hongchan Choi, Song Hui Chon, Yoomi Hur, Hyung-suk

Kim, Keunsup Lee, Kyogu Lee, Jieun Oh, Hwan Shim, Sook-Young Won and Woon

Seung Yeo for their friendship and support.

I also want to thank Honglak Lee at University of Michigan and Jiquan Ngiam in

Andrew Ng’s machine learning group. My thesis research was rooted from a project

with them. Since then, they have provided me with invaluable advice and resources

to conduct my thesis research.

Finally, I thank my parents for their endless love and support. They always

encouraged me and led me to think in a positive way. Also, thanks to my brother

and his family who were proud of me all the time. I am thankful to my adorable two

sons, Patrick and Lyle, for being just as they are. Lastly, I would like to thank my

better half, Hea Jin, who has supported and encouraged me for the last five years.

This thesis would not have been possible without her endless support, patience and

love.

vii

Contents

Abstract iv

Acknowledgements vi

1 Introduction 1

1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Content-based Music Information Retrieval . . . . . . . . . . . . . . . 3

1.2.1 Content-based MIR system . . . . . . . . . . . . . . . . . . . 4

1.3 Feature Representations By Learning . . . . . . . . . . . . . . . . . . 7

1.4 Applications to Music Classification . . . . . . . . . . . . . . . . . . . 10

1.5 Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2 Feature Learning Framework 13

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.2 Previous Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.3 Proposed Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3.1 Preprocessing . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.3.2 Feature Learning: Common Properties . . . . . . . . . . . . . 23

2.3.3 Feature Learning: Algorithms . . . . . . . . . . . . . . . . . . 27

2.3.4 Feature Summarization . . . . . . . . . . . . . . . . . . . . . . 33

2.3.5 Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.4 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.4.1 Datasets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.4.2 Preprocessing Parameters . . . . . . . . . . . . . . . . . . . . 37

viii

2.4.3 Feature-Learning Parameters . . . . . . . . . . . . . . . . . . 38

2.4.4 Classifier Parameters . . . . . . . . . . . . . . . . . . . . . . . 38

2.5 Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

2.5.1 Visualization . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

2.5.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . 41

2.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3 Music Annotation and Retrieval 52

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

3.2 Previous Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.3 Proposed Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3.3.1 Single Layer Model . . . . . . . . . . . . . . . . . . . . . . . . 56

3.3.2 Extension to Deep Learning . . . . . . . . . . . . . . . . . . . 57

3.4 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

3.4.1 Datasets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

3.4.2 Preprocessing Parameters . . . . . . . . . . . . . . . . . . . . 60

3.4.3 MFCC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

3.4.4 Feature-Learning Parameters . . . . . . . . . . . . . . . . . . 61

3.4.5 Classifier Parameters . . . . . . . . . . . . . . . . . . . . . . . 61

3.5 Evaluation and Discussion . . . . . . . . . . . . . . . . . . . . . . . . 62

3.5.1 Visualization . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

3.5.2 Evaluation Metrics . . . . . . . . . . . . . . . . . . . . . . . . 62

3.5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . 64

3.6 Conclusion and Future Work . . . . . . . . . . . . . . . . . . . . . . . 69

4 Piano Transcription Using Deep Learning 71

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

4.2 Previous Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.3 Proposed Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.3.1 Feature Representation By Deep Learning . . . . . . . . . . . 74

4.3.2 Training Strategy . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.3.3 HMM Post-processing . . . . . . . . . . . . . . . . . . . . . . 77

ix

4.4 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.4.1 Datasets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.4.2 Pre-processing . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.4.3 Unsupervised Feature Learning . . . . . . . . . . . . . . . . . 81

4.4.4 Evaluation Metrics . . . . . . . . . . . . . . . . . . . . . . . . 82

4.4.5 Training Scenarios . . . . . . . . . . . . . . . . . . . . . . . . 82

4.5 Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

4.5.1 Validation Results . . . . . . . . . . . . . . . . . . . . . . . . 83

4.5.2 Test Results: Comparison With Other Methods . . . . . . . . 84

4.6 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . 86

5 Conclusions 88

5.1 Contributions and Reviews . . . . . . . . . . . . . . . . . . . . . . . . 88

5.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

A Real-time Music Tagging Visualizer 92

A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

A.2 Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

A.3 Implementation Details . . . . . . . . . . . . . . . . . . . . . . . . . . 94

B Supplementary Materials 96

Bibliography 97

x

List of Tables

2.1 Comparison of different feature-learning algorithms on the GTZAN

genre dataset. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

2.2 Comparison of different feature-learning algorithms on the ISMIR2004

genre dataset. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

2.3 Confusion Matrix for ten genres for the best algorithm (sparse coding

gives 89.7% accuracy) on the GTZAN gene dataset. . . . . . . . . . . 48

2.4 Comparison with state-of-the-art algorithms on the GTZAN genre

dataset . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

2.5 Comparison with state-of-the-art algorithms on the ISMIR2004 genre

dataset . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

3.1 Examples of music annotation. The natural language template was

borrowed from Turnbull [103]. The words in bold are the annotation

output generated by our system. . . . . . . . . . . . . . . . . . . . . . 53

3.2 Example of text-query based music retrieval. These are retrieval out-

puts generated by our system. . . . . . . . . . . . . . . . . . . . . . 54

3.3 This table describes the acoustic patterns of the feature bases that

are actively “triggered” in songs with a given tag. The corresponding

feature bases are shown in Figure 3.2. . . . . . . . . . . . . . . . . . 64

3.4 Performance comparison for different input data and feature-learning

algorithms. These results are all based on linear SVMs. . . . . . . . . 65

xi

3.5 Performance comparison for linear SVM and neural networks with ran-

dom initialization (Mel-SRBM-NN*) and pre-training by DBN (Mel-

SRBM-DBN*). The figures (1, 2 and 3) indicate the number of hidden

layers. The receptive field size was set to 6 frames. . . . . . . . . . . 68

3.6 Performance comparison: state-of-the-art (top) and proposed methods

(bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.1 Frame-level accuracy on the Poliner and Ellis, and Marolt test set. The

upper group was trained with the Poliner and Ellis train set while the

lower group was with other piano recordings or uses different methods.

S1 and S2 refer to training scenarios. †These results are from Poliner

and Ellis [85]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

4.2 Frame-level accuracy on the MAPS test set in F-measure. “ft” stands

for fine-tuned. †These results are from Vincent et. al. [107]. . . . . . 87

xii

List of Figures

1.1 Data processing pipeline in content-based MIR . . . . . . . . . . . . . 5

1.2 Examples of hand-engineered audio features: MFCC and Chroma. The

figures on the left compare spectrogram, mel-frequency spectrogram,

MFCCs and reconstructed spectrogram from the MFCCs in the order

of top to bottom. Note that the MFCCs extract spectral envelope while

removing harmonic patterns. The figures on the right, borrowed from

[70], show a filter bank output (from 88 bandpass filters whose center

frequencies correspond to piano notes) and two versions of chroma

features formed by projecting the filter bank output onto 12 pitch

classes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.3 Statistically learned basis functions from natural sounds are compared

to physiologically measured responses. Each basis function (colored

in red) is overlaid on a measured impulse response obtained from cat

auditory nerve fibers (colored in blue). The original figure is from

Smith and Lewicki [96]. . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.4 Feature representations learned from a music dataset. (a) Feature bases

learned with a sparse RBM (b) mel-frequency spectrogram (left) and

encoded feature representations with regards to the learned feature

bases (right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.1 Proposed data processing pipeline for feature learning . . . . . . . . 18

xiii

2.2 Effect of a time-frequency Automatic Gain Control (AGC). The sub-

band envelopes obtained from the original spectrogram are remapped

to linear frequency (middle). This is used to equalize the original spec-

trogram (bottom). Note that the high-frequency content is boosted by

the AGC in the output spectrogram. . . . . . . . . . . . . . . . . . . 20

2.3 Comparison of (equalized) spectrogram and mel-frequency spectro-

gram. 513 bins in spectrogram (FFT size is 1024) are mapped to

128 bins in mel-frequency spectrogram. . . . . . . . . . . . . . . . . . 22

2.4 Mel-frequency spectrogram (top) and PCA-whitened mel-frequency

spectrogram (bottom). The figure indicates that 4 frames in the mel-

frequency spectrogram are chosen as a receptive field size and projected

to a single vector in the PCA space. The PCA indices are sorted such

that components with high eigenvalues have lower index numbers. . . 25

2.5 Undirected graphical model of a RBM. . . . . . . . . . . . . . . . . . 32

2.6 Comparison of mel-frequency spectrogram and the learned feature rep-

resentation (hidden layer activation) using sparse RBM. The dictionary

size was set to 1024 and target activation (sparsity) was to 0.01. That

means that only 1% out of 1024 features are activated on average. . 34

2.7 Comparison of hidden layer activation and its max-pooled version. For

visualization, the hidden layer features masked by the maximum value

in their pooling region were set to zero. As a result, the max-pooled

feature maintains only locally dominant activation. The hidden layer

feature representation was learned using sparse RBM. Dictionary size

was set to 1024, sparsity was to 0.01 and max-pooling size was to 43

frames (about 1 second). . . . . . . . . . . . . . . . . . . . . . . . . 36

2.8 Top 20 most active feature bases (dictionary elements) for ten genres

at the GTZAN set. (a) Sparse RBM with 1024 hidden units and 0.03

sparsity (b) Sparse coding with 1024 dictionary elements and λ=1.5. . 40

2.9 Comparison of spectrogram and mel-frequency spectrogram. The dic-

tionary size is set to 1024. . . . . . . . . . . . . . . . . . . . . . . . . 41

2.10 Effect of Automatic Gain Control. The dictionary size is set to 1024. 42

xiv

2.11 Effect of receptive field size. The dictionary size is set to 1024. . . . 43

2.12 Effect of dictionary size. The receptive field size is set to 4 frames. . 44

2.13 Effect of sparsity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

2.14 Effect of max-pooling . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3.1 Feature-learning architecture using deep learning for multi-labeling

classification. A deep belief network is used on top of the song-level

feature vectors and then the network is fine-tuned with the tag labels. 59

3.2 Top 20 most active feature bases (dictionary elements) learned by a

sparse RBM for different emotions, vocal quality, instruments and us-

age categories of the CAL500 set. . . . . . . . . . . . . . . . . . . . . 63

3.3 Effect of number of frames. The dictionary size is set to 1024. . . . . 66

3.4 Effect of sparsity (sparse RBM). . . . . . . . . . . . . . . . . . . . . . 67

3.5 Max-pooling of sparsity (sparse RBM). . . . . . . . . . . . . . . . . . 67

4.1 Classification-based polyphonic note transcription. Each binary clas-

sifier detects the presence of a note. . . . . . . . . . . . . . . . . . . 72

4.2 Feature bases learned from a piano dataset using a sparse RBM. They

were sorted by the frequency of the highest peak. Most bases capture

harmonic distributions which correspond to various pitches while some

contain non-harmonic patterns. Note that the feature bases show ex-

ponentially growing curves for each harmonic partial. This verifies the

logarithmic scale in the piano sound. . . . . . . . . . . . . . . . . . . 75

4.3 Network configurations for single-note and multiple-note training. Fea-

tures are obtained from feed-forward transformation as indicated by

the bottom-up arrows. They can be fine-tuned by back-propagation as

indicated by the top-down arrows. . . . . . . . . . . . . . . . . . . . . 78

4.4 Signal transformation through our system. The original spectrogram

gradually changes to the final output via the deep network, SVM clas-

sifiers and HMM-based smoothing . . . . . . . . . . . . . . . . . . . . 80

4.5 Frame-level accuracy on validation sets in two scenarios. The first and

second-layer DBN features are referred to as L1 and L2. . . . . . . . . 84

xv

4.6 Onset accuracy on validation sets in two scenarios. . . . . . . . . . . 85

4.7 Frame-level accuracy VS sparsity (hidden layer activation in RBMs) . 85

A.1 A screen shot of music tagging visualizer . . . . . . . . . . . . . . . . 93

A.2 Diagram of software architecture . . . . . . . . . . . . . . . . . . . . 94

xvi

Chapter 1

Introduction

1.1 Background

The music industry has dramatically evolved over the last decade. Physical media

such as CDs are already crowded out by MP3s and now compressed audio files are

being replaced by readily accessible audio streaming, rendering music indeed ubiqui-

tous. Along with the transition of audio formats, the scale of music content also has

rapidly increased by online music and other content services. For example, Last.fm

has more than 12 million audio tracks available from artists on all the major com-

mercial and independent labels.1 On Youtube, 72 hours of video data is uploaded

every minute,2 where music content is much of it. In addition, the advent of social

media services has promoted diversity in music content by enabling people to share

their own original music or cover songs, and facilitating easy access to various types

of music contents such as music video or live performance video at concerts or TV

stations. These significant changes in the music industry have prompted new strate-

gies to deliver music content, for example, by having a large volume of music content

searchable by different query methods (e.g., text, humming, music track, etc.) or

providing personalized music recommendation.

1http://www.last.fm, accessed in May, 20122http://www.youtube.com/t/press_statistics, accessed in Aug, 2012

1

CHAPTER 1. INTRODUCTION 2

Online music service providers and researchers in the Music Information Retrieval

(MIR) community have approached this problem in a variety of ways. The most

common approach is using textual metadata. A music track is usually described with

artist and track information. In addition to the basic text data, some social music ser-

vices allow users to specify tags that describe music in multi-faceted contexts. Other

music services such as Pandora have trained music experts create more sophisticated

music analysis data.3 This rich text data contains diverse categorical information

which can be used for organizing a large collection of music or measuring artist or

music similarity for advanced music services. Another type of metadata is user data,

for example, play history, song rating and personal music library. Collaborative fil-

tering using the user preference data is known to be particularly effective in music

recommendation [14].

Most music services rely on these types of metadata only and have succeeded to

some degree [13]. However, these approaches have limitations, which progressively

stand out as the scale of music content increases and the types become diverse. The

most significant one is that, while some popular artists or songs have highly rich meta

data, the majority of music are insufficiently annotated or accessed. This is often

called popularity bias or demonstrated as a long-tail distribution. This causes the

“cold start” problem, which refers to failing to retrieve unused or rarely annotated

items [89]. Another weakness is that textual metadata is often created by many

people so that it can be difficult to maintain consistency and accuracy. Although

expert analysis data can be an alternative, they are somewhat limited and costly.

In addition to the metadata, great efforts have been made to use the content itself

to help searching or organizing music [109, 13]. The content-based approach is usually

conducted by building a system that predicts certain types of meta data from audio

data. The major advantage of this approach is that it can be used for unlabeled music

content. That is, the system can automatically annotate unlabeled songs or measure

song similarity based on acoustic information only. Also, as long as there is computing

power, the system can perform such tasks in a fast and inexpensive manner. These

3Trained music analysts in Pandora annotate each song with up to 450 musical attributes. Seehttp://www.pandora.com/about/mgp, accessed in Sep, 2012

CHAPTER 1. INTRODUCTION 3

merits of the content-based approach, however, assume that the system is reliable.

For example, the content-aware system should be comparable to human judgment

or satisfy certain evaluation criteria in terms of accuracy. This requirement poses an

intriguing research challenge, that is, building an intelligent machine that understands

music as humans do.

Recently there was an argument the this content-based approach has drawbacks

over “human signals” such as user ratings when it comes to measuring similarity

between multimedia content or making recommendations on it [94]. This is undeniable

to some extent when considering highly subjective and context-sensitive aspects of

music and other media such as images or movies. For example, “indie music” can

be hardly identified by the content. However, the content-based approach is not

exclusive to the metadata approach but can compensate for it, for example, regarding

the popularity bias problem. Inversely, as we collect more human signals, they can

be used to improve the content-based approach by providing more ground truth.

This thesis addresses the problem of music search and organization with the

content-based approach, particularly focusing on music classification. The follow-

ing sections will detail the content-based approach and then introduce the motivation

of this thesis.

1.2 Content-based Music Information Retrieval

Sound is a physical phenomenon often described as oscillatory changes in pressure.

As sound comes into our ears, however, it evokes a vast array of sensations. The

stimulus first causes tonotopic responses along the cochlear, activates multiple lay-

ers of neurons along the auditory pathway and finally elicits high-level abstractions.

As a consequence, what we normally extract from a sound is information that the

sound contains rather than quantities of it physical characteristics. For example, we

recognize words or emotions from speech sounds, find melody, rhythm or genres from

musical sounds and identify types or locations from environmental sounds.

People have long been interested in enabling computers to understand sounds as

humans do. Although the majority of research efforts has been made for speech, there

CHAPTER 1. INTRODUCTION 4

have been efforts to handle this topic for other sounds, calling it machine listening

or machine hearing as an analogy for computer vision or machine vision [22, 60].

Specifically, Ellis addressed machine listening as a general machine perception area

specific to sound that includes speech, music and environmental ones [22]. Lyon

described the area with the term, machine hearing, particularly focusing on a cochlear

filter model as a front-end sound processor. Content-based MIR is a category that

mainly handles musical sounds [109, 13].

Among all sorts of sounds, music is distinguished by many aspects. First, music

is highly structured. For example, musical notes are usually arranged under a tuning

system, harmonic rules, tempo and rhythms. Also, most songs are composed with a

musical form that describes the layout of them. Second, multiple sound sources (i.e.,

instruments) are played simultaneously in music. In addition, every instrument has

different characteristic timbre, register and expressiveness. Third, music is dynamic.

Musical signals often reach the lower and upper bounds of human hearing range in

loudness and frequency. Lastly, music stirs up various emotional and aesthetic states

and thus has been involved with social interaction and other types of art in different

cultural settings. Hence, music is associated with various types of information from

symbolic music representations (e.g., melody, note and instrumentation) to high-level

categorical concepts in different contexts (e.g., genre, emotion and usage). Also, musi-

cal signals require different methods of processing from those developed for speech or

other sounds. Techniques and strategies for analyzing musical signals and extracting

such information are the main concerns of content-based MIR.

1.2.1 Content-based MIR system

A practical content-based MIR system is usually designed to perform a specific task,

that is, extract certain information. Therefore, every system use different tools and

algorithms to process the musical signal and associate it with desired information [69].

However, their key data processing stages, the pathway from sound to information,

is common to most systems, as summarized in Figure 1.1.

CHAPTER 1. INTRODUCTION 5

Sound Time-frequencytransform

Learning Algorithms InformationFeature

Extraction

Figure 1.1: Data processing pipeline in content-based MIR

The first key stage is the front-end processor that performs time-frequency trans-

form. Sound is acquired as a waveform in the time domain from a sensor (e.g., micro-

phones) in real-time or stored sound media (e.g., audio files or sound tracks in video

files). The majority of content-based MIR systems convert the waveform to a time-

frequency representation, for example, by Fourier transform, constant-Q transform

or other filter banks. Since these transforms have more correlated basis functions to

sound than the waveform, they provide more interpretable representations of sound.4

The time-frequency representations are, however, still complex and high-

dimensional so that they usually contain more acoustic variations than necessary;

some of which are essential to performing a desired task whereas others are redun-

dant or interfering. Therefore it is desirable to extract only features relevant to the

task in a succinct form. Thus feature extraction is the second key stage in content-

based MIR system and reduces audio content for processing by subsequent learning

algorithms, which we refer to as feature representations.

Popularly used feature representations in MIR are divided into two categories. One

category summarizes statistical characteristics from spectrogram. These are used as

timbral features (e.g., spectral centroid, roll-off, kurtosis, etc.) or an onset detection

indicator (e.g., spectral flux). The other category represents acoustic characteristics

in an elaborate way. The most commonly used are mel-frequency Cepstral Coefficients

(MFCCs) and Chroma. The MFCC was originally developed for speech recognition

4Waveforms can be represented as a weighted sum of shifted impulses in discrete-time domain.Thus, shifted impulses can be seen as the basis functions of waveforms. On the other hand, time-frequency representations use some form of a sinusoid with linearly or logarithmically scaled fre-quencies, as basis functions. Since sound is by nature described by periodicity, the sinusoidal basisfunctions are more correlated to sound than the shifted impulses.

CHAPTER 1. INTRODUCTION 6

(a) MFCC (b) Chroma

Figure 1.2: Examples of hand-engineered audio features: MFCC and Chroma. Thefigures on the left compare spectrogram, mel-frequency spectrogram, MFCCs andreconstructed spectrogram from the MFCCs in the order of top to bottom. Notethat the MFCCs extract spectral envelope while removing harmonic patterns. Thefigures on the right, borrowed from [70], show a filter bank output (from 88 bandpassfilters whose center frequencies correspond to piano notes) and two versions of chromafeatures formed by projecting the filter bank output onto 12 pitch classes.

as a method to extract vocal formants but is also frequently used to capture the

spectral envelopes of musical signals as a timbral feature. Chroma is a music-specific

feature. It extracts harmonic information by projecting spectral energy to 12 pitch

classes. These 12 pitch classes correspond to the notes in a western octave. Figure

1.2 illustrates these two audio features.

The last key stage is the learning algorithm. In general, supervised machine-

learning algorithms are used to convert the extracted audio features and available

ground truth so that the system makes predictions for new sound. Support vector

CHAPTER 1. INTRODUCTION 7

machine and neural networks are widely used for classification. Hidden Markov mod-

els are frequently chosen when temporal dependency needs to be considered. On

the other hand, some similarity-based tasks train the system in an unsupervised set-

ting. They first measure distances between two feature distributions (e.g., cosine, L1,

L2 and KL-divergence) and apply nearest neighbor algorithms for classification or

similarity-based music search.

1.3 Feature Representations By Learning

The data processing pipeline in Figure 1.1 suggests that the success of content-based

MIR systems is determined by a combination of good feature representations and

appropriate learning algorithms. In particular, the former is important because good

features facilitate learning in the next step, for example, by making different classes

of musical information easily separable (e.g., linearly separable) in the feature space.

Conventionally, feature representations used in content-based MIR systems are

engineered by relying on human efforts, using domain knowledge such as acoustics or

audio signal processing.5 For example, the aforementioned MFCC has been crafted

based on psychoacoustic observations (e.g., logarithmic perception in frequency or

amplitude) and audio signal processing techniques in the speech recognition commu-

nity [11, 20, 41, 93, 25]. On the other hand, Chroma was designed based on musical

acoustics (e.g., musical tuning) by the MIR community [30, 108, 24, 70]. In addition,

there are a number of engineered audio features used in content-based MIR, such

as auditory filterbank temporal envelopes (AFTE) [66], octave-based spectral con-

trast (OSC) [46], Daubechies wavelet coefficient histogram (DWCH) [58], auditory

temporal modulation [82] and so forth.

Although these hand-engineered features are used extensively and proven to be

effective, this approach has limitations as well. First, the features are usually crafted

through time-consuming human refinement, specifically numerous trials and errors.

5This approach is common to most machine perception tasks including speech recognition orcomputer vision. For example, widely used computer vision features, such as Scale-Invariant FeatureTransform (SIFT) and Histogram of Oriented Gradient (HOG), were also hand-engineered based ondomain-specific knowledge.

CHAPTER 1. INTRODUCTION 8

Figure 1.3: Statistically learned basis functions from natural sounds are compared tophysiologically measured responses. Each basis function (colored in red) is overlaidon a measured impulse response obtained from cat auditory nerve fibers (colored inblue). The original figure is from Smith and Lewicki [96].

Second, the features require expert domain knowledge or often rely on ad-hoc ap-

proaches. As discussed above, some of hand-engineered features heavily rely on spe-

cific acoustic knowledge. This makes it difficult to use them in more general-purpose

systems.

As an alternative to the hand-engineering approach, there has been increasing

interest in learning the feature representations. Instead of designing computational

steps that extract specific aspects of data, a new approach develops general-purpose

algorithms that automatically learn feature representations from data. The under-

lying idea is that if data has some structures, the algorithm will find a set of basis

functions that explain the structures and thus better represent any example of the

data. This approach was originally inspired by computational neuroscience.

One of the primary goals in computational neuroscience is to understand human

perception mechanism using computational models. This was usually tackled by re-

verse engineering, that is, mimicking the functionality of physiological units in compu-

tational models. Alternatively, some computational neuroscientists questioned on an

underlying theoretical principle of human sensory systems, that is, why human sensory

system has the very specific structure and characteristics–for example, why does the

CHAPTER 1. INTRODUCTION 9

auditory system have cochlea filters with specific frequency responses, among many

other choices? They attempted to answer the question using information-theoretic

views.

Barlow first proposed a hypothesis that the human sensory system encodes incom-

ing signals so that the information contained in the signals is maximally extracted

with minimum resources. This is the principle of redundancy reduction [3]. This

efficient sensory coding was further investigated using statistical learning algorithms

by Olshausen and Field. They introduced a sparse coding algorithm that encodes a

sensory signal into a sparse representation by representing the signal as a linear com-

bination of basis functions with a sparsity constraint [12, 78]. They showed that basis

functions learned from natural images resembled the characteristics of receptive fields

in the primary visual cortex, for example, edge detectors at different location and

angles. In the auditory domain, Lewicki and Smith applied sparse coding into natu-

ral sounds and speech [57, 96]. They demonstrated that the basis functions learned

from data are very similar to the impulse responses of mammalian cochlear filters, as

shown in Figure 1.3. The key idea the visual and auditory work suggest is that sen-

sory processing is adapt to the stimuli under a general computational principle. That

it, the incoming data is encoded in a sparse way and thus our brains use minimum

energy to recognize it (i.e., spikes).

Machine-learning researchers figured out that this data representation scheme is

useful for machine recognition as well and have developed various representational

algorithms in machine-learning context. This area of machine learning is often called

unsupervised feature learning, or deep learning when a multi-layer structure is used

[73]. They demonstrated that the feature-learning algorithms automatically discovers

various structures in image, video, speech and music data, for example, edges, corners

and shapes from natural images [53, 54], timbral patterns or phonemes from speech

[56, 55] and harmonic patterns from music [95, 1, 9]. Furthermore, they showed

that the learned representations outperform the hand-engineered features in several

benchmark tests [54, 55, 16, 17].

In addition, this learning approach has a couple of practical advantages. First,

since they are unsupervised, the learning algorithms do not require labels. Unlabeled

CHAPTER 1. INTRODUCTION 10

data is much easier and cheaper than labeled data. Second, they are universal feature

learners. That is, they can be applied to any type of data, whether it is image, speech

or music. Thus we can minimize use of domain specific knowledge and considerations.

Third, new feature representations can be developed quickly. Thus the time for feature

learning time depends only on the amount of training data and computing power.

1.4 Applications to Music Classification

Music can be described by a variety of semantics such as mood, emotion, genre

and artist. Such high-level concepts are often determined by fundamental elements

in music: timbre, pitch, harmony and rhythm. Thus, numerous efforts have been

directed at developing feature representations that captures the musical elements of

audio. In the past, the audio features were usually hand-crafted relying on acoustic

knowledge or signal processing techniques. In this thesis, we will apply learning

algorithms to find new feature representations, instead. Throughout the following

chapters, we will focus on two issues:

• Can the learning algorithms capture meaningful and rich acoustic patterns?

In other words, are the learned features interpretable as musical elements and

furthermore can they be associated with musical semantics and categories?

• Do the learned feature representations perform well against hand-engineered

audio features, or other feature-learning approaches, in practical music classifi-

cation tasks?

To this end, we will present a data processing pipeline to effectively learn feature

representation and evaluate it on publicly available datasets. We will examine the

pipeline parameters that affect the feature patterns and classification performance.

Strategically, we will leverage successful work in other domains, particularly computer

vision. Hence, we will focus on adapting well-known feature-learning algorithms to

the music domain rather than developing new algorithms (Note that the learning

algorithms can be applied to any type of data).

CHAPTER 1. INTRODUCTION 11

Mel

−fre

quen

cy b

in

20

40

60

80

100

120

(a) Feature bases

Time [frame]

Fre

quency [kH

z]

50 100 150 200

20

40

60

80

100

120

Time [frame]

Dic

tionary

Index

50 100 150 200

200

400

600

800

1000

0.5

1

1.5

2

2.5

3

3.5

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

(b) Input data (Left) and feature representation (right)

Figure 1.4: Feature representations learned from a music dataset. (a) Feature baseslearned with a sparse RBM (b) mel-frequency spectrogram (left) and encoded featurerepresentations with regards to the learned feature bases (right)

As a preview, Figure 1.4 shows feature bases learned by an unsupervised learning

algorithm and a new feature representation shown as high-dimensional sparse activa-

tion. The feature bases extract rich acoustic patterns associated with musical signals.

This will be described in the next chapter.

1.5 Organization

This thesis has five chapters. After this introduction, Chapter 2 presents a framework

to learn audio features for music classification. Specifically, we propose a data pro-

cessing pipeline that effectively captures a range of acoustic patterns by unsupervised

CHAPTER 1. INTRODUCTION 12

algorithms and construct a song-level feature. Also we introduce feature-learning

algorithms and compare them. Finally, we evaluate the framework on music genre

datasets, investigating the effect of individual modules in the pipeline. Chapter 3

applies the same framework to music annotation and retrieval tasks. These are more

general music classification problem. Since this task provides various descriptions of

music, we can interpret the learned features as richer musical semantics. Additionally,

we adopt a deep learning algorithm to improve classification performance. Chapter

4 presents classification-based polyphonic piano transcription as another application.

This is also based on feature learning with a deep structure. Finally, Chapter 5 con-

cludes this thesis by summarizing the presented content and discussing future work.

Chapter 2

Feature Learning Framework for

Music Classification

2.1 Introduction

Music classification refers to tasks that make predictions of a label (e.g., genre, artist

and mood classification) given a musical signal. As the size of music collections

in online music services has increased and the types become more varied by virtue

of social media services, labeling the music content has become an important need

for music browsing, search or recommendation. However, the majority of the music

content are insufficiently annotated due to popularity bias, or inconsistently labeled

by online crowds. This has led the necessity of automatically analyzing and classifying

music using the audio content, thereby placing content-based music classification as

one of the key problems in MIR.

In general, the content-based music-classification system is composed of two pieces.

The first piece extracts salient descriptions of timbre (or instrumentation), pitch (or

harmony) and rhythm from audio signals to characterize the music. The second

piece of the system performs supervised training to associate the features with high-

level musical labels, and then classifies new songs into one or multiple categories.

The majority of previous music classification systems have focused on developing

good musical features while choosing widely used classifiers, for example, k-nearest

13

CHAPTER 2. FEATURE LEARNING FRAMEWORK 14

neighbor, support vector machine and Gaussian mixture model.

For example, Tzanetakis and Cook presented various signal processing methods

to extract timbre, rhythm and pitch features in the seminal work on music genre

classification [104]. Specifically, they suggested spectral centroid/roll-off/flux, time-

domain zero-crossings, MFCC, low-energy features for timbre, a wavelet transform-

based beat histogram for rhythm, and outcomes of a multi-pitch detection algorithm.

They additionally considered long-term features that summarize short-term timbral

features using mean and variance over a “texture” window. McKinney and Breebaart

investigated feature representations in the auditory domain [66]. They proposed psy-

choacoustic features based on estimates of the perception of roughness, loudness and

sharpness, and auditory filterbank temporal envelopes using gamma-tone filters. Sim-

ilarly, a number of novel audio features have been developed based on different choices

of time-frequency representations, psychoacoustic adaptation and long-term summa-

rizations. They include octave-based spectral contrast [46], Daubechies wavelet co-

efficient histogram [58], stereo panning spectrum features [105], auditory temporal

modulations [82] and so on.

While these methods extract important audio features in novel ways, the under-

lying approach is that the features are engineered by human efforts. In other words,

they use hand-crafted features based on acoustic knowledge and signal processing

techniques. Although these hand-crafted features are invaluable and are successfully

used in many music classification tasks, this approach has drawbacks. First, the fea-

tures are tuned through numerous trials so that their development is time-consuming.

Second, the approach often relies on highly complicated domain knowledge on an ad-

hoc basis. This hinders the features from being used for other purposes. Similar

problems were addressed in other domains such as computer vision [52].

As an alternative to this approach, there has been increasing interest in learning

features automatically from data. That is, instead of extracting features with humans

domain knowledge, a new approach discovers the features using a learning algorithm.

The idea is that if data has structure, the algorithm can learn a set of basis functions

and encode any example of the data into a feature vector using the basis functions.

Researchers have shown that these learning algorithms not only discover underlying

CHAPTER 2. FEATURE LEARNING FRAMEWORK 15

structures from image or audio but also provide new features representations that can

substitute for hand-crafted features [54, 16, 17].

This learning approach has been recently applied to music domain as well, par-

ticularly, music genre classification [55, 32, 35, 21, 90]. In general, researchers con-

structed different learned features by choosing different time-frequency transforms of

input data, feature-learning algorithms and further summarization techniques for long

sequences of data. While they have incrementally showed better results in classifi-

cation accuracy over hand-engineered features, few studies conducted comprehensive

analysis on the various factors of the data processing pipeline. This includes data

normalization, selection of input size and meta parameters in feature learning al-

gorithms. Also, while most visualized the learned feature bases and attempted to

interpret them in the view of acoustics, they did not provide deep insight on how the

new features are associated with high-level categorical concepts.

In this chapter, we propose a data processing pipeline to effectively learn features

for music classification. We analyze individual modules by visualizing the interme-

diate data representations along the pipeline. Also, we examine in details how each

processing module or it parameters affect classification performance. Our methods

are evaluated on two public datasets, GTZAN and ISMIR2004, described in Section

2.4.1. Our experiment shows that our method achieves 89.7% accuracy on GTZAN

and 86.8% on ISMIR2004. This is higher than previously reported efforts using feature

learning. In addition to the performance evaluation, we illustrate the learned features

by associating them with high-level categorical concepts using statistical means.

The reminder of this chapter is organized as follows. In Section 2.2, we review

previous feature learning work for music classification. In Section 2.3, we propose a

data processing pipeline for feature learning and describe each processing unit and

the feature learning algorithms. In Section 2.4, we explain our experiment detailing

datasets and parameters. In Section 2.5, we show the results and discuss the effect

of parameters in the data processing pipeline. Finally we wrap up this chapter in

Section 2.6.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 16

2.2 Previous Work

Recently feature learning has of great interest to the MIR community. Researchers

highly leveraged recent development of feature learning algorithms in machine learn-

ing and computer vision. The algorithms that they used can be divided into two

groups depending on the number of feature layers. One group pursued hierarchi-

cal feature representations using multi-layer algorithms. This is often called deep

learning. The other group focused on a high-dimensional sparse representation using

single-layer algorithms.

The most common deep learning approach is a Deep Belief Networks (DBNs). This

is a multi-layer model built by greedy layer-wise training of a number of single-layer

learning modules called Restricted Boltzmann Machines (RBMs) (see Section 2.3.3 for

details) [36]. Hamel et. al. applied the DBN to several content-based MIR tasks such

as musical instrument identification, music genre and tag classification [34, 32]. They

utilized a DBN mainly for “pre-training” deep neural networks. For example, they

attempted to find a nonlinear mapping between spectrogram and musical genres using

frame-level DBN training and then “fine-tuning” the network with genre labels. They

used the top hidden-layer representation as novel audio features (supervised by genres)

and fed these features into SVM with the RBF kernel, achieving 84.3% accuracy on

the GTZAN dataset. A similar approach was applied to music emotion recognition

as well by Schmidt et. al. [91, 92]. In order to model emotion in a continuous space,

they instead attached a regression algorithm to the top of the DBN. They showed

that DBN-based features were superior to MFCC or other hand-engineered features.

While DBNs were successful in controlled domains, for example, classifying single

frames of spectrogram, they have limitations when scaling to a high-dimensional data

such as multiple frames or song segments. Lee et. al. proposed Convolutional Deep

Belief Networks (CDBN) in order to improve their scalability [54]. In CDBNs, each

layer learns features from lower layers at a different scale using probabilistic max-

pooling. They applied their network to various audio classification tasks such as

music genre and artist classification and showed promising results [55]. Dieleman et.

al. [21] also adopted the CDBN for artist, genre and key recognition tasks. But they

CHAPTER 2. FEATURE LEARNING FRAMEWORK 17

trained it on two hand-engineered features (EchoNest chroma and timbre features)

using a multi-modal deep learning approach [77].

Deep learning algorithms have advantages in that they can find highly non-linear

or hierarchical feature representations. However, they usually have a number of hyper-

parameters to be tuned in a cross-validation stage and the range of parameter values

are often selected based on computational constraints. Alternatively, another group

of researchers focused on single-layer learning algorithms. In particular, Coates et.

al. showed that single-layer feature learning algorithms can outperform multi-layer

algorithms on benchmark image datasets when data is appropriately preprocessed

and sufficient features (i.e., a large dictionary) are learned [16].

Similar approaches have been attempted for music classification. For example,

Henaff et. al. used a sparse coding algorithm as single-layer feature learning for

music genre classification [35]. For input they applied constant-Q transforms in two

ways, one on single frames and the other on octave fragments within a frame, and

then aggregated them into a segment-level feature for classification. They achieved

83.4% accuracy on octave features and 79.4% on frame features with only a lin-

ear SVM. Schluter and Osendorfer compared three single-layer algorithms (K-means,

RBM, mean-covariance RBM) and stacked versions of the two RBMs (i.e., DBNs)

in similarity-based music classification, applying them to mel-frequency spectrogram

with 40 or 70 bins and aggregating learned features in song-level to compute similar-

ity measures [90]. They showed that mcRBM performs best and found no noticeable

improvement using DBNs in their experiment. Wulng and Riedmiller focused input

size selection for feature learning while simply choosing K-means for feature learning

[111]. Specifically, they took sub-band patches from either single or multiple frames

of spectrogram (both regular and constant-Q). They showed that narrow sub-band

patches with multiple frames performed better than wide-band patches with a single

frame. In addition, constant-Q transform is superior to regular STFT when other con-

ditions remain the same. With the best parameters, they achieved 85.25% accuracy

on GTZAN.

Our method also follows the single-layer learning approach. In particular, we

focus on high-dimensional sparse feature learning and song-level summarization using

CHAPTER 2. FEATURE LEARNING FRAMEWORK 18

Waveform AutomaticGain Control

Time-FrequencyRepresentation

Feature Learning Algorithm

Max-Pooling Aggregation

AmplitudeCompression

Preprocessing

MultipleFrames

PCA Whitening

Summarization

Feature Learning

Song-LevelFeature Vector Classifier

Local Sparse Feature Vector

Figure 2.1: Proposed data processing pipeline for feature learning

max-pooling. While a similar method has been exploited in computer vision [16],

we suggest effective pre-processing techniques and feature learning strategies for the

audio domain. Also, we provide insight on the learned features by showing how

the short-term acoustic patterns are associated with musical semantics (see Section

2.5.1). In addition, we thoroughly analyze the effect of preprocessing and feature

representation parameters including dictionary size, sparsity and max-pooling.

2.3 Proposed Method

We propose a data processing pipeline to build a feature representation by unsuper-

vised learning. This is composed of three sections as shown in Figure 2.1. The first

section preprocesses the raw waveform and returns a normalized time-frequency rep-

resentation. The second section performs feature learning and extracts local features.

Finally, the third section summarizes the local features and produces a song-level

CHAPTER 2. FEATURE LEARNING FRAMEWORK 19

feature. Each block is detailed below.

2.3.1 Preprocessing

Musical signals are a highly variable and complex data. Ideally, we wish to have

feature-learning algorithm that capture all possible variations from the raw data (i.e.,

waveforms) without any preprocessing. In practice, however, appropriate peripheral

processing such as a time-frequency transform or normalization significantly helps

feature learning. In this section, we discuss time-frequency representations and sug-

gest normalization techniques to facilitate learning algorithms. Note that this section

somewhat uses acoustic knowledge or insight from it. However, we try to minimize it

by preserving the original input data intact as much as possible.

Automatic Gain Control

Sound is inherently dynamic. The constantly varying intensity and bandwidth in a

musical signal is a specific instance of this dynamism. The human hearing system

has a dynamic-range compression mechanism to control the level of incoming sounds.

This automatic gain control (AGC) is performed separately on each band of the

cochlear filter, thereby providing more regulated filter levels to auditory nerves [61].

Inspired by this auditory processing, we apply a time-frequency AGC as a front-end

preprocessing step.

The use of a time-frequency AGC assumes that a time-frequency representation

will be used as input data for feature-learning algorithms. In general, there is less

high-frequency energy than low-frequency energy in most classes of sounds, including

music due to timbral natures of sound sources or phase cancelation when mixing

multiple sources. As a result, high-frequency content is likely to be ignored by feature-

learning algorithm. A time-frequency AGC equalizes the spectral distribution so that

a learning algorithm can effectively capture dependencies in the spectral domain. In

addition to the spectral imbalance, the overall volume of sound files in a dataset

is usually not normalized because they are often obtained under different recording

conditions. As a byproduct, a time-frequency AGC regularizes the overall gain as

CHAPTER 2. FEATURE LEARNING FRAMEWORK 20

Input Spectrogram

kH

z

0 0.5 1 1.5 2 2.50

5

10

Sub−band Envelopes

kH

z

0 0.5 1 1.5 2 2.50

5

10

Equalized Spectrogram

Seconds

kH

z

0 0.5 1 1.5 2 2.50

5

10

−20

0

20

40

−40

−30

−20

−10

0

−20

0

20

40

Figure 2.2: Effect of a time-frequency Automatic Gain Control (AGC). The sub-bandenvelopes obtained from the original spectrogram are remapped to linear frequency(middle). This is used to equalize the original spectrogram (bottom). Note that thehigh-frequency content is boosted by the AGC in the output spectrogram.

well.

We used the time-frequency AGC proposed by Ellis [23]. It first computes an

FFT and maps the magnitude to a small number of sub-bands. Then it extracts

amplitude envelopes from each band using an envelope follower and remaps the sub-

band envelopes back to the linear-frequency scale as shown in the middle of Figure

2.2. Finally, it divides the original spectrogram by the remapped envelopes. As a

CHAPTER 2. FEATURE LEARNING FRAMEWORK 21

result, the time-frequency AGC equalizes input signals so that their energy spectrum

is more uniform, as shown in the bottom of Figure 2.2.

Time-frequency Representation

Musical sounds are generally characterized by harmonic or non-harmonic elements.

Thus, time-frequency transforms, whose basis functions are given as sinusoids, pro-

vide more interpretable representations of musical sounds. There are many choices

of time-frequency transforms such as STFT, constant-Q transform (more generally,

wavelet transform) and various filter banks. They are often modified by adapting

psycho-acoustic considerations. For example, the STFT is often mapped to percep-

tual frequency scales such as mel [99], Bark [113] and ERB [67] or to perceptual

amplitude levels such as loudness [97]. Constant-Q considers the perceptual aspect

by definition, that is, having higher frequency resolutions at lower frequencies. Some

filter banks are designed from scratch to imitate the cochlear responses in human ears

[84, 61].

In our data processing pipeline, we choose mel-frequency spectrogram as the pri-

mary time-frequency representation. The mel-frequency spectrogram is obtained from

spectrogram by mapping FFT frequency bins to a small number of mel frequency bins.

This emphasizes low-frequency content and squeezes high-frequencies into a smaller

number of bins as shown in Figure 2.3. There are two advantages of this frequency

mapping. First, it reduces the dimensionality of the data. This facilitates feature-

learning algorithms even with wide receptive fields, for example, if multiple frames

are used as input to the feature-learning algorithms (this will be explained in Section

2.3.2). Second, it alleviates unnecessarily detailed variations in the high frequency

content. In general, high-frequency content is statistically more variable or random.

The human ear is often agnostic to such details. Thus, high-frequency content is often

modeled as a wide sub-band energy in audio coding. The mel-frequency spectrogram

performs a similar mapping. Note that we chose a larger number of mel-frequency

bins in our experiments so that the reconstructed output from the mel-frequency spec-

trogram preserves the original quality as much as possible and thus the underlying

structure of the musical data can be discovered via feature-learning algorithms.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 22

Equalized Spectrogram

kH

z

0 0.5 1 1.5 2 2.50

5

10

Mel−freq. Spectrogram

Seconds

Mel−

freq b

ins

0 0.5 1 1.5 2 2.5

20

40

60

80

100

120

−20

0

20

40

−40

−20

0

Figure 2.3: Comparison of (equalized) spectrogram and mel-frequency spectrogram.513 bins in spectrogram (FFT size is 1024) are mapped to 128 bins in mel-frequencyspectrogram.

Magnitude Compression

As a final preprocessing step, we compress the amplitude of the mel-frequency spec-

trogram using an approximate log scale, log10(1 + C|X(t, f)|), where |X(t, f)| is the

time-frequency representation and C controls the degree of compression [69]. In gen-

eral, the linear content of each bin of spectrogram or mel-frequency spectrogram

has an exponential distribution. Scaling with a log function makes it have a more

Gaussian-like distribution. This enables the magnitude to be well-fitted with PCA

whitening, which has an implicit Gaussian assumption.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 23

2.3.2 Feature Learning: Common Properties

Now the input data is ready to be used for feature-learning algorithms. As previously

stated, the input data is pre-processed in a highly constrained manner so that it still

contains salient acoustic characteristics such as harmonic and transient distributions,

low and high energy, and their temporal dependency in the spectral domain. We

are going to capture the characteristic patterns with different unsupervised learn-

ing algorithms, for example, K-means, restricted Boltzmann machine, sparse coding,

auto-encoder and so forth. Although they perform the task using different mathemat-

ical schemes, their basic objective functions and meta parameters are shared. This

section discusses the common properties in a general perspective before introducing

individual algorithms. Firstly, we introduce PCA whitening used prior to feature

learning as another preprocessing step.1

PCA Whitening

PCA is a popular algorithm for reducing the dimensionality or removing pair-wise

correlation. PCA extracts a set of basis functions in a way that maximizes the

variance of the projected subspace. In this sense, it can be regarded as a feature-

learning algorithm. However, the basis functions are limited to be orthogonal and

thus it can learn only the second-order dependencies (a Gaussian distribution) of the

input data. For this reason, PCA is often used as a preprocessing step followed by

Independent Component Analysis (ICA) or other learning algorithms that capture

high-order dependency. With an additional normalizing step that makes the projected

space have unit variances, this processing is called PCA whitening [42]. Algorithm 1

describes the procedure to learn the transformation matrix for PCA whitening given

a matrix X where each column is a sample from input data. As a result, subtracting

the mean of X from each column of a new input data and multiply U′

performs

PCA whitening. Figure 2.4 shows an example of the PCA-whitened mel-frequency

1In a previous work, we placed PCA whitening in the pre-processing stage due to its rule inthe data processing pipeline [71]. However, we moved it to the feature-learning stage because thewhitening matrix is obtained in a data-driven manner similarly to other feature-learning algorithmsand, furthermore, PCA whitening itself is often used as a feature-learning algorithm [33].

CHAPTER 2. FEATURE LEARNING FRAMEWORK 24

Algorithm 1 PCA Whitening

1. Subtract the mean at each row of X. This returns a matrix X′

that has a zeromean for each row.

2. Compute the covariance matrix of X′

and perform eigen decomposition

X ′X ′T = UV UT , (2.1)

This returns the eigenvectors ui from each column of U and eigenvalues vi fromthe diagonal of V for i = 1, 2, 3...R where R is the input data dimension.

3. Choose ρ and find the maximum k such that

k∑i=1

vi

R∑i=1

vi

< ρ. This determines the

amount of dimensionality reduction.

4. Form a normalization (and diagonal) matrix D whose diagonal elements are1

vi+ε, where ε is a regularization parameter to prevent diagonal elements from

being excessively increased.

5. Multiply D and U , and define the output as U′. This is used as the whitening

matrix.

spectrogram.

Objective Function

Feature-learning algorithms commonly learn a basis functions, which is often called

a dictionary. They represent the input data with a linear combination of the basis

functions learned using the following objective function:

minD,s(i)

∑i

∥∥x(i) −Ds(i)∥∥22 (2.2)

where x(i) is an example of the input data, D is a matrix containing the basis functions,

and s(i) is an encoded feature vector based on the learned basis functions. In a

probabilistic setting, such as RBM, the objective function is described in a maximum

CHAPTER 2. FEATURE LEARNING FRAMEWORK 25

Figure 2.4: Mel-frequency spectrogram (top) and PCA-whitened mel-frequency spec-trogram (bottom). The figure indicates that 4 frames in the mel-frequency spectro-gram are chosen as a receptive field size and projected to a single vector in the PCAspace. The PCA indices are sorted such that components with high eigenvalues havelower index numbers.

likelihood form:

maxD

P (x|D) = maxD

∑h

P (x|D,h)P (h), (2.3)

where x is a random vector corresponding to the input data and h is a hidden random

vector that corresponds to s(i) above. This optimization problem is usually solved

by a type of bi-directional relaxation similar to Expectation-Maximization (EM).

This is described in Algorithm 2. Once we obtain the dictionary D, the feature

representations for new input data, for example, test data, can be calculated by

solving the second step above with fixed D.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 26

Algorithm 2 Feature Learning Procedure

1. Initialize the basis functions D using either random numbers or randomly se-lected input data with appropriate normalization.

2. Fix D and solve Equation 2.2 to obtain s(i) or infer P (h|x, D) given x(i) inEquation 2.3.

3. Update D using the previous results. For example, solve the least squaresproblem given x(i) and fixed s(i)in Equation 2.2.

4. Repeat step 2 and 3 until the objective function converges.

Sparsity

In addition to the basic objective function defined in Equation 2.2 or 2.3, most feature-

learning algorithms have additional sparsity constraint on s(i) or h as this turns

out to be the core factor to be able to discover meaningful structures [12, 78, 96,

53]. Usually, an L1-norm or the mean of extracted feature vectors is used as an

additional regularization term. Specifically, the second step in Algorithm 2 includes

the additional sparsity term in the objective function.

Receptive Field Size

Receptive field originally refers to a region of space where the presence of a stimulus

changes the activity of a neuron in the human sensory system. In the feature-learning

context, this term is often used to indicate the input size from which a feature is

learned such as image patch size. In our setting, we select the receptive field from

the mel-frequency spectrogram. There are two aspects of the receptive field size: the

number of frames (time-wise) and the number of subbands (frequency-wise). The

number of frames is changed by selecting a different number of consecutive frames

along the time axis. Since neighboring frames in musical sounds tend to depend on

each other, for example, musical notes are usually sustained, the receptive field size

that covers multiple frames allows the system to learn temporal dependencies. The

number of subbands in the receptive field determines whether features are learned

CHAPTER 2. FEATURE LEARNING FRAMEWORK 27

separately for each subband or just once for a whole band. Using separate receptive

fields, however, may require an additional learning layer that captures dependency

among the sub-bands. In order to reduce complexity, we simply take the whole band

in our experiment. Figure 2.4 indicates the receptive field size and shows how multiple

frames of the mel-frequency spectrogram are used as input to the learning algorithm.

Dictionary Size

Dictionary size refers to the number of the basis functions. This is also called feature

size or hidden layer size in different algorithms or contexts. Dictionary size is one of

the most important meta parameters of learning algorithms because it directly deter-

mines the diversity of learned features in single-layer algorithms and also corresponds

to the input dimension of the classifier in our pipeline. In general, a large number

of dictionary size is preferred in single-layer algorithms as it significantly influences

classification performance [16].

2.3.3 Feature Learning: Algorithms

This section describes individual feature-learning algorithms that we evaluated in our

experiment.

K-means Clustering

K-means clustering is a well-known unsupervised learning algorithm. It learns K

cluster and their centroids from the input data and assigns the membership of a

given input to one of the K clusters. The training is performed by alternating two

steps:

1. Compute distances between data examples and the current centroids and find

the nearest centroids for each data example.

2. Update centroids with the mean of the examples assigned to each centroid.

As a result, given a set of centroids, K-means performs a hard assignment as follow:

CHAPTER 2. FEATURE LEARNING FRAMEWORK 28

fk(x) =

1 if k = argminj||x− c(j)||220 otherwise .

(2.4)

This computation is exactly the same as steps 2 and 3 of Algorithm 2 if a set

of K centroids is regarded as a dictionary and the membership is represented as an

extremely sparse feature vector s that has all zeros except a single “1” corresponding

to the assigned centroid. For example, if K is 5 and an example is closest to the 1st

centroid, the feature vector s will be represented as [1 0 0 0 0]. In this sense, K-means

can be regarded as having a sparsity constraint by nature, specifically, providing the

maximally sparse representation.

K-means Clustering: Soft Encoding

Coates et. al. recently used a variation of K-means encoding, noting that the hard

assignment by K-means makes the feature vector too terse [16]. Given a set of cen-

troids (learned using the K-means clustering), the new encoding method performs

a non-linear mapping that attempts to be “gentler” while maintaining sparsity as

follows:

fk(x) = max(0, µ(z)− zk), (2.5)

where zk = ||x − c(k)||2 and µ(z) is the mean of the elements of z. This encoding

returns 0 for any feature fk where the distance to the centroid c(k) is “above average,”

thereby setting half the elements of the feature vector s to zero. They showed that this

simple tweaking of the encoding significantly improve image recognition performance.

We refer to this soft encoding as K-means (soft) and the hard assignment version as

K-mean (hard).

Sparse Coding

Sparse coding is an unsupervised learning algorithm to represent input data efficiently

with a set of basis functions that we call a dictionary. The goal of sparse coding is

to find a dictionary such that an input vector x ∈ Rn is represented as a linear

CHAPTER 2. FEATURE LEARNING FRAMEWORK 29

combination of the dictionary elements:

x =k∑j=1

sjdj , (2.6)

where dj is a dictionary element vector and sj is the corresponding coefficient. The

dictionary size k is often set to be greater than input dimensionality n, rendering

Equation 2.6 an over-complete representation. However, with this condition (k >

n), the coefficients s are no longer uniquely determined. Therefore, an additional

constraint called sparsity is introduced to resolve this problem. Ideally, an L0 norm

of the coefficients s, the number of non-zero values, is used as sparsity constraint.

However, it is not differentiable and difficult to optimize in general, and is known

to be NP-hard. Instead, the L1 norm of the coefficients s, is commonly used as a

relaxation. As a result, a dictionary to build a over-complete representation is learned

using the following objective function:

minD,s(i)

∑i

∥∥Ds(i) − x(i)∥∥22

+ λ∥∥s(i)∥∥

1, (2.7)

where D is the dictionary matrix si is the sparse code and λ controls the amount of

sparsity. In addition, the dictionary D is constrained to be normalized such that, for

each dictionary element, ‖dj‖22 = 1.

There are many algorithms to find the dictionary and sparse code. They are all

based on the alternating minimization rule in Section 2.3.2. Specifically, the second

step, which finds sparse code s(i) given a dictionary D, is called the L1 minimization

problem or inference step in the probabilistic context:

mins(i)

∑i

∥∥Ds(i) − x(i)∥∥22

+ λ∥∥s(i)∥∥

1. (2.8)

The third step, which updates the dictionary given the inferred sj, is formed as a

CHAPTER 2. FEATURE LEARNING FRAMEWORK 30

least squares problem, often with a unit L2 norm constraint for dj:

minD

∑i

∥∥Ds(i) − x(i)∥∥22

subject to ‖dj‖22 = 1, ∀j.(2.9)

Following Coates [17], we solved the L1 minimization problem using a coordinate

descent algorithm [110] and the least squares problem simply using a inverse of the

dictionary matrix followed by the unit normalization. After we learn the dictionary,

we use the absolute value of the sparse code s(i) as learned features to incorporate it

into the max-pooling and aggregation stage.

Auto-Encoder

Auto-encoder is an unsupervised neural network algorithm where the output is con-

figured to reconstruct the input [74, 37, 4]. A single-layer neural network is described

as follows:

h(i) = g(W1x(i) + b1) (2.10)

y(i) = W2h(i) + b2 , (2.11)

where x is the input data, h is the hidden layer output, y is the output, W1 and W2

are network parameters, b1 and b2 are bias terms, and g(z) = 1/(1 + exp(−z)) is the

logistic sigmoid function. The auto-encoder is trained to minimize the reconstruction

error:

E(W1,W2, b1, b2) =∑i

∥∥y(i)(x(i))− x(i)∥∥2, (2.12)

where W1 and W2 are often tied to be transpose to each other, that is, W T1 = W2.

In addition, we introduce a constraint on the hidden layer to promote sparsity.

We use the KL divergence between a target activation ρ and the mean activation ρj

CHAPTER 2. FEATURE LEARNING FRAMEWORK 31

for the hidden layer [74]:

K∑j=1

KL(ρ||ρj) =K∑j=1

ρ logρ

ρj+ (1− ρ) log

1− ρ1− ρj

(2.13)

where ρj = 1m

m∑i=1

h(x(i)). This way, sparsity is controlled by setting a small value to

the target activation. Finally, the sparse auto-encoder is trained by minimizing the

sum of the reconstruction error and the KL divergence:

minW1,W2

E ′(W1,W2, b1, b2) = minW1,W2

E(W1,W2, b1, b2) +K∑j=1

KL(ρ||ρj) . (2.14)

This is usually solved by gradient descent because both terms are differentiable with

regard to the network parameters and bias terms:

W1 ← W1 + µ∂E ′(W1,W2, b1, b2)

∂W1

(2.15)

W2 ← W2 + µ∂E ′(W1,W2, b1, b2)

∂W2

(2.16)

b2 ← b1 + µ∂E ′(W1,W2, b1, b2)

∂b1(2.17)

b2 ← b2 + µ∂E ′(W1,W2, b1, b2)

∂b2, (2.18)

where µ is the learning rate.

Note that W2 plays a role of feature bases in feature-learning context and h(i)(x(i))

is the learned features.

Restricted Boltzmann Machine

A restricted Boltzmann Machine (RBM) is a probabilistic version of the auto-encoder.

The RBM is defined as a bipartite undirected graphical model as shown in Figure 2.5

CHAPTER 2. FEATURE LEARNING FRAMEWORK 32

[98]. It consists of visible nodes x and hidden nodes h where the visible nodes repre-

sent input vectors and the hidden nodes represent the features learned by training the

RBM. The joint probability for the hidden and visible nodes is defined in Equation

2.19 when the visible notes are real-valued Gaussian units and the hidden notes are

binary units. The RBM has symmetric connections between the two layers denoted

by a weight matrix W , but no connections within the hidden nodes or visible nodes.

This particular configuration makes it easy to compute the conditional probability

distributions, when nodes in either layer are observed (Equation 2.20 and 2.21).

− logP (x,h) ∝ E(x,h) =1

2σ2xTx− 1

σ2

(cTx + bTh + hTWx

)(2.19)

p(hj|x) = g(1

σ2(bj + wT

j x)) (2.20)

p(xi|h) = N ((ci + wTi h), σ2), (2.21)

where σ2 is a scaling factor, b and c are bias terms, and W is a weight matrix. The

parameters are estimated by maximizing the log-likelihood of the visible nodes. This

is approximated by block Gibbs sampling between two layers, particularly, using the

contrastive-divergence learning rule which involves only a single iteration of Gibbs

sampling [36].

x1 x2 x3

h1 h2 h3 h4

W

b

c

Figure 2.5: Undirected graphical model of a RBM.

We further regularize this model with a sparsity term by encouraging each hidden

CHAPTER 2. FEATURE LEARNING FRAMEWORK 33

unit to have a pre-determined expected activation using a regularization penalty:

∑j

(ρ− 1

m(m∑l=1

E[hj|xl]))2, (2.22)

where {x1, ...,xm} is the training set and ρ determines the target sparsity of the hidden

unit activations [53]. This term is added to the maximum-likelihood estimation.

Finally, the parameters are obtained from the following optimization problem:

minW,b,c

m∑l=1

− log∑h

P (x,h) + λ∑j

(ρ− 1

m(m∑l=1

E[hj|xl]))2 (2.23)

Similarly to the sparse auto-encoder, we can interpret W as a dictionary and h as

sparse code. Also, both the auto-encoder and the RBM do not need any iterative com-

putation because they have an explicit encoding scheme h = g( 1σ2 (b + W Tx)). This

property is particularly useful in real-time applications because the feature can be ex-

tracted using a simple feed-forward computation. Figure 2.3 shows a mel-frequency

spectrogram and the corresponding hidden-layer activation. Note that the encoded

feature representation is very sparse compared to the mel-frequency spectrogram (only

1% of hidden units are activated in this example).

2.3.4 Feature Summarization

Max-Pooling and Aggregation

The feature-representation algorithms provide sparse feature vectors for short-term

frames. Since a song is a very long sequence of data, however, we need to summarize

them to construct a song-level feature. A typical approach is aggregating them as a

histogram, that is, by averaging them over a song or a song segment. [32, 35]. How-

ever, averaging every single short-terms feature can dilute their local discriminative

power. Therefore, we first perform pooling to find only dominant features over a

small segment and then aggregate the output over a song.

Pooling summarizes local features obtained at neighboring locations into a group

of statistic that is often called a bag of features (BoF) [10]. For example, Hamel et.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 34

Mel−freq. Spectrogram

Seconds

Me

l−fr

eq

bin

s

0 0.5 1 1.5 2 2.5

20

40

60

80

100

120

−40

−20

0

Feature Bases

Hidden Layer Index

Me

l−fr

eq

bin

s

20

40

60

80

100

120

−1

−0.5

0

0.5

1

Hidden Layer Activation

Seconds

Hid

de

n L

aye

r In

de

x

0 0.5 1 1.5 2 2.5

50

100

150

200

250

0.2

0.4

0.6

0.8

Figure 2.6: Comparison of mel-frequency spectrogram and the learned feature repre-sentation (hidden layer activation) using sparse RBM. The dictionary size was set to1024 and target activation (sparsity) was to 0.01. That means that only 1% out of1024 features are activated on average.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 35

al evaluated various types of pooling such as mean, min, max, variance and other

high-order statistics and their combinations on PCA-whitened frame-level features

for music tag classification [33]. We perform max-pooling by taking the maximum

value at each dimension over a segment. Max-pooling is often used in convolutional

neural networks to reduce feature scales and also make them invariant to local shifts.

In our experiment, max-pooling works as a form of temporal masking because it

discards small details around a high peak of feature activation. Figure 2.7 compares

the hidden layer activation obtained by training sparse RBM with its max-pooled

version. This shows that max-pooling removes small noisy feature activation and

makes peaks more distinct. The max-pooled features are summed over a song as a

histogram of dominant feature activations. This summarization eventually provides

a single song-level feature vector.

2.3.5 Classification

The feature representation data processing pipeline produces a single feature vector

for a song. They are used for a classifier that performs supervised training where a

single or multiple labels are given. In our experiment, we use a linear support vector

machine (SVM) as a reference classifier to evaluate the feature representations learned

by different algorithms and parameters. The linear SVM is trained by minimizing a

L2-regularized L2 hinge loss given the training data [28]:

minw

M∑j=1

(1

2wTj wj + C

N∑i=1

max(0, 1− y(i)j wTj x(i)))2 , (2.24)

where wj is SVM parameter, C is a regularization weight, x(i) is the song-level fea-

ture vector augmented with ’1’ for the bias term, y(i)j ∈ {−1, 1} is the label and j

corresponds to one of the M output classes. After training, we make predictions by

choosing j that produces the maximum value of wTj x(i). We implemented the SVM

using the publicly available Matlab optimization library, minFunc.2

2Matlab library found in http://www.cs.ubc.ca/~schmidtm/Software/minFunc.html

CHAPTER 2. FEATURE LEARNING FRAMEWORK 36

05

1015

20

0

5

10

15

200

0.5

1

Seconds

Hidden Layer Activation

Hidden Layer Index

05

1015

20

0

5

10

15

200

0.5

1

SecondsHidden Layer Index

Figure 2.7: Comparison of hidden layer activation and its max-pooled version. Forvisualization, the hidden layer features masked by the maximum value in their poolingregion were set to zero. As a result, the max-pooled feature maintains only locallydominant activation. The hidden layer feature representation was learned using sparseRBM. Dictionary size was set to 1024, sparsity was to 0.01 and max-pooling size wasto 43 frames (about 1 second).

CHAPTER 2. FEATURE LEARNING FRAMEWORK 37

2.4 Experiments

2.4.1 Datasets

We evaluated our proposed feature representation on two popularly used genre datasets:

GTZAN and ISMIR2004.

GTZAN

GTZAN is a public music genre dataset released by Tzanetakis [104]. This has been

used as a benchmark dataset to evaluate genre classification algorithms [80, 58, 62, 5,

51, 82, 83, 29, 6, 45]. Recent work based on feature-learning algorithms also evaluated

these methods on this dataset [32, 35]. GTZAN contains 10 different genres and 100

song segments evenly for each genre: Blues, Classical, Country, Disco, Hiphop, Jazz,

Metal, Pop, Reggae, and Rock. All song segments are 30-second long.

ISMIR2004

ISMIR2004 is another publicly available genre dataset released after the MIREX 2004

competition.3 This contains 6 different genres: Classical (320 samples), Electronic

(115 samples), Jazz/Blues (26 samples), Metal/Punk (45 samples), Rock/Pop (101

samples), and World (122 samples). Each sample has the full length of a song. In

this experiment, we took only a 30-second segment after the first 30 seconds of each

song. This dataset has been also a secondary benchmark dataset to evaluate genre

classification algorithms [59, 79, 39, 82].

2.4.2 Preprocessing Parameters

We first resampled the waveform data to 22.05kHz and computed an FFT with a 46ms

Hann window and 50% overlap. This produces a 513 dimensional vector (up to half

the sampling rate) for each frame. As for the time-frequency AGC, we mapped the

spectrogram to 10 sub-bands and calculated each envelope using temporal smoothing.

3http://ismir2004.ismir.net/genre_contest/index.htm

CHAPTER 2. FEATURE LEARNING FRAMEWORK 38

After equalizing the spectrogram with the envelopes (by remapping them back to the

spectrogram domain), we then converted the output to a mel-frequency spectrogram

with 128 bins. In the magnitude compression, the strength C was fixed to 10.

2.4.3 Feature-Learning Parameters

We sampled 100,000 data examples at random positions for each dataset for PCA

whitening and subsequent feature learning. The receptive field was selected as a

128 × n (n=1, 2, 4 and 6) patch from the preprocessed data. In PCA whitening,

we reduced the dimensionality of the sampled data by retaining 90% of the variance.

Before the whitening, we added 0.01 to the variance for regularization. We used

dictionary size (or hidden layer size) and sparsity (when applicable) as the primary

feature-learning parameters. The dictionary size was varied over 128, 256, 512, 1024

and 2048. The sparsity parameter was set to ρ = 0.005, 0.007, 0.01, 0.02, 0.03 and

0.05 for sparse RBM and sparse Auto-Encoder and λ = 1.0, 1.5, 2.0 and 2.5 for sparse

coding. Then, max-pooling was performed over segments of length 0.1, 0.25, 0.5, 1,

2, 4, 8 and 16 seconds.

2.4.4 Classifier Parameters

We first standardized the song-level features obtained from the feature representation

stage by subtracting the mean and dividing by the standard deviation of those in the

training set. We fixed a classifier to a linear Support Vector Machine (SVM) in

order to focus only on evaluation features. In GTZAN, we performed 10-fold cross-

validation following the practice of previous work. In ISMIR2004, we used the original

split between training, development and test sets.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 39

2.5 Evaluation

2.5.1 Visualization

We show feature bases learned by the sparse RBM in Figure 2.6. While the majority

of previous work show such feature bases patterns captured by different feature-

learning algorithms, there are few attempt to associate them with high-level musical

semantics such as genre. We suggest a systematic way to find the relationship and

better understand the learned features. This is performed by finding the top-K active

feature bases given a genre label as follows:

1. Select a subset of the dataset that belong to a specific genre.

2. Given learned feature bases, compute feature representations of the subset.

3. Summarize all the local features using a histogram.

4. Sort the histogram in descending order and choose the top-K elements that have

the highest values.

5. Repeat the steps above for each genre.

This returns the most actively triggered K feature bases for the music genre. Figure

2.8 shows the most active top-20 feature bases learned on GTZAN for ten music

genres (one for sparse RBM and the other for sparse coding). The correspond-

ing feature bases are vividly distinguished by different timbral patterns, such as

harmonic/non-harmonic, wide/narrow band, strong low/high-frequency content and

steady/transient ones. Different genres of music trigger specific types of feature bases

more actively than others. For example, classical and jazz music mainly trigger

harmonic features. Metal music is likely to have more high-frequency energy and

wide-band patterns. In addition, disco, reggae and hiphop music contain significant

portions of transient patterns. This indicates the feature-learning algorithms effec-

tively encode the input data into high-dimensional sparse feature vectors such that

the feature vectors are “selectively” activated by a given genre of music.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 40

blues

Mel−

frequency b

in

20 40 60 80

20

40

60

80

100

120

classical

20 40 60 80

20

40

60

80

100

120

country

20 40 60 80

20

40

60

80

100

120

disco

20 40 60 80

20

40

60

80

100

120

hiphop

20 40 60 80

20

40

60

80

100

120

jazz

Mel−

frequency b

in

20 40 60 80

20

40

60

80

100

120

metal

20 40 60 80

20

40

60

80

100

120

pop

20 40 60 80

20

40

60

80

100

120

reggae

20 40 60 80

20

40

60

80

100

120

rock

20 40 60 80

20

40

60

80

100

120

(a) Sparse RBM

blues

Mel−

frequency b

in

20 40 60 80

20

40

60

80

100

120

classical

20 40 60 80

20

40

60

80

100

120

country

20 40 60 80

20

40

60

80

100

120

disco

20 40 60 80

20

40

60

80

100

120

hiphop

20 40 60 80

20

40

60

80

100

120

jazz

Mel−

frequency b

in

20 40 60 80

20

40

60

80

100

120

metal

20 40 60 80

20

40

60

80

100

120

pop

20 40 60 80

20

40

60

80

100

120

reggae

20 40 60 80

20

40

60

80

100

120

rock

20 40 60 80

20

40

60

80

100

120

(b) Sparse Coding

Figure 2.8: Top 20 most active feature bases (dictionary elements) for ten genres atthe GTZAN set. (a) Sparse RBM with 1024 hidden units and 0.03 sparsity (b) Sparsecoding with 1024 dictionary elements and λ=1.5.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 41

2.5.2 Results and Discussion

We analyze genre classification accuracy for different preprocessing options and meta

parameters of learning algorithms.

Spectrogram versus mel-frequency spectrogram

Figure 2.9 compares the linear spectrogram and mel-frequency spectrogram as the

input time-frequency representation. We compare different receptive field sizes for

both the sparse RBM and sparse coding. The mel-frequency spectrogram outper-

forms the linear spectrogram for all cases. This confirms that high frequency content

tends to have unnecessary details and thus squeezing them (as done by mel-frequency

spectrogram) helps the learning algorithm focus on the more important variations.

In addition, since the mel-frequency spectrogram has a much smaller size (by 4 times

in the experiment), it is computationally more efficient.

1 2 3 4 5 670

75

80

85

90

95

Number of frames

Accu

racy [

%]

spec

mel−spec

(a) RBM

1 2 3 4 5 670

75

80

85

90

95

Number of frames

Accu

racy [

%]

spec

mel−spec

(b) Sparse Coding

Figure 2.9: Comparison of spectrogram and mel-frequency spectrogram. The dictio-nary size is set to 1024.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 42

Effect of Automatic Gain Control (AGC)

Figure 2.10 shows the effect of the time-frequency AGC for different receptive field

sizes. Again, they are compared separately for the sparse RBM and sparse coding.

For both feature-learning algorithms, the AGC consistently increases the accuracy by

about 2 ∼ 5% regardless of the receptive field size.

1 2 3 4 5 670

75

80

85

90

95

Number of frames

Accu

racy [

%]

no AGC

AGC

(a) RBM

1 2 3 4 5 670

75

80

85

90

95

Number of frames

Accu

racy [

%]

no AGC

AGC

(b) Sparse Coding

Figure 2.10: Effect of Automatic Gain Control. The dictionary size is set to 1024.

Effect of Receptive Field Size

Figure 2.11 compares the results for different receptive field sizes and separately for

five feature-learning algorithms. In general, accuracy increased when multiple frames

were chosen. This indicates that learning temporal dependencies indeed helps to

discriminate different genres. However, as the receptive field size increases, the result

becomes saturated from the four frames. This is expected at some point because

variations in the receptive field grow exponentially and thus learning algorithms are

no longer capable of capturing the variations. Among the five algorithms, sparse

coding outperforms the others for all sizes. It is notable that K-means (hard) get

CHAPTER 2. FEATURE LEARNING FRAMEWORK 43

1 2 4 675

80

85

90

Number of frames

Accura

cy [%

]

SC

RBM

Kmeans(hard)

Kmeans(tri)

AE

Figure 2.11: Effect of receptive field size. The dictionary size is set to 1024.

degraded as the numbers of frames increases whereas K-means (soft) is comparable

to other algorithms that have sparsity control. This may be related to the lack of

sparsity control, that is, too extreme fixed sparsity in K-mean (hard).

Dictionary Size and Sparsity

Dictionary size is one of the most crucial parameters because it determines not only

diversity of the feature patterns but also the input dimension of the classifier that

follows in our data processing pipeline. This is illustrated in Figure 2.12. The ac-

curacy dramatically increases as the dictionary size increases from 128 to 1024 and

saturates by 2048. In addition, Figure 2.13 shows how the sparsity affects perfor-

mance for each dictionary size and separately with regard to the RBM and sparse

coding. In the RBM, sparsity is controlled by a target activation for the hidden layer

units. Thus, as the sparsity value decreases, hidden layer activation, that is, feature

representations become sparser. In sparse coding, sparsity is controlled as a weight

for sparsity constraint (i.e., L1 norm of the feature activation). Thus, as the weight

CHAPTER 2. FEATURE LEARNING FRAMEWORK 44

128 256 512 1024 204872

74

76

78

80

82

84

86

88

90

Dictionary Size [sec]

Accura

cy [%

]

SC

RBM

Kmeans(hard)

Kmeans(tri)

AE

Figure 2.12: Effect of dictionary size. The receptive field size is set to 4 frames.

(λ) increases, the feature vectors become sparser. The plots show a relation between

optimal sparsity and dictionary size. For small dictionary sizes, the accuracy tends

to be higher when the features are relatively less sparse. For large dictionary sizes, on

the other hand, the accuracy tends to be higher when they are sparser. As a result,

the best genre-classification accuracy is achieved by “high-dimensional sparse feature

representations.”

Max-pooling Size

Figure 2.14 plots the results for different max-pooling sizes with regard to two dictio-

nary sizes, 512 and 2048. In general, as the max-pooling size becomes larger, accuracy

increases up to a point and then decreases after that. Compared to the point where

max-pooling is not performed (the leftmost one), the peak classification performance

gains 6 ∼ 7 % accuracy improvement, showing that max-pooling is a very effective

technique to boost discriminative power. Another noticeable result is that the optimal

max-pooling size depends on the dictionary size. For example, when the dictionary

CHAPTER 2. FEATURE LEARNING FRAMEWORK 45

0.005 0.007 0.01 0.02 0.03 0.0570

72

74

76

78

80

82

84

86

88

90

Sparsity ( Target Activation )

Accu

racy [

%]

256

512

1024

2048

(a) RBM

1.0 1.5 2.0 2.570

75

80

85

90

Sparsity ( Lambda )

Accu

racy [

%]

256

512

1024

2048

(b) Sparse Coding

Figure 2.13: Effect of sparsity

size is 512, the best accuracy is obtained around 1 second. However, as the dictionary

size increases to 2048, the best max-pooling size moves toward 4 second, boosting the

accuracy overall by 3 % or so. This shows that the optimal max-pooling size depends

on the dictionary size.

Comparison of feature-learning algorithms

Table 2.4 summarizes the best GTZAN genre-classification accuracy for five feature-

learning algorithms. The results show that sparse coding slightly outperforms all other

algorithms particularly for large dictionary sizes. However, the overall differences are

not significant (within 1∼2 %) except K-means (hard). This indicates that the feature

representation by K-means (hard) is too sparse and thus some amount of sparsity is

necessary to achieve good performance. In this sense, the soft encoding in K-means

(soft) is seen as a simple and effective alternative to the hard assignment in K-means

(hard). Table 2.5 summarizes the best genre classification accuracy on ISMIR2004.

Overall, the results are very similar to those on GTZAN except that K-means (soft)

CHAPTER 2. FEATURE LEARNING FRAMEWORK 46

0.023 0.1 0.25 0.5 1 2 4 8 1675

80

85

90

Max−pooling [sec]

Accu

racy [

%]

SC

RBM

Kmeans(hard)

Kmeans(tri)

AE

(a) dictionary size = 512

0.023 0.1 0.25 0.5 1 2 4 8 1675

80

85

90

Max−pooling [sec]

Accu

racy [

%]

SC

RBM

Kmeans(hard)

Kmeans(tri)

AE

(b) dictionary size = 2048

Figure 2.14: Effect of max-pooling

achieves outstanding accuracy for small dictionary sizes.

While sparse coding is slightly superior in terms of accuracy to other algorithms,

the encoding in sparse coding requires iterative computation steps to solve the L1

minimization. This slows down feature extraction for new input in the testing phase,

prohibiting it from being used in real-time applications 4. On the other hand, sparse

RBM and auto-encoder have an explicit encoding scheme that has a single feed-

forward computation step. K-means also can extract features simply by computing

distances from centroids. Thus, these algorithms are more useful when real-time

processing is required or computation resources are limited.

In terms of training time or complexity, auto-encoder and RBM were most difficult

to train because they need a number of hyper-parameters and cross-validating all of

the meta parameters can be time-consuming. Thus, we fixed several hyper-parameters

(e.g., weight cost and scale) to a constant after finding appropriate values in our initial

experiments. On the other hand, K-means clustering (both hard and soft encoding)

4A variant of sparse coding, called Predictive Sparse Coding [47], has a separate encoding schemeto overcome this problem.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 47

Algorithms Best Accuracy (%)

Dictionary Size 128 256 512 1024 2048

K-mean (hard) 75.9 79.6 82.6 85.1 86.0

K-mean (soft) 78.5 82.2 84.7 86.6 88.4

Sparse RBM 74 81.7 85.8 87.7 88.8

Sparse Auto-encoder 79.2 83.3 85.9 87.4 88.8

Sparse Coding 77.1 83.5 86.3 88.5 89.7

Table 2.1: Comparison of different feature-learning algorithms on the GTZAN genredataset.

requires only a single parameter (dictionary size) and the training algorithm is simple

and fast as well. Therefore, the K-means, especially with soft encoding, is seen to be

the most efficient solution for feature learning in this regard.

Confusion Matrix

Table 2.3 shows a confusion matrix for the best accuracy obtained with sparse coding

on GTZAN. Classical, blues, jazz and metal have high accuracy, above 94 %, whereas

country, disco and rock have relatively low accuracy. In particular, rock is significantly

lower than others, being confused mainly with country and metal. This makes sense

because rock is used as a broad musical category and the three genres share a guitar

in common. Other incorrect results also occur among relatively similar genres, for

example, disco and pop. The low accuracy in disco and reggae can be explained

by the lack of rhythmic feature extraction in our method. Although the song-level

feature implicitly captures rhythmic characteristic by averaging activation of transient

patterns (usually appearing in note onsets), this seems to be not sufficient to capture

the sophisticated rhythms in disco or reggae.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 48

Algorithms Best Accuracy (%)

Dictionary Size 128 256 512 1024 2048

K-mean (hard) 77.8 79.6 82.0 82.17 83.4

K-mean (soft) 81.9 83.8 84.5 85.7 86.1

Sparse RBM 77.4 81.9 83.8 85.0 85.7

Sparse Auto-encoder 79.5 82.3 84.6 85.0 84.9

Sparse Coding 79.4 81.8 84.6 85.7 86.8

Table 2.2: Comparison of different feature-learning algorithms on the ISMIR2004genre dataset.

Blues Classical Country Disco Hiphop Jazz Metal Pop Reggae Rock

Blues 94 1 2 1 0 2 0 0 0 0

Classical 0 99 1 0 0 0 0 0 0 0

Country 1 2 86 2 0 0 0 2 1 6

Disco 0 0 2 87 3 1 1 2 1 3

Hiphop 0 0 1 2 88 0 3 4 2 0

Jazz 1 4 0 0 0 94 1 0 0 0

Metal 0 0 0 1 0 0 97 0 0 2

Pop 0 0 2 3 0 0 1 90 3 1

Reggae 0 0 1 2 3 1 0 1 89 3

Rock 1 0 11 3 0 2 5 2 3 73

Table 2.3: Confusion Matrix for ten genres for the best algorithm (sparse coding gives89.7% accuracy) on the GTZAN gene dataset.

CHAPTER 2. FEATURE LEARNING FRAMEWORK 49

Classifiers Features Accuracy (%)

CSC Many hand-engineered features [45] 92.7

SRC Auditory temporal modulations [81] 92

Linear SVM Proposed method (sparse coding) 89.7

Linear SVM Learned using K-means [111] 85.3

RBF-SVM Learned using deep belief network [34] 84.3

Linear SVM Learned using predictive sparse coding [35] 83.4

AdaBoost Many hand-engineered features [5] 83

SVM Daubechies Wavelet Coefficients [58] 78.5

Log. Regression Spectral Covariance [6] 77

Linear SVM Auditory Temporal Modulations [81] 70

GMM Many hand-engineered features [104] 61

Table 2.4: Comparison with state-of-the-art algorithms on the GTZAN genre dataset

Classifiers Features Accuracy (%)

SRC Auditory temporal modulations [81] 93.6

Linear SVM Proposed method (sparse coding) 86.8

GMM NMF-based features [39] 83.5

NearestNeighbor Spectral Similarity / fluctuation patterns [79] 82.3

RBF-SVM Auditory temporal modulations [82] 81.0

Linear SVM Many hand-engineered features [59] 79.7

Table 2.5: Comparison with state-of-the-art algorithms on the ISMIR2004 genredataset

CHAPTER 2. FEATURE LEARNING FRAMEWORK 50

Comparison to state-of-the-art algorithms

Table 2.4 and 2.5 compare our best accuracy with previous results on GTZAN and

ISMIR2004, respectively. State-of-the-art algorithms achieved greater than 90% on

both datasets using hand-engineered features. However, they use complicated non-

linear classifiers. On the other hand, our result is the third from the top using only

using a linear classifier. Among those based on feature learning, we achieved the

highest accuracy, which is 4.4% greater than the second (85.3 %) or 5.4 % greater

than the deep learning method with a non-linear classifier (84.3 %)

2.6 Conclusion

We presented a data processing pipeline to learn high-dimensional sparse features

from a short-term receptive field on mel-frequency spectrogram and summarize them

to form a song-level feature vector. We showed that the framework can effectively

capture manifold acoustic patterns such as harmonic/non-harmonic, steady/transient,

low/high-frequency energy and wide/narrow band. As the learned patterns are se-

lectively activated for different genres of music due to the sparsity constraint, we can

regard them as low-level musical neurons.

We also conducted comprehensive analysis on how each processing module af-

fects performance. The results show that, while there are small degrees of differences

among feature-learning algorithms, the use of time-frequency AGC, log-frequency

scale (as in mel-frequency spectrogram) and meta parameters (e.g., receptive field

size, dictionary size and max-pooling size) significantly improve classification accu-

racy. This indicates that, in order to achieve good performance, appropriate prepro-

cessing, high-dimensional feature learning and dominant feature selection (by sparsity

and max-pooling) are of paramount importance over using different feature-learning

algorithms.

Finally, we compared our method to previous state-of-the-arts based on hand-

crafted features or feature learning. The results show that, with a simple linear

classifier, our approach achieved high accuracy comparable to the top group and the

CHAPTER 2. FEATURE LEARNING FRAMEWORK 51

best accuracy among those based feature learning on the GTZAN and ISMIR2004

datasets.

Chapter 3

Music Annotation and Retrieval

Using Feature Learning

3.1 Introduction

We presented a feature representation framework and applied it to music genre classi-

fication in Chapter 2. Genre is the most commonly used and acknowledged categorical

concept to distinguish different kinds of music. Thus, the majority of music service

providers organize the music content based on genre so that users can conveniently

browse the music collection. As such, automatic genre classification has been one of

the most popular topics in MIR research.

However, many criticize the inconsistency and ambiguity of genre labels. A genre

usually has a number of sub-genres and they are often created by fusing different

genres. This makes the taxonomy of genre hierarchical and somewhat entangled.

Thus, a song can be recognized as different genres or having multiple genres [65]. Also,

some genre decisions are based upon the extrinsic descriptions of music, for example,

artist’s overall musical characteristic (even if individual songs belong to different

genres), locality of the artists home country or languages and other cultural signals,

rather than the intrinsic musical attributes such as rhythm, timbre or tonality. For

these reasons, it has been argued that the 1-of-K classification problem for automatic

genre classification is an ill-defined problem [2].

52

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 53

James Brown - Give it up or turnit a loose

This is a very danceable song that is arousing/awakening, excit-ing/thrilling and happy. It features strong vocal and fast tempo. Itis a song with high energy and high beat that you might like listen to whileat a party.

Nine Inch Nails - Head like a hole

This is a electronica song that is angry/aggressive but not touch-ing/loving. It features sequencer, drum machine and aggressive vocal.It is a song with high and heavy beat that you might like listen to whiledriving.

Cardigans - Lovefool

This is a pop song that is happy and carefree/lighthearted andlight/playful. It features high-pitched and altered with effects vocal.It is a song with positive feelings that you might like listen to while at aparty.

Table 3.1: Examples of music annotation. The natural language template was bor-rowed from Turnbull [103]. The words in bold are the annotation output generatedby our system.

In addition, as the scale of music collections grows, there is an increasing need

to search for music beyond browsing. This requires associating music with other

high-level semantics, for example, mood, emotion, voice quality, usage and their com-

binations.

Researchers have attempted to overcome the limitations of genre classification by

annotating music with multiple tags that include a variety of words that describe

musical semantics or characteristics. This task is referred to as music annotation or

tag classification and has been actively studied for a past few years.1 While music

genre classification can be treated as a multi-class learning problem, that is, choosing

only one out of many classes, music annotation produces multiple tags to describe

music in diverse contexts as illustrated in Table 3.1. The predicted tags can be

regarded as the most probable words given a song (generative view) or as results of

1Tag classification has been included in MIREX (Music Information Retrieval Evaluation eX-change) since 2008.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 54

Query Top 5 Retrieved Songs

Angry/Aggressive (Emotion)

Nirvana - Aneurysm

Aphex Twin - Come to daddy

Rocket City Riot - Mine tonite

Metallica - One

Liz Phair - Supernova

Female Lead Vocals (Instrument)

Norah Jones - Don’t know why

Dido - Here with me

Sheryl Crow - I shall believe

No doubt - Simple kind of life

Carpenters - Rainy days and mondays

Alternatives (Genre)

Nirvana - Aneurysm

Oasis - Supersonic

Big Star - In the street

Gin Blossoms - Hey jealousy

Rocket City Riot - Mine tonite

Very Danceable (Usage)

LL Cool J - Mama said knock you out

2pac - Trapped

Jane’s addiction - Been caught stealing

Jackson 5 - Abc

Boogie Down Productions - The bridge is over

Table 3.2: Example of text-query based music retrieval. These are retrieval outputsgenerated by our system.

binary classification for the presence of each tag (discriminative view).

Meanwhile, annotation tasks usually produce a confidence score for each tag pre-

diction. For example, this confidence score can be a probability of a tag or a distance

from the decision boundary. This can be used to retrieve music using a text query

by sorting the confidence levels of the query word, i.e., the tags for each song in a

database. Therefore, music retrieval (query-by-text) can be performed as a byproduct

of the music annotation. Table 3.2 shows outcomes of the text-based music retrieval

for several descriptive words.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 55

3.2 Previous Work

Music annotation is similar to genre classification in terms of the data processing

pipeline. Both of systems extract various audio features to characterize music at the

song level and then associate these features with labels. The main difference is that

genre classification is based on a winner-take-all scheme whereas music annotation

allows multiple labels to be selected. Thus, the majority of prior work in music

annotation has focused on multi-labeling algorithms while relying on popular audio

features from genre classification, for example, MFCC, dynamic MFCC and Auditory

Filterbank Temporal Envelope (AFTE).

One class of multi-labeling algorithms is based on fitting data to a generative

model that estimates a multinomial probability distribution for tags given a song.

This include a multi-class naıve Bayes approach [101], a Gaussian mixture model

using a weighted mixture hierarchies expectation-maximization (MixHier) [103] and

Codeword Bernoulli Average (CBA) model [38]. The MixHier model was progressively

improved by considering temporal dependency (Dynamic Texture Mixture) [18] or by

building up existing algorithms such as Decision Fusion [19]. These algorithms were

mainly evaluated on the CAL500 dataset that Turnbull et. al. released as a music

tagging dataset [102].

Another class of multi-labeling algorithms is based on a discriminative model that

finds a decision boundary of each tag using a binary classifier. This classification can

be implemented with SVM [100], AdaBoost [7] and logistic regression [26, 112]. It was

noted that these classifiers focus on binary classification accuracy rather than directly

optimizing the continuous confidence scores and thus the training can be suboptimal

for retrieval tasks [38]. However, recent results show that the distance from the

classifier’s decision boundary can be a reliable means to measure the confidence [26,

112].

While the previous work focused on multi-labeling algorithms, some researchers

were concerned with developing new audio features. In particular, they attempted to

discover audio features using learning algorithms instead of using features designed

with acoustic knowledge. However, most of the feature-learning approaches focused on

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 56

music genre classification as reviewed in the previous chapter whereas there have been

few attempts to apply the learning-based features to music annotation and retrieval.

Hamel and Eck developed DBN-based features for music classification and tagging

[32]. However, their experiment was not fully dedicated to annotation (i.e, tagging)

and retrieval. In addition, their sequel auto-tagging system focused on different types

of temporal pooling on a PCA-whitening domain rather than exploiting manifold

patterns of learned features [33].

In this chapter, we continue to evaluate our data processing pipeline in Chapter2

for music annotation and retrieval. However, leveraging the rich descriptions of tags

(rather than simply using genre), we will interpret the locally learned acoustic features

in more diverse musical semantics. In addition, we will extend the data processing

pipeline by applying DBNs to top of the song-level feature representation.

3.3 Proposed Method

3.3.1 Single Layer Model

We presented a feature-learning framework in Chapter 2. It constructs a song-level

feature in three steps. First, preprocessing is performed to normalize raw audio data

and return mel-frequency spectrogram. Second, high-dimensional sparse features are

locally learned on multiple frames of the mel-frequency spectrogram. Third, the local

feature vectors are max-pooled over a segment and the results are averaged over a

song. As a result, it produces a single high-dimensional vector for each song using an

unsupervised algorithm. For genre classification, we performed multi-class supervised

training with the feature vectors and genre labels using linear SVMs.

We apply this feature-learning framework to music annotation and retrieval as

well. In order to incorporate the framework into the multi-labeling problem, we

slightly modify the supervised training part. Previously, we adopted one-versus-all

scheme for multi-class classification. That is, we used one binary SVM for one genre

label so that only one label is set to 1 and all others to -1 for each song, as specified

in Equation 2.24. This can easily generalize to a multi-labeling problem simply by

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 57

setting more than one label to 1. For example, in genre classification, a typical label

vector y(i) is [1 -1 -1 -1 -1] or [-1 -1 1 -1 -1] where only one genre is selected among

five genres. In music annotation, we can set it to [1 -1 1 -1 -1] or [-1 1 1 -1 1]

where we set 1 when the corresponding tag is selected among five tags. In this way,

using the same objective function shown in Equation 2.24, we can concurrently train

multiple classifiers and perform multi-label classification. In addition, we can compute

distances from the decision boundary in a linear SVM (corresponding to a tag) to a

set of song-level features. They can be used as confidence levels for text-query based

music retrieval.

3.3.2 Extension to Deep Learning

A linear classifier has limitations in finding a complex decision boundary for a label in

the feature space, in our case, the presence of a tag in the song-level feature space. In

order to find a more accurate boundary, nonlinear classifiers using kernels or neural

networks are often chosen. In practice, the kernel approach, often an SVM with Radial

Basis function (RBF) or polynomial kernels, are preferred when building deep neural

networks because network performance is limited as the number of layers increases

[4].2 However, Hinton et. al. recently introduced a greedy layer-wise unsupervised

learning algorithm for the deep neural networks called Deep Belief Networks (DBN)

[36]. This has shown great promise as a strategy to train deep networks. Since then,

a great deal of research has been conducted in the area of neural networks. In Section

2.2, we briefly reviewed the DBN in feature-learning context. Here we apply the DBN

as a way of improving classifier performance.

The DBN is a generative model with many hidden layers. It is trained in an

unsupervised way by “greedy layer-wise stacking” of RBMs, which were introduced

in Section 2.3.3. First, a single layer RBM is trained to model the data. This RBM

learns a set of weights W and biases b, c in Equation 2.19. Then, we fix them as

the parameters of the first layer of the DBN. To learn the next layer of weights

2Deep neural networks are known to have poor training and generalization ability as the numberof hidden layers is greater than one or two. This is often attributed to gradient-based optimizationstarting from random initialization, which may get stuck near poor local optimum.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 58

and biases, we compute the hidden layer representation discovered by the first layer

RBM using Equation 2.20 and apply these outputs as inputs to a new binary-binary

RBM (which has binary input units instead of Gaussian) to learn another layer of

the representation. This returns parameters for the next layer and deeper layers

are learned in a similar fashion. Hinton et. al. showed that the preceding learning

algorithm for a DBN always improves a variational lower bound on the log-likelihood

of the data when training more layers [36].

The DBN learns the most probable parameters to model the input data. The

parameters can be used to initialize deep neural networks instead of using random

initialization. This is often called “pre-training”. After this step, the deep network

can be trained with tags in a supervised way using back-propagation. This is often

called “fine-tuning” as a subsequent step to the pre-training. This approach for

learning deep networks has been shown to be essential for training deep networks.

We apply the DBN to model the song-level feature vectors. Figure 3.1 illustrates

a deep network built on top of our feature-learning framework to form a complete

system to estimate multiple tags. After we obtain the song-level features, we continue

to perform unsupervised learning by greedy layer-wise training up to the final hidden

layer and then fine-tune the network with tag labels using back-propagation. In

our experiment, we employed the L2-regularized L2 hinge-loss of the linear SVM in

Equation 2.24 with additional regularization terms of the network parameters as a

penalty function. This makes the penalty term consistent between classifiers. That

way, performance difference can be attributed only to the inclusion of the hidden

layer. Note that this architecture can be seen as a special case of a convolutional

deep network that connects a song (a long sequence of data) to a set of tags [50, 54].

However, the layers below and above the song-level feature have very different flavors.

Specifically, below the song-level feature, we focus on high-dimensional sparse feature

learning using a single layer on local data. Also, the layer remains unsupervised.

On the other hand, above the song-level feature, we perform dense feature learning

through multiple layers without sparsity and they are eventually supervised using the

tag information.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 59

Tag 1 Tag 2 Tag 3 Tag 4

...

Mel!freq. Spectrogram

Seconds

Mel!f

req b

ins

0 0.5 1 1.5 2 2.5

20

40

60

80

100

120

Feature Bases

Hidden Layer Index

Mel!f

req b

ins

20

40

60

80

100

120

Hidden Layer Activation

Seconds

Hidd

en La

yer I

ndex

0 0.5 1 1.5 2 2.5

50

100

150

200

250

Song-Level Feature Vector

...

Mel!freq. Spectrogram

Seconds

Mel!

freq

bins

0 0.5 1 1.5 2 2.5

20

40

60

80

100

120

Feature Bases

Hidden Layer Index

Mel!

freq

bins

20

40

60

80

100

120

Hidden Layer Activation

Seconds

Hidd

en L

ayer

Inde

x

0 0.5 1 1.5 2 2.5

50

100

150

200

250

... ...

Local Sparse Features

Mel-Frequency Spectrogram

Max-Pooling / Aggregation

Feature Encoding

Hidden Layers

Figure 3.1: Feature-learning architecture using deep learning for multi-labeling clas-sification. A deep belief network is used on top of the song-level feature vectors andthen the network is fine-tuned with the tag labels.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 60

3.4 Experiments

3.4.1 Datasets

We evaluated our multi-labeling system on CAL500, which is one of the most popular

datasets in music annotation and retrieval. CAL500 contains 502 western songs, each

of which was manually annotated with one or more tags out of 174 possibilities.

The tags are grouped into 6 categories: Mood, Genre, Instrument, Song, Usage, and

Vocal [102]. In our experiments, we used 97 tags with at least 30 example songs and

performed 5 fold cross-validation to compare results with those reported in previous

works. In order to apply the full path of our pipeline, we obtained MP3 files of the

502 songs and used the decoded waveforms.3

3.4.2 Preprocessing Parameters

We first resampled the waveform data to 22.05kHz and applied the time-frequency

AGC using 10 sub-bands and temporal smoothing the envelope on each band. We

computed an FFT with a 46ms Hann window and 50% overlap. This produces a 513

dimensional vector (up to half the sampling rate) for each frame. We then converted

it to a mel-frequency spectrogram with 128 bins. For the magnitude compression, C

was set to 10 (see Section 2.3.1).

3.4.3 MFCC

We also evaluated MFCC as a “hand-crafted” feature in order to compare it to

our proposed feature representation. Instead of using the MFCC provided from the

CAL500 dataset, we computed our own MFCC to match parameters as close as pos-

sible to the proposed feature. We used the same AGC and FFT parameters but 40

bins for the mel-frequency spectrogram and then applied log and DCT. In addition,

we formed a 39-dimensional feature vector by combining the delta and double delta

vector and normalized it by making the 39-dimensional vector have zero mean and

3Note that the decoded waveforms may be different from the original waveforms although theywill be perceptually very similar to each other.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 61

unit variance. The MFCC vector was also fed into either the classifier directly or the

feature-learning step.

3.4.4 Feature-Learning Parameters

For the PCA whitening and feature-learning steps, we sampled 100000 data examples,

approximately 200 examples at random positions within each song 4. Each example

is selected as a 128 × n (n=1, 2, 4, 6, 8 and 10) patch from the mel-frequency

spectrogram. Using PCA whitening, we reduced the dimensionality of the examples

to retain 90% of the variance. Before the whitening, we added 0.01 to the variance

for regularization. We used dictionary size (or hidden layer size) and sparsity (when

applicable) as the primary feature-learning meta parameters. The dictionary size was

fixed to 1024. The sparsity parameter was set to ρ = 0.007, 0.01, 0.02, 0.03, 0.05, 0.07

and 0.1 for sparse RBM and λ = 1.0, 1.5, 2.0 and 2.5 for sparse coding. Max-pooling

was performed over segments of length 0.05, 0.1, 0.25, 0.5, 1, 2, 4, 8, 16, 32 and 64

seconds. Note that we performed the PCA whitening and feature learning only with

a training set, that is, separately for each fold of cross validation.

3.4.5 Classifier Parameters

We first normalized the song-level features of the training set by subtracting the mean

and dividing by the standard deviation. We then trained the classifiers with the

features and hard annotation. All parameters in preprocessing and feature-learning

stages were tuned using linear SVMs. Using the best parameter set, we replaced the

single-layer classifier with deep neural networks with up to three hidden layers. In

order to verify the effectiveness of the DBN, we compared the pre-training by DBN

with random initialization for the deep neural networks. In the deep neural network,

all hidden layer sizes were fixed to 512 units. Finally, to suppress frequently used tags

and thus makes more balanced predictions, we adjusted the distance to the decision

4In our previous work, we sampled examples only once from the whole dataset. In this experiment,we sampled them from the training set and learned parameters separately for each fold. Thus, theresults are slightly different from those in [71].

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 62

boundary by subtracting the mean times diversity factor (1.25) from the classifier

output, following the heuristic in [38].

3.5 Evaluation and Discussion

3.5.1 Visualization

Figure 3.2 shows feature bases learned from the CAL500 dataset using a sparse RBM.

Using the procedure in Section 2.5.1, given a tag label we searched 20 most active

feature bases and organized them for each group of tags. As previously shown with

the genre dataset, they are distinguished by different timbral patterns. In particular,

by virtue of the rich words of musical semantics in the dataset, the results effectively

demonstrate that the semantic descriptions are associated with the local acoustics

patterns.Table 3.3 summarizes the relationships for selected tags. It suggests that

songs with a specific tag activate certain feature bases more frequently. This demon-

strates that the feature bases are selectively activated depending on the semantics of

music, thereby helping discriminate music at the high-level.

3.5.2 Evaluation Metrics

We evaluated the annotation task using precision, recall and F-score, following pre-

vious work. Precision and recall were computed based on the methods described by

Turnbull [103]. The F-score was computed by first calculating individual F-scores for

each tag and then averaging the individual F-scores, similarly to what was done by

Ellis [26]. It should be noted that averaging individual F-scores tends to generate

lower average F-score than computing the F-score from mean precision and recall val-

ues. As for the retrieval, we used the area under the receiver operating characteristic

curve (AROC), mean average precision (MAP) and top-10 precision (P10) [26].

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 63

Angry/Agressive

Mel−

frequency

bin

20

40

60

80

100

120

Calming/Soothing

20

40

60

80

100

120

Exciting/Thrilling

20

40

60

80

100

120

Happy

20

40

60

80

100

120

Sad

20

40

60

80

100

120

(a) Emotion

Aggressive

Mel−

frequency

bin

20

40

60

80

100

120

High−pitched

20

40

60

80

100

120

Low−pitched

20

40

60

80

100

120

Rapping

20

40

60

80

100

120

Screaming

20

40

60

80

100

120

(b) Vocal

ElectricGuitar(distorted)

20

40

60

80

100

120

DrumMachine

20

40

60

80

100

120

FemaleLeadVocals−Solo

20

40

60

80

100

120

MaleLeadVocals−Solo

20

40

60

80

100

120

Trumpet

20

40

60

80

100

120

(c) Instrument

Driving

Me

l−fr

eq

ue

ncy b

in

20

40

60

80

100

120

Reading

20

40

60

80

100

120

Romancing

20

40

60

80

100

120

Sleeping

20

40

60

80

100

120

Wakingup

20

40

60

80

100

120

(d) Usage

Figure 3.2: Top 20 most active feature bases (dictionary elements) learned by a sparseRBM for different emotions, vocal quality, instruments and usage categories of theCAL500 set.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 64

Tag Description on feature bases

Angry/Aggressive Wideband energy with strong high-frequencycontent or extreme low-frequency content

Calm/Soothing and Sleeping Low-frequency content with harmonic patterns

Low-pitched and High-pitched Juxtaposed with low- and high-pitched harmonicpatterns

Rapping and DrumMachine Extremely low-freq. energy and several wide-band and transient patterns

Exciting/Thrilling Non-harmonic and transient patterns

Table 3.3: This table describes the acoustic patterns of the feature bases that areactively “triggered” in songs with a given tag. The corresponding feature bases areshown in Figure 3.2.

3.5.3 Results and Discussion

We examine the effect of preprocessing and feature-learning algorithms on the anno-

tation and retrieval performance. Also, we show improved results with deep neural

networks. Finally, we compare our best results to those of state-of-the-art algorithms.

Input Data, Algorithms and AGC

Table 3.4 summarizes results on features obtained with different types of input data

and feature-learning algorithms. First of all, the mel-frequency spectrogram signifi-

cantly outperforms MFCC regardless of the type of learning algorithms. This indi-

cates that capturing rich acoustic patterns (not just timbre with MFCC but also pitch

and harmony) from mel-frequency spectrogram is necessary to effectively associate

sound with musical semantics. Among the feature-learning algorithms, K-means and

sparse RBM generally perform better than sparse coding, which is somewhat dif-

ferent from the genre classification result in the previous chapter. In addition, the

results show that the time-frequency AGC significantly improves both annotation and

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 65

Annotation Retrieval

Data+Algorithm Prec. Recall F-score AROC MAP P10

With AGC

MFCC only 0.399 0.223 0.242 0.713 0.446 0.467

MFCC+K-means 0.446 0.240 0.270 0.732 0.471 0.492

MFCC+SC 0.437 0.232 0.260 0.713 0.452 0.476

MFCC+SRBM 0.441 0.235 0.263 0.725 0.463 0.485

Mel-Spec+K-means 0.467 0.253 0.289 0.740 0.487 0.515

Mel-Spec+SC 0.458 0.250 0.283 0.733 0.481 0.509

Mel-Spec+SRBM 0.474 0.258 0.290 0.741 0.489 0.513

Without AGC

MFCC only 0.399 0.222 0.239 0.712 0.444 0.460

MFCC+K-means 0.438 0.237 0.267 0.727 0.465 0.489

Mel-Spec+SRBM 0.458 0.246 0.275 0.727 0.478 0.506

Table 3.4: Performance comparison for different input data and feature-learning al-gorithms. These results are all based on linear SVMs.

retrieval performance, regardless of the input features.

Receptive Field Size

Figure 3.3 plots F-score and AROC for different receptive field size, that is, the

number of frames taken from the mel-frequency spectrogram. It shows that the

performance significantly increases between 1 and 4 frames and then saturates beyond

4 frames. It is interesting that the best results are achieved at 6 frames (about

0.16 second long). We think this is related to the representational power of the

algorithm. That is, when the number of frames is small, the algorithm is capable of

capturing the variation of input data. However, as the number of frames grows, the

algorithm becomes incapable of representing the exponentially increasing variation,

in particular, temporal variation.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 66

1 2 4 6 8 100.26

0.265

0.27

0.275

0.28

0.285

0.29

0.295

Number of frames

F−

sco

re

SC

RBM

Kmeans

(a) F-score

1 2 4 6 8 100.71

0.715

0.72

0.725

0.73

0.735

0.74

0.745

0.75

Number of frames

AR

OC

SC

RBM

Kmeans

(b) AROC

Figure 3.3: Effect of number of frames. The dictionary size is set to 1024.

Sparsity and max-pooling size

Figure 3.5 plots the F-score for a set of sparsity values and max-pooling sizes. It shows

a clear trend that higher accuracy is achieved when the feature vectors are sparse

(around 0.02) and max-pooled over segments of about 16 seconds.5 These results

indicate that the best discriminative power in song-level classification is achieved by

capturing only a few important features over both timbral and temporal domains.

Deep Learning

Table 3.5 shows results when deep neural networks are used. The preprocessing

and feature-learning parameters were chosen from the best result with linear SVMs.

In general, pre-training outperforms random initialization regardless of the number

of hidden layers. With random initialization, the best result was obtained with a

single hidden layer (Mel-SRBM-NN1) and, as more hidden layers are used, the result

became worse. This is probably because it becomes more difficult to find a good

5We found that the average length of songs on the CAL500 dataset is approximately 250 seconds,which suggests that aggregating about 16 (≈ 250/16) max-pooled feature vectors over an entire songis an optimal choice.

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 67

0.007 0.01 0.02 0.03 0.05 0.07 0.10.28

0.282

0.284

0.286

0.288

0.29

Sparsity

F−

sco

re

(a) F-score

0.007 0.01 0.02 0.03 0.05 0.07 0.10.732

0.734

0.736

0.738

0.74

0.742

0.744

SparsityA

RO

C

(b) AROC

Figure 3.4: Effect of sparsity (sparse RBM).

0.1 0.25 0.5 1 2 4 8 16 32 640.265

0.27

0.275

0.28

0.285

0.29

Max−pooling [sec]

F−

sco

re

(a) F-score

0.1 0.25 0.5 1 2 4 8 16 32 640.728

0.73

0.732

0.734

0.736

0.738

0.74

0.742

0.744

Max−pooling [sec]

AR

OC

(b) AROC

Figure 3.5: Max-pooling of sparsity (sparse RBM).

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 68

Annotation Retrieval

Classifiers Prec. Recall F-score AROC MAP P10

Linear SVM

Mel-SRBM-SVM 0.479 0.257 0.289 0.741 0.489 0.513

Neural networks with hinge loss (random initialization)

Mel-SRBM-NN1 0.470 0.256 0.291 0.756 0.505 0.531

Mel-SRBM-NN2 0.481 0.244 0.285 0.749 0.498 0.531

Mel-SRBM-NN3 0.467 0.236 0.280 0.728 0.479 0.500

Neural networks with hinge loss (pre-trained by DBNs)

Mel-SRBM-DBN1 0.488 0.260 0.295 0.757 0.508 0.532

Mel-SRBM-DBN2 0.476 0.256 0.291 0.754 0.509 0.546

Mel-SRBM-DBN3 0.476 0.253 0.286 0.752 0.507 0.531

Table 3.5: Performance comparison for linear SVM and neural networks with randominitialization (Mel-SRBM-NN*) and pre-training by DBN (Mel-SRBM-DBN*). Thefigures (1, 2 and 3) indicate the number of hidden layers. The receptive field size wasset to 6 frames.

local optimum as the number of parameter increases. With pre-training by DBN, the

best result was obtained with one and two hidden layers (Mel-SRBM-DBN1 and Mel-

SRBM-DBN2). They achieved the highest accuracy in both annotation and retrieval.

These results show that the DBN is an effective strategy for training a deep neural

network as a classifier.

Comparison to state-of-the-art algorithms

Table 3.6 compares our best results to those of state-of-the-art algorithms from the

group that developed CAL500. The group used MFCC features as input data and

modeled the features using either Gaussian Mixture Model (GMM) as a bag of frames

[103] or Dynamic Texture Mixture (DTM) considering temporal dependency [18].

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 69

Annotation Retrieval

Methods Prec. Recall F-score AROC MAP P10

HEM-GMM [103] 0.374 0.205 0.213 0.686 0.417 0.425

HEM-DTM [18] 0.446 0.217 0.264 0.708 0.446 0.460

BoS-DTM-GMM-LR [26] 0.434 0.272 0.281 0.748 0.493 0.508

DF-GMM-DTM [19] 0.484 0.230 0.291 0.730 0.470 0.487

DF-GMM-BST-DTM [19] 0.456 0.217 0.270 0.731 0.475 0.496

Proposed methods

Mel-Spec-SRBM-SVM 0.474 0.258 0.290 0.741 0.489 0.513

Mel-Spec-SRBM-DBN1 0.488 0.260 0.295 0.757 0.508 0.532

Table 3.6: Performance comparison: state-of-the-art (top) and proposed methods(bottom).

They progressively improved the performance by adding Bag of Systems (BoS) [26]

or Decision Fusion (DF). The results show that our method produces comparable

results to theirs with only a linear SVM and furthermore outperforms the prior arts

in F-score and all retrieval metrics with a nonlinear classifier (neural network pre-

trained by DBN).

3.6 Conclusion and Future Work

We have extended the feature-learning framework to a multi-labeling classification

system by adding a deep neural network as a classifier. Using the rich descriptions

of tag words in the CAL500 dataset, we illustrated the relationship between local

acoustic-feature patterns and high-level semantics in music. By pre-training the deep

neural network with DBNs, we showed that our system outperformed state-of-the-art

algorithms for both annotation and retrieval tasks on the CAL500 dataset. To ensure

the discriminative power of our proposed feature representation method, we need to

CHAPTER 3. MUSIC ANNOTATION AND RETRIEVAL 70

evaluate it on larger datasets, such as, the Million Song Dataset [8] or Magnatagatune

[49].

Chapter 4

Piano Transcription Using Deep

Learning

4.1 Introduction

Music transcription is the task of inferring a symbolic representation (e.g., musical

notes) from audio recordings. Musical notes are often played simultaneously and

thus individual notes interfere with each other by virtue of their harmonic relations.

In addition, timbre, tuning and room conditions vary all the time. The polyphonic

nature of music and the acoustic variations make music transcription a challenging

problem.

A number of methods have been proposed since Moorer first attempted to use

computers to transcribe two voices of different musical instruments [68]. State-of-

the-art algorithms can be divided into three categories: iterative F0 searches, joint

source estimation and classification-based methods. Iterative F0-searches first find

the predominant F0 and then subtract its relevant sources (e.g., harmonic partials)

from the input signal. They repeat this procedure on what remains until no additional

F0s are found [48]. Joint source estimation examines possible combinations of sound

sources by hypothesizing that the input signal is approximated by a weighted sum of

the sound sources with different F0s [31, 95].

While these two categories are based on a generative approach that find a group of

71

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 72

Waveforms Feature Representations

Binary Classifier

Binary Classifier

Binary Classifier. . .

. . .

C4# note on/off

C4 note on/off

B3 note on/off

Figure 4.1: Classification-based polyphonic note transcription. Each binary classifierdetects the presence of a note.

notes that constitute an observed polyphonic tone, classification-based methods are

based on a discriminative approach which detects the presence of a note from the poly-

phonic tone. They usually use multiple binary classifiers, each of which are trained

with short-time acoustic features and single note labels (i.e., note on/off). They

collectively detect multiple notes simultaneously as shown in Figure 4.1. Although

classification-based methods typically make relatively less use of acoustic knowledge,

they showed comparable results to iterative F0 searches and joint source estimation,

particularly for piano music [64, 85].

However, there are two main issues to address in the classification-based methods.

First, the discriminative approach usually requires a large dataset to generalize well

[75]. This requires good feature representations invariant to diverse variations in the

large dataset. Second, each classifier is usually trained separately for each note. This

can be computationally expensive especially when the dimensionality of features is

high or the classifiers have many meta parameters that require cross-validation.

In this chapter, we present a new classification-based algorithm for polyphonic

piano note transcription that considers these two issues. Specifically, we extend a

previous classification-based method in two ways: (1) by using feature representations

learned from spectrogram data and (2) by jointly training the classifiers for multiple

notes. We obtain the learned features using deep belief networks and, in turn, use the

network to concurrently train multiple binary classifiers. We evaluate our approach on

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 73

several public piano transcription datasets and show that our approach outperforms

compared music transcription methods 1.

4.2 Previous Work

The discriminative approach is based on finding an optimal boundary for the presence

of each note by supervised learning. A single note has variations due to timbre, tuning,

room acoustics (e.g., reverberation or background noise) and mixing with other notes.

Previous work has attempted to find the complex boundary by developing robust

features to the acoustic variations and feeding them into sophisticated classifiers,

mainly focusing on polyphonic piano transcription.

Marolt used an auditory filter model and adaptive oscillator networks as a front-

end processing [64]. The auditory filter model emulates the functionality of human

ear producing quasi-periodic firing activities of inner hair cells on a set of frequency

channels and the adaptive oscillator networks track partials and group harmonically

related ones. He used the output of the oscillator network module as a note-detection

feature and applied it to neural networks for supervised training. Poliner and Ellis

proposed a simpler discriminative model. They used a normalized spectrogram as

a timbre-invariant feature and applied it to support vector machine (SVM) with an

RBF kernel [85]. Then, they temporally smoothed the resulting prediction of the

note presence using Hidden Markov Models (HMMs). Boogaart and Lienhart used

a similar approach. They adopted multiple frames of Gabor transform as an input

feature and Adaboost as a note classifier, instead [106].

These methods in common used novel features and non-linear classifiers to find

the complex boundary of note presence. The features were designed by using different

time-frequency transforms and refining them in a hand-tuned manner. In addition,

they sampled positive and negative examples separately for each binary note classifier

and trained them independently. Since the majority of pianos have 88 notes, this

training strategy can be computationally expensive especially when the feature is

high-dimensional or classifiers have many meta parameters.

1This chapter is based on our previous work [72].

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 74

There is increasing interest in learning the feature from data using unsupervised

learning algorithms as an alternative to the hand-tuning approach. Researchers dis-

covered that the learning algorithms can find the underlying structure of data and

provide the outcome as a new feature representation. Furthermore, this adaptively

learned feature was shown to be highly effective in a number of classification tasks.

We apply this feature-learning approach to classification-based polyphonic piano tran-

scription. We will show that this approach is highly effective with only a linear

classifier. Also, we propose a strategy to efficiently train note classifiers.

4.3 Proposed Method

4.3.1 Feature Representation By Deep Learning

A restricted Boltzmann machine (RBM) is an unsupervised learning algorithm that

has two layers: a visible layer and a hidden layer. The visible layer corresponds to

the input data while the hidden layer represents features discovered by training the

RBM. In Chapter 2 and 3, we used a sparse version of RBM where the binary units of

the hidden layer are constrained to be parsimoniously activated. We showed that the

sparse RBM captures acoustic patterns that explain musical signals and the outcome

was used as a novel feature representation for music classification.

We apply the sparse RBM to piano sounds in order to discover their harmonic

patterns and use them for classification-based polyphonic piano transcription. Specif-

ically, we model single frames of normalized spectrogram with the sparse RBM; the

visible layer corresponds to a vector of the spectrogram frame. Figure 4.2 illustrates

the feature bases learned from a large piano dataset using the sparse RBM. Most bases

capture harmonic distributions, which correspond to various pitches while some con-

tain non-harmonic patterns. Also, note that the feature bases show exponentially

growing curves for each harmonic partial. This verifies the structure of the piano

sound, i.e., the logarithmic scale of musical notes.

On top of the first RBM, we stack another RBM to find more complex dependency

in piano sounds, for example, different combinations of the features shown in Figure

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 75

Figure 4.2: Feature bases learned from a piano dataset using a sparse RBM. Theywere sorted by the frequency of the highest peak. Most bases capture harmonicdistributions which correspond to various pitches while some contain non-harmonicpatterns. Note that the feature bases show exponentially growing curves for eachharmonic partial. This verifies the logarithmic scale in the piano sound.

4.2. This is performed by greedy layer-wise training, that is, using the feature data

from the first RBM to train another RBM. This deep learning algorithm is called a

deep belief network (DBN) [36]. In Chapter 3, we used DBNs as a means to achieve

“better initialization” for neural networks in the context of supervised training. Thus

we compared it to randomly initialized neural networks. Here we intend to use them as

a multi-layer feature-learner. Thus, we first examine the DBN output (the top hidden

layer of the pre-trained network) as a feature representation for note classifiers. Then,

we will fine-tune the network by back-propagating the errors from the classifiers. In

our experiments, we evaluated up to two layers of DBNs and compared the pre-trained

network to the fine-tuned one.

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 76

4.3.2 Training Strategy

A piano usually has 88 notes, each of which is detected by the corresponding binary

classifier. They can be trained either separately for each note or jointly as a multi-

labeling problem. We term the two training strategies as single-note training and

multiple-note training, respectively, and describe them below.

Single-note Training

The majority of previous classification-based methods trained classifiers individually

for each note. For example, Poliner and Ellis’ piano transcription system consists

of 87 independent support vector machine (SVM) classifiers with an RBF kernel.

They formed the training data by selecting spectrogram frames that include the note

(positive examples) and those that do not include it (negative examples). They

randomly sampled 50 positive (when available) and negative examples from each

piano song per note and trained the SVM with separate training data for each note.

We also examine this single-note training strategy. However, instead of a nor-

malized spectrogram along each frequency axis that they used, we apply DBN-based

feature representations of spectrogram frames. In addition, we constrained the SVM

to a linear kernel because they reported that the RBF kernel provided only modest

performance gains with significantly more computation [86] and also a linear SVM

is more suitable to large-scale data. The left column of Figure 4.3 illustrates our

approach for single-note training. Our proposed method transforms the spectrogram

frames into mid-level features via one or two layers of learned networks and then feeds

them into the classifier. As an additional step, we fine-tune the network using the

error from the linear SVM. We compare this with a baseline model that directly feed

spectrogram frames into the SVM.

Multiple-note Training

While examining the single-note training, we observed that the trained classifiers tend

to be somewhat “aggressive”. In other words, they produced more “false alarm” errors

(detection of inactive notes as active ones) than “miss” errors (failure to detect active

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 77

notes). In particular, this significantly degraded onset accuracy. Also, the training

was slow because the deep networks had to be fine-tuned separately for each note.

For this reason, we attempted to train all binary classifiers concurrently, referring to

this as multiple-note training.

The idea is that we can view the polyphonic piano transcription as a multi-labeling

problem, that is, labeling 88 binary note-on/off tags given an audio feature. We have

already handled this problem for music annotation in Chapter 3, where we performed

the training by summing multiple SVM objectives using shared features and the

binary label vectors.2 This allows cross-validation to be jointly performed for the

combined SVMs, thereby saving a significant amount of training time. On the other

hand, this requires a different way of sampling examples because the training data is

shared by all binary classifiers. Since we combined all 88 notes in our experiments, all

spectrogram frames except silent ones are a positive example for at least one SVM.

Thus we sampled the training data by simply selecting every K spectrogram frame.

K was set to 16 as a trade-off between data reduction and performance. Note that

this makes the ratio of positive and negative examples for each SVM determined

by occurrences of the note in the whole training set, thereby having significantly

more negative examples than positive ones for most SVMs. It turned out that this

“unbalanced” data ratio makes the classifiers “less aggressive,” as a result, increasing

overall performance.

The right column of Figure 4.3 illustrates the multiple-note training. Before the

fine-tuning, the binary classifiers, in fact, do not influence each other. However, when

the fine-tuning is performed, the errors of the classifiers collaboratively update the

shared network. In other words, the presence of multiple notes such as C3 or C4

jointly updates the learned features, improving the overall performance.

4.3.3 HMM Post-processing

The note on/off classification described above treats training examples independently

without considering dependency between neighboring frames. We temporally smooth

2This approach is described as a multi-labeling problem in Section 3.3.1. The objective functionis denoted in Equation 2.24.

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 78

...

... Input

Multiple-NoteTraining

Linear SVM

(Baseline)

Linear SVM

+ Hidden Layers

...

Single-NoteTraining

Output

...

...

...

...

...

...

Output

Input

HiddenLayers

...

Figure 4.3: Network configurations for single-note and multiple-note training. Fea-tures are obtained from feed-forward transformation as indicated by the bottom-uparrows. They can be fine-tuned by back-propagation as indicated by the top-downarrows.

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 79

the output of classifiers using HMM-based post-processing, following Poliner and Ellis.

We model each note independently with a two-state HMM and modified the SVM

output (distance to the decision boundary) to obtain a posterior probability:

p(yi = 1|xi) = sigmoid(α(θTxi)), (4.1)

where xi is a feature vector, θ are SVM parameters, yi is a label and α is a scaling

constant. α was chosen from a pre-determined list of values as part of the cross-

validation stage. For each note class, the smoothing process was performed by running

a Viterbi search based on a 2x2 transition matrix, a note on/off prior (for obtained

from the training data) and the posterior probability.3

Figure 4.4 shows the overall signal transformation through the DBN networks,

SVM classifiers and HMM post-processing. Note that the original spectrogram grad-

ually changes, getting more similar to the final output.

4.4 Experiments

4.4.1 Datasets

We used three publicly available piano datasets to evaluate our approach.

Poliner and Ellis

This data set consists of 124 MIDI files of classical piano music. They were rendered

into 124 synthetic piano sound and 29 real piano recordings [85]. We used the first

60-second excerpt of each song.

MAPS

MIDI-Aligned Piano Sounds (MAPS) is a large piano dataset that includes various

patterns of playing and pieces of music [27]. We used 9 sets of piano pieces, each with

3We used HMMTool box found at http://www.cs.ubc.ca/~murphyk/Software/HMM/hmm.html

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 80

Input (Spectrogram ) Hidden layer activation

HMM outputSVM output

Time (10ms)

Freq

uenc

y [k

Hz]

100 200 300 4000

0.5

1

1.5

Time (10ms)H

idde

n un

it in

dex

100 200 300 400

50

100

150

200

250

Time (10ms)

MID

I not

e nu

mbe

r

100 200 300 400

40

60

80

100

Time (10ms)

MID

I not

e nu

mbe

r

100 200 300 400

40

60

80

100

Figure 4.4: Signal transformation through our system. The original spectrogramgradually changes to the final output via the deep network, SVM classifiers andHMM-based smoothing

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 81

30 songs. They were created by various high-quality software synthesizers (7 sets)

and Yamaha Disklavier (2 sets). We used the first 30-second excerpt of each song in

the validation and test sets but the same length at a random position for the training

set.

Marolt

Marolt provided a small number of piano examples that consists of 3 synthetic piano

and 3 real piano recordings [64]. This small dataset was used only in the testing

phase.

4.4.2 Pre-processing

We first computed spectrograms from the datasets with a 128ms window and 10ms

overlaps. To remove note dynamics, we normalized each column by dividing entries

by their sum, and then compressed it using a cube root, commonly used as an ap-

proximation to the loudness sensitivity of human ears. Furthermore, we applied PCA

whitening to the normalized spectrogram, retaining 99% of the training data variance

and adding 0.01 to the variance before the whitening. This yielded roughly 50-60%

dimensionality reduction and lowpass filtering in the PCA domain. The ground truth

was created from the MIDI files. We extended note offset times by 100ms in all train-

ing data to account for room reverberation in the piano recordings. The extended

note length was experimentally determined.

4.4.3 Unsupervised Feature Learning

We trained the first and second-layer DBN representations using the pre-processed

spectrogram. The hidden layer size was chosen as 256 and the expected activation

of hidden units (sparsity) was cross-validated over 0.05, 0.1, 0.2 and 0.3, while other

parameters were kept fixed.

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 82

4.4.4 Evaluation Metrics

We primarily used the following metric of accuracy:

Accuracy =TP

FP+FN+TP, (4.2)

where TP (true positive) is the number of correctly predicted examples, FP (false pos-

itives) is the number of note-off examples transcribed as note-on, FN (false negative)

is the number of note-on examples transcribed as note-off. This metric is used for

both frame-level and onset accuracy. Frame-level accuracy is measured by counting

the correctness of frames every 10ms, and onset accuracy is by searching a note onset

of the correct pitch within 100 ms of the ground-truth onset. In addition, we used the

F-measure for frame-level accuracy to compare our results to those published using

the metric.

4.4.5 Training Scenarios

Our method is evaluated in two different scenarios. In the first scenario, we primarily

used the Poliner and Ellis set, splitting it into training, validation and test data fol-

lowing [85]. In order to avoid overfitting to the specific piano set, we selected 26 songs

from two synthesizer pianos sets in MAPS and used them as an additional validation

set. For convenience, we refer to this subset as MAPS2. In the second scenario, we

used five remaining synthesizer piano sets in MAPS for training to examine if our

method generalizes well when trained on diverse types of timbre and recording condi-

tions. For validation, we randomly took out 26 songs from the five piano sets, calling

them MAPS5 to distinguish it from the actual training data. We additionally used

MAPS2 for validation in the second scenario as well.4

4The lists of MAPS songs for training, validation and test are specified in http://ccrma.

stanford.edu/~juhan/ismir2011.html

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 83

4.5 Evaluation

4.5.1 Validation Results

We compare the baseline feature (normalized spectrogram by cube root) to the first-

and second-layer DBN features and their fine-tuned versions on validation sets in the

two scenarios. The results are shown in Figure 4.5 and 4.6.

In scenario 1, DBN features generally outperform the baseline. In single-note

training, fine-tuned L1-features give the highest accuracy on both validation sets. In

multiple-note training, unsupervised L1- or L2-features achieve slightly better results.

In a comparison of the two training methods, either one appears to be not superior

to the other, showing subtle differences: Multiple-note training gives slightly better

results when the same piano set are used for validation (Poliner and Ellis), whereas

single-note training does a little better job when different pianos set (MAPS2) are

used.

In scenario 2, the results show that DBN L1-features always achieve better results

than the baseline but DBN L2-features generally give worse accuracy. Fine-tuning

always improves results on both validation sets, although the increment is very lim-

ited on MAPS2 in multiple-note training. In comparison of the two training methods,

multiple-note training outperforms single-note training for both validation sets, par-

ticularly giving the best accuracy on MAPS2. The superiority of multiple-note train-

ing is even more apparent in onset accuracy as shown in Figure 4.6. This is because

the multiple-note training is less aggressive and therefore it has less false alarms.

Figure 4.7 shows the influence of sparsity (hidden layer activation in RBMs)

on frame-level accuracy. The accuracy is the average value on two validation sets

(MAPS5 and MAPS2) when L1 features are used in multiple-note training and sce-

nario 2. The results indicate that relatively less sparse features perform better before

fine-tuning; however, with fine-tuning, sparse features achieve the highest accuracy

as well as the best improvement.

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 84

40

50

60

70

80

90

Acc

urac

y (%

)

Single−note training Multiple−note training

Poliner Ellis

MAPS2

Poliner Ellis

MAPS2

Baseline

L1

L1−finetuned

L2

L2−finetuned

(a) Scenario 1

40

50

60

70

80

90

Acc

urac

y (%

)

Single−note training Multiple−note training

MAPS5

MAPS2

MAPS5

MAPS2

(b) Scenario 2

Figure 4.5: Frame-level accuracy on validation sets in two scenarios. The first andsecond-layer DBN features are referred to as L1 and L2.

4.5.2 Test Results: Comparison With Other Methods

The validation results show that a single layer of DBN is the best-performing feature

representation and multiple-note training is better than single-note training. Thus, we

chose DBN L1-features and multiple-training to test our system. Also, we evaluated

both unsupervised and fine-tuned features.

Table 4.1 shows results on the Poliner and Ellis test set, and Marolt set. We

divided the table into two groups to make a fair comparison. The upper group uses

the same dataset for both training and testing (the Poliner and Ellis set) whereas

the lower group assumes that the piano tones in the test sets were “unheard” in

training or uses different transcription algorithms. In the upper group, Poliner and

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 85

10

20

30

40

50

60

70

80

Acc

urac

y (%

)

Single note training Multiple note training

MAPS5

MAPS2

MAPS5MAPS2

(a) Scenario 1

10

20

30

40

50

60

Acc

urac

y (%

)

Single−note training Multiple−note training

MAPS5

MAPS2

MAPS5

MAPS2

(b) Scenario 2

Figure 4.6: Onset accuracy on validation sets in two scenarios.

0 0.05 0.1 0.15 0.2 0.25 0.362

64

66

68

Sparsity

Acc

urac

y (%

)

L1

L1−finetuned

Figure 4.7: Frame-level accuracy VS sparsity (hidden layer activation in RBMs)

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 86

Algorithms P. and E. Marolt

Poliner and Ellis [85] † 67.7% 44.6%

Proposed (S1-L1) 71.5% 47.2%

Proposed (S1-L1-fine-tuned) 72.5% 46.45%

Marolt [64] † 39.6% 46.4%

Ryyananen and Klapuri [87] † 46.3% 50.4%

Proposed (S2-L1) 63.8% 52.0%

Proposed (S2-L1-fine-tuned) 62.5% 51.4%

Table 4.1: Frame-level accuracy on the Poliner and Ellis, and Marolt test set. Theupper group was trained with the Poliner and Ellis train set while the lower groupwas with other piano recordings or uses different methods. S1 and S2 refer to trainingscenarios. †These results are from Poliner and Ellis [85].

Ellis’ transcription system adopted a normalized spectrogram and a non-linear SVM.

Our method outperformed their approach for both test sets. In the lower group, our

method trained with MAPS (scenario 2) also produced better accuracy than the two

published results on both sets. Note that, in both groups, unsupervised features give

better results than fine-tuned features when different piano sets are used for training

and testing. As for onset accuracy, we achieved 62% in training scenario 1 on the

Poliner and Ellis test set, which is very close to the Poliner and Ellis’ result (62.3%).

Table 4.2 compares our method with other algorithms evaluated on the MAPS

test set, composed of 50 songs selected from the two Disklavier piano sets by [107].

The fine-tuned DBN-features in our method give the highest frame-level accuracy

among compared methods.

4.6 Discussion and Conclusions

We have applied DBNs to classification-based polyphonic piano transcription. The

results show that learned feature representations by DBNs, particularly L1 features,

CHAPTER 4. PIANO TRANSCRIPTION USING DEEP LEARNING 87

Algorithms Precision Recall F-measure

Marolt [64] † 74.5% 57.6% 63.6%

Vincent et al. [107] † 71.6% 65.5% 67.0%

Proposed (S2-L1) 80.6% 67.8% 73.6%

Proposed (S2-L1-ft.) 79.6% 69.9% 74.4%

Table 4.2: Frame-level accuracy on the MAPS test set in F-measure. “ft” stands forfine-tuned. †These results are from Vincent et. al. [107].

provide better transcription performance than the baseline and our classification ap-

proach outperforms compared piano transcription methods. Our evaluation shows

that fine-tuning generally improves accuracy, particularly when sparse features are

used. However, unsupervised features often work better when the system is tested on

different piano sets. This indicates that unsupervised features generalize well, being

robust to acoustic differences among piano sounds.

We also suggested multiple-note training. Compared to single-note training, this

method improved not only transcription accuracy but also reduced training time

by concurrently cross-validating multiple classifiers. In our computing environment,

multiple-note training was more than five times faster than single-note training when

the DBNs are fine-tuned.

Our method is based on frame-level feature learning and binary classification

under simple two-state note event modeling. We think that more refinements will be

possible by modeling richer states to represent dynamic properties of musical notes.

Chapter 5

Conclusions

We have presented feature representations using learning algorithms and successfully

applied them to several content-based MIR tasks; genre classification, music anno-

tation/retrieval, and polyphonic piano transcription. These tasks were in common

posed as classification problems, where it is essential to have good features in or-

der to facilitate supervised training and achieve high performance. Conventionally,

audio features have been hand-tuned on an ad-hoc basis using domain knowledge.

Throughout this thesis, we developed novel audio features based on learning algo-

rithms and showed promising results over the hand-tuned features. In this chapter,

we summarize our contributions and reviews the results. Lastly we discuss ideas for

future work.

5.1 Contributions and Reviews

Learning feature representations has been suggested as a new paradigm in machine

learning and actively exploited in a variety of machine perception tasks such as com-

puter vision and speech recognition. We dare say that the work presented in this

thesis makes meaningful contributions toward the effort, particularly in the area of

music information retrieval. They can be summarized as follows:

- We proposed a data processing pipeline to effectively learn features for music

classification. In particular, we showed that appropriate preprocessing such

88

CHAPTER 5. CONCLUSIONS 89

as automatic gain control, time-frequency transform and receptive field size

(selecting multiple frames) is essential to facilitate the feature learning and

improve performance.

- We demonstrated that feature-learning algorithms capture the rich timbral pat-

terns of musical signals. For music genre classification and annotation/retrieval,

we took advantage of them by associating the learned features with high-level

musical semantics such as genre, emotion, song/voice quality and usage. For

polyphonic piano transcription, we illustrated that the learned patterns mainly

contain harmonic information.

- We showed that the learned feature representations can be superior to popularly

used hand-engineered features. Using the proposed data processing pipeline, we

achieved comparable results to state-of-art algorithms or outperformed them on

publicly available datasets in the three different content-based MIR tasks.

Note that we relied on some acoustic knowledge in preprocessing such as automatic

gain control or mel-frequency scale in spectrogram. This is against the key idea

of feature learning. It could have been more ideal if we applied feature-learning

algorithm directly to raw waveforms without any preprocessing. However, some front-

end processing such as AGC is practically necessary to normalize datasets and the use

of acoustic knowledge was also constrained so that the processed outputs are highly

reconstructible (We have audio examples reconstructed from the output of each stage

in the data processing pipeline. Appendix B has the information). We leave “feature

learning from raw audio data” as future work.

We evaluate several different feature-learning algorithms and compare them to

prior state-of-the-arts in Chapter 2 and 3. A notable result is that there is no abso-

lutely outstanding feature-learning algorithm. For example, sparse coding achieved

the best accuracy in music genre classification whereas it produced worst results

among compared algorithms in music annotation/retrieval. On the other hand, RBM

CHAPTER 5. CONCLUSIONS 90

was intermediate in genre classification whereas it outperformed others in music an-

notation/retrieval. Furthermore, our experiments showed that selection of meta pa-

rameters, in particular, dictionary size, sparsity and max-pooling size is far more

important than using different algorithms. This indicates that new feature algo-

rithms need to be developed focusing on more practical aspects such as training time,

fast encoding or less meta parameters than algorithm itself [76, 15].

5.2 Future Work

While carrying out our experiments, we had ideas to further explore using feature

learning in audio applications. Here we list some of them as future work.

Learning Features from Waveforms

As stated above, we applied feature-learning algorithms to preprocessing outputs (i.e.

mel-frequency spectrogram). We would like to learn features from raw audio data

without any use of acoustic knowledge. There are a few previous works that attempted

to learn the front-end processing from waveforms using RBM [43, 44] or sparse coding

[9, 63]. However, they need to be refined more to be used in various content-based

MIR tasks. Moreover, since musical signals are highly complex (e.g., mixed with many

sound sources), algorithms in this approach will need more constraints to remove

unnecessary variations.

Hierarchical Feature Learning

In Chapter 2 and 3, we focused on single-layer feature learning that captures local

timbral or pitch/harmonic dependency, and summarized the outcomes directly into a

song-level feature. This approach lacks of capturing mid-level patterns such as rhythm

or chord progression. These mid-level features could be learned in a hierarchical and

convolutional manner, for example, by adding another feature-learning layer on top

of the max-pooled output.

CHAPTER 5. CONCLUSIONS 91

Similarity-based Audio Retrieval

We evaluated our data processing pipeline only in classification settings. Since the

pipeline produces a song-level feature vector, it can be directly applied to similarity-

based tasks. For example, similar songs can be searched by computing a distance

between two song-level feature vectors such as cosine, L1, L2 and KL-divergence.

Furthermore, as a more elaborated way, we could apply semantic hashing by using

a deep learning model on the song-level feature vector (similar to the deep learning

model in Chapter 3 but using only unsupervised learning algorithms) [88].

Onset Detection

We observed that the local feature representations under high sparsity are likely to

be more strongly activated when the input signal is non-stationary, for example, note

onsets, pitch-modulation or other abrupt temporal changes. In other words, the sum

of the local feature representation over the feature dimension tends to fluctuate more

at such non-stationary points. This is somewhat similar to what is observed in human

sensory system; many areas of the brain (i.e., neurons) are mostly silent but highly

activated when an unexpected or abrupt stimulus comes in. In addition, a recent

work presented a potential use of deep learning for tempo estimation [40]. Hence, we

need to further explore feature learning in this aspect.

Appendix A

Real-time Music Tagging

Visualizer

A.1 Introduction

The majority of MIR research presents the results with evaluation metrics (e.g., F-

score and AROC), showing the metrics as average values over a dataset. While this

approach is convenient for comparing algorithms, it has limitations in demonstrating

how the algorithm works and providing attractive presentation. For this reason, we

implemented our music annotation work as a real-time visualization system.

The system basically displays each stage of the data processing pipeline in real-

time with the original waveform, mel-frequency spectrogram, learned feature repre-

sentation and tag prediction for a song. The tag size increases or decreases in time

depending on the confidence level of prediction using aggregated features up to the

current time. Figure A.1 shows a screen shot of the visualizer. The waveform, spectro-

gram and feature representation are rendered as a short-term history and six different

categories of tags (emotion, vocal, genre, song, instrument and usage from CAL500)

are located as a cloud on an edge or a corner.

The primary purpose of this visualizer is to provide better insight on our proposed

data processing pipeline. In particular, we show learned feature representation as

“musical neuron activation” and tag prediction time- and size-varying. Not only that

92

APPENDIX A. REAL-TIME MUSIC TAGGING VISUALIZER 93

Figure A.1: A screen shot of music tagging visualizer

but also the visualizer can show subtle differences in prediction for individual songs,

for example, to what extent wrong predictions are dissimilar from the ground truth.

In addition, we can see how accurately the system makes predictions for “easy” or

“hard” examples. These are the merits of the visualizer that cannot be found in

simply showing average values of evaluation metrics.

A.2 Architecture

Figure A.2 shows the diagram of the software architecture. It is composed of three

main threads that communicate with each other via data buffers. Audio Manager

first reads an audio frame from a wave file and stores it into two separate buffers;

one is used for data processing and the other is for audio playback. This is associ-

ated with an audio callback function to read next frames. Processing Pipeline plays

a role of “musical sensory system.” It performs the sequence of computation in our

APPENDIX A. REAL-TIME MUSIC TAGGING VISUALIZER 94

audio bufferAudio

ManagerProcessing

Pipeline

audio buffer�����

PlayBack

Waveform buffer

View Manager

Mel-spec buffer

Hidden-Layer buffer

Input control(e.g. select song)

key command

Tag buffer

Figure A.2: Diagram of software architecture

proposed data processing pipeline; time-frequency AGC, FFT, mel-frequency map-

ping, feature encoding and tag prediction. We used sparse RBM and linear SVM

for learning algorithms (including PCA). The feature and tags were computed using

trained parameters from the algorithms. We produced four different outputs from

the pipeline. Each of them is stored in the corresponding buffer every frame time.

Finally, View Manager fetches the buffers and visualizes them at different locations.

A.3 Implementation Details

We built the software architecture based on the following software libraries.

• SFML (Simple and Fast Multimedia Library)1 : This is a free multimedia C++

API and used for 3-D visualization (OpenGL) in View Manager.

• RtAudio2 : This is an audio library for real-time audio input/output and used

for reading wav file and playback audio frames in Audio Manager.

1SFML: http://www.sfml-dev.org2RtAudio: http://www.music.mcgill.ca/~gary/rtaudio

APPENDIX A. REAL-TIME MUSIC TAGGING VISUALIZER 95

• Eigen3: This is a C++ template library for linear algebra. We used this for ma-

trix computation, element-wise math functions and FFT in Processing Pipeline.

• libsndfile4: This is a C library for reading and writing audio files. We used this

for reading .wav files.

3Eigen: http://eigen.tuxfamily.org/4libsndfile: http://www.mega-nerd.com/libsndfile/

Appendix B

Supplementary Materials

Audio and video examples to support this thesis are maintained in this website:

https://ccrma.stanford.edu/~juhan/thesis. It contains the following content.

• Dictionary learning animation

• Music tagging visualizer demo video

• Audio examples: reconstructed outputs from the following stages in the data

processing pipeline

- Original waveform

- AGC output

- Mel-frequency spectrogram

- PCA Whitening

- Sparse RBM

96

Bibliography

[1] Samer A. Abdallah and Mark D. Plumbley. Unsupervised analysis of

polyphonic music by sparse coding. IEEE Transactions on Neural Networks,

2006.

[2] Jean-Julien Aucouturier and Franois Pachet. Representing musical genre: A

state of the art. Journal of New Music Research, 2003.

[3] Horace B. Barlow. Possible principles underlying the transformation of

sensory messages. Sensory Communication, pages 217–234, 1961.

[4] Yoshua Bengio and Hugo Larochelle Pascal Lamblin, Dan Popovici. Greedy

layer-wise training of deep networks. Advances in Neural Information

Processing Systems 19, 2007.

[5] James Bergstra, Norman Casagrande, Dumitru Erhan, Douglas Eck, and

Balazs Kegl. Aggregate features and adaboost for music classification.

Machine Learning, 2006.

[6] James Bergstra, Michael Mandel, and Douglas Eck. Scalable genre and tag

prediction using spectral covariance. In Proceedings of the 11th International

Conference on Music Information Retrieval (ISMIR), 2010.

[7] T. Bertin-Mahieux, D. Eck, F. Maillet, and P. Lamere. Autotagger: a model

for predicting social tags from acoustic features on large music databases. In

Journal of New Music Research, 2008.

97

BIBLIOGRAPHY 98

[8] Thierry Bertin-Mahieux, Dan Ellis, Brian Whitman, and Paul Lamere. The

million song dataset. In Proceedings of the 12th international society for music

information retrieval conference, 2011.

[9] Thomas Blumensath and Mike Davies. Sparse and shift-invariant

representations of music. IEEE transactions on audio, speech and language

processing, 2006.

[10] Y-Lan Boureau, Jean Ponce, and Yann LeCun. A theoretical analysis of

feature pooling in visual recognition. Proceedings of the 27th International

Conference on Machine Learning (ICML), 2010.

[11] J. S. Bridle and M. D. Brown. An experimental automatic word-recognition

system. JSRU Report No. 1003, Joint Speech Research Unit, 1974.

[12] David .J. Field Bruno. A. Olshausen. Emergence of simple-cellreceptive field

properties by learning a sparse code for natural images. Nature, pages

607–609, 1996.

[13] Michael A. Casey, Remco VeltKamp, Masataka Goto, Marc Leman,

Christophe Rhodes, and Malcolm Slaney. Content-based music information

retrieval: Current directions and future challenges. Proceedings of the IEEE,

96:668–696, 2008.

[14] Oscar Celma. Music recommendation and discovery in the long tail. PhD

thesis, Universitat Pompeu Fabra,Barcelona, Spain, 2008.

[15] Adam Coates, Blake Carpenter, Carl Case, Sanjeev Satheesh, Bipin Suresh,

Tao Wang, David J. Wu, and Andrew Y. Ng. Text detection and character

recognition in scene images with unsupervised feature learning. 2011.

[16] Adam Coates, Honglak Lee, and Andrew Ng. An analysis of single-layer

networks in unsupervised feature learning. Journal of Machine Learning

Research, 2011.

BIBLIOGRAPHY 99

[17] Adam Coates and Andrew Y. Ng. The importance of encoding versus training

with sparse coding and vector quantization. In Proceedings of the 28th

International Conference on Machine Learning (ICML), 2011.

[18] Emanuele Coviello, Antoni B. Chan, and Gert R. G. Lanckriet. Time series

models for semantic music annotation. IEEE Transactions on Audio, Speech,

and Language Processing, 2011.

[19] Emanuele Coviello, Riccardo Miotto, and Gert R. G. Lanckriet. Combining

content-based auto-taggers with decision-fusion. In Proceedings of the 12th

International Conference on Music Information Retrieval (ISMIR), 2011.

[20] Steven B. Davis and Paul Mermelstein. Comparison of parametric

representations for monosyllabic word recognition in continuously spoken

sentences. IEEE Transactions on Acoustics, Speech and Signal Processing,

1980.

[21] Sander Dieleman, Philmon Brakel, and Benjamin Schrauwen. Audio-based

music classification with a pretrained convolutional network. In Proceedings of

the 12th International Conference on Music Information Retrieval (ISMIR),

2011.

[22] Dan Ellis. A history and overview of machine listening. web resource,

available,

http://www.ee.columbia.edu/~dpwe/talks/gatsby-2010-05.pdf, 2010.

[23] Dan Ellis. Time-frequency automatic gain control. web resource, available,

http://labrosa.ee.columbia.edu/matlab/tf_agc/, 2010.

[24] Dan Ellis and Graham Poliner. Identifying cover songs with chroma features

and dynamic programming beat tracking. In Proceedings of the IEEE

International Conference on Acoustics, Speech and Signal Processing

(ICASSP), Honolulu, Hawaii, USA, 2007.

BIBLIOGRAPHY 100

[25] Daniel P. W. Ellis. PLP and RASTA (and MFCC, and inversion) in Matlab,

2005. online web resource.

[26] Katherine Ellis, Emanuele Coviello, and Gert R. G. Lanckriet. Semantic

annotation and retrieval of music using a bag of systems representation. In

Proceedings of the 12th International Conference on Music Information

Retrieval (ISMIR), 2011.

[27] Valentin Emiya, Roland Badeau, and Bertrand David. Multipitch estimation

of piano sounds using a new probabilistic spectral smoothness principle. IEEE

Transaction on Audio, Speech and Language Processing, 2010.

[28] Rong-En Fan, Kai-Wei Chang, Cho-Jui Hseigh, Xiang-Rui Wang, and

Chih-Jen Lin. LIBLINEAR: a library for large linear classification. Journal of

Machine Learning Research, 2008.

[29] Zhouyu Fu, Guojun Lu, Kai Ming Ting, and Dengsheng Zhang. On feature

combination for music classification. Proceedings of International Workshop

on Statistical Pattern Recognition, 2010.

[30] Takuya Fujishima. Realtime chord recognition of musical sound: a system

using common lisp music. In Proceedings of the International Conference on

Computer Music (ICMC), 1999.

[31] Masataka Goto. A predominant-f0 estimation method for cd recordings:map

estimation using em algorithm for adaptive tone models. In Preceedings of

IEEE International Conference on Acoustics, Speech and Signal Processing,

2001.

[32] Philippe Hamel and Douglas Eck. Learning features from music audio with

deep belief networks. In In Proceedings of the 11th International Conference

on Music Information Retrieval (ISMIR), 2010.

[33] Philippe Hamel, Simon Lemieux, Yoshua Bengio, and Douglas Eck. Temporal

pooling and multiscale learning for automatic annotation and ranking of music

BIBLIOGRAPHY 101

audio. In Proceedings of the 12th International Conference on Music

Information Retrieval (ISMIR), 2011.

[34] Philippe Hamel, Sean Wood, and Douglas Eck. Automatic identification of

instrument classes in polyphonic and poly-instrument audio. In In Proceedings

of the 10th International Conference on Music Information Retrieval

(ISMIR), 2009.

[35] Mikael Henaff, Kevin Jarrett, Koray Kavukcuoglu, and Yann LeCun.

Unsupervised learning of sparse features for scalable audio classification. In

Proceedings of the 12th International Conference on Music Information

Retrieval (ISMIR), 2011.

[36] Geoffrey E. Hinton, Simon Osindero, and Yee-Whye Teh. A fast learning

algorithm for deep belief nets. Neural computation, 18:1527–1554, 2006.

[37] Geoffrey E. Hinton and Ruslan R. Salakhutdinov. Reducing the

dimensionality of data with neural networks. Science, 2006.

[38] Matt Hoffman, David Blei, and Perry Cook. Easy as CBA: A simple

probabilistic model for tagging music. In Proceedings of the 10th International

Conference on Music Information Retrieval (ISMIR), 2009.

[39] Andr Holzapfel and Yannis Stylianou. Musical genre classication using

nonnegative matrix factorization-based features. IEEE Transactions on

Acoustics, Speech and Signal Processing, 2008.

[40] Eric J. Humphrey, Juan Pablo Bello, and Yann LeCun. Moving beyond

feature design: Deep architectures and automatic feature learning in music

informatics. In Proceedings of the 13th International Conference on Music

Information Retrieval (ISMIR), 2012.

[41] M. J. Hunt, M. Lennig, and P. Mermelstein. Experiments in syllable-based

recognition of continuous speech. In ICASSP, 1980.

BIBLIOGRAPHY 102

[42] Aapo Hyvarinen, Jarmo Hurri, and Patrik O. Hoyer. Natural Image Statistics.

Springer-Verlag, 2009.

[43] Navdeep Jaitly and Geoffrey Hinton. Learning a better representation of

speech sound waves using restricted boltzmann machines. In Proceedings of

the IEEE International Conference on Acoustics, Speech and Signal Processing

(ICASSP), 2011.

[44] Navdeep Jaitly and Geoffrey E. Hinton. A new way to learn acoustic events.

NIPS 2011 Workshop on Deep Learning and Unsupervised Feature Learning,

2011.

[45] Kaichun K. Chang Jyh-Shing Roger Jang and Costas S. Iliopoulos. Music

genre classication via compressive sampling. In Proceedings of the 11th

International Conference on Music Information Retrieval (ISMIR), 2010.

[46] D.-N. Jiang, L. Lu, H.-J. Zhang, and J.-H. Tao. Music type classification by

spectral contrast feature. In Proceedings of International Conference on

Multimedia Expo (ICME), 2002.

[47] Koray Kavukcuoglu, Marc’Aurelio Ranzato, and Yann LeCun. Fast inference

in sparse coding algorithms with applications to object recognition. Technical

Report CBLL-TR-2008-12-01, Computational and Biological Learning Lab,

Courant Institute, NYU, 2008.

[48] Anssi Klapuri. A perceptually motivated multiple-f0 estimation method. In

Proceedings of IEEE Workshop on Applications of Signal Processing to Audio

and Acoustics, 2005.

[49] Edith Law and Luis Von Ahn. Input-agreement: a new mechanism for

collecting data using human computation games. In Proc. Intl. Conf. on

Human factors in computing systems, CHI. ACM, 2009.

[50] Yann LeCun, Leon Bottou, Yoshua Bengio, and Patrick Haffner.

Gradient-based learning applied to document recognition. Proceedings of the

IEEE, 1998.

BIBLIOGRAPHY 103

[51] Chang-Hsing Lee, Jau-Ling Shih, Kun-Ming Yu, and Hwai-San Lin.

Automatic music genre classification based on modulation spectral analysis of

spectral and cepstral features. IEEE Transaction on Multimedia, 2009.

[52] Honglak Lee. Unsupervised feature learning via sparse hierarchical

representations. Ph.D thesis, Stanford University, 2010.

[53] Honglak Lee, Chaitanya Ekanadham, and Andrew Y. Ng. Sparse deep belief

net model for visual area V2. In Advances in Neural Information Processing

Systems 20, pages 873–880. 2008.

[54] Honglak Lee, Roger Grosse, Rajesh Ranganath, and Andrew Y. Ng.

Convolutional deep belief networks for scalable unsupervised learning of

hierarchical representations. In Proceedings of the 26th International

Conference on Machine Learning, pages 609–616, 2009.

[55] Honglak Lee, Yan Largman, Peter Pham, and Andrew Y. Ng. Unsupervised

feature learning for audio classification using convolutional deep belief

networks. In Advances in Neural Information Processing Systems 22, pages

1096–1104. 2009.

[56] Jong-Hwan Lee, Ho-Young Jung, Te-Won Lee, and Soo-Young Lee. Speech

feature extraction using independent component analysis. In Proceedings of

the IEEE International Conference on Acoustics, Speech and Signal Processing

(ICASSP), 2000.

[57] Michael S. Lewicki. Efficient coding of natural sounds. Nature Neuroscience,

2002.

[58] Tao Li, Mitsunori Ogihara, and Qi Li. A comparative study of content-based

music genre classification. In Proceedings of the 26th international ACM

SIGIR conference on Research and development in informaion retrieval, 2003.

[59] T. Lidy, A. Rauber, A. Pertusa, and J. Inesta. Combining audio and symbolic

descriptors for music classification from audio,. In Music Information

Retrieval Information Exchange (MIREX), 2007.

BIBLIOGRAPHY 104

[60] Dick Lyon. Machine hearing: an emerging field. IEEE Signal Processing

Magazine, 2010.

[61] Richard Lyon. Filter cascades as analogs of the cochlea. Neuromorphic

systems engineering: neural networks in silicon, 1998.

[62] Micahel Mandel and Dan Ellis. Song-level features and svms for music

classification. In Proceedings of the 6th International Conference on Music

Information Retrieval (ISMIR), 2005.

[63] Pierre-Antoine Manzagol, Thierry Bertin-Mahieux, and Douglas Eck. On the

use of sparse time-relative auditory codes for music. In Proceedings of the 9th

International Conference on Music Information Retrieval (ISMIR), 2008.

[64] Matija Marolt. A connectionist approach to automatic transcription of

polyphonic piano music. IEEE Transactions on Multimedia, 2004.

[65] Cory McKay and Ichiro Fujinaga. Musical genre classification: Is it worth

pursuing and how can it be improved? In Proceedings of the 6th International

Conference on Music Information Retrieval (ISMIR), 2006.

[66] Martin F. McKinney and Jeroen Breebaart. Features for audio and music

classification. In Proceedings of the 4th International Conference on Music

Information Retrieval (ISMIR), 2003.

[67] Brian. C. J. Moore and Brian R. Glasberg. A revision of zwicker’s loudness

model. Acta Acustica, 1996.

[68] J. Andy Moorer. On the transcription of musical sound by computer.

Computer Music Journal, 1987.

[69] Meinard Muller, Dan Ellis, Anssi Klapuri, and Gal Richard. Signal processing

for music analysis. IEEE Journal on Selected Topics in Signal Processing,

2011.

BIBLIOGRAPHY 105

[70] Meinard Muller and Sebastian Ewert. Chroma Toolbox: MATLAB

implementations for extracting variants of chroma-based audio features. In

Proceedings of the 12th International Conference on Music Information

Retrieval (ISMIR), Miami, USA, 2011.

[71] Juhan Nam, Jorge Herrera, Malcolm Slaney, and Julius O. Smith. Learning

sparse feature representations for music annotation and retrieval. In

Proceedings of the 13th International Conference on Music Information

Retrieval (ISMIR), 2012.

[72] Juhan Nam, Jiquan Ngiam, Honglak Lee, and Malcolm Slaney. A

classification-based polyphonic piano transcription approach using learned

feature representation. In Proceedings of the 12th International Conference on

Music Information Retrieval (ISMIR), 2011.

[73] Andrew Ng. Unsupervised feature learning and deep learning. web resource,

available, http://icml2011speechvision.files.wordpress.com/2011/06/

visionaudio.pdf, 2011.

[74] Andrew Y. Ng. CS294A lecture note: Sparse autoencoder.

[75] Andrew Y. Ng and Michael I. Jordan. On discriminative vs. generative

classifiers: A comparison of logistic regression and naive bayes. Advances in

Neural Information Processing Systems 14, 2001.

[76] J. Ngiam, P. Koh, Z. Chen, S. Bhaskar, and A.Y. Ng. Sparse filtering.

Proceedings of the 25th Conference on Neural Information Processing Systems

(NIPS), 2011.

[77] Jiquan Ngiam, Aditya Khosla, Mingyu Kim, Juhan Nam, Honglak Lee, and

Andrew Y. Ng. Multimodal deep learning. In Proceedings of the 28th

International Conference on Machine Learning (ICML), June 2011.

[78] Bruno A. Olshausen and David J. Field. Sparse coding with an overcomplete

basis set: a strategy employed by v1. Vision Research, 37:3311–3325, 1997.

BIBLIOGRAPHY 106

[79] Elias Pampalk, Arthur Flexer, and Gerhard Widmer. Improvements of

audio-based music similarity and genre classification. In Proceedings of the 6th

International Conference on Music Information Retrieval (ISMIR), 2005.

[80] Elias Pampalk, Andreas Rauber, and Dieter Merkl. Content-based

organization and visualization of music archives. In Proceedings of ACM

Multimedia, 2002.

[81] Yannis Panagakis, Constantine Kotropoulos, and Gonzalo R. Arce. Music

genre classification using locality preserving nonnegative tensor factorization

and sparse representations. In Proceedings of the 10th International

Conference on Music Information Retrieval (ISMIR), 2009.

[82] Yannis Panagakis, Constantine Kotropoulos, and Gonzalo R. Arce. Music

genre classification via sparse representation of auditory temporal

modulations. In Proceedings of the 17th European Signal Processing

Conference (EUSIPCO), 2009.

[83] Yannis Panagakis, Constantine Kotropoulos, and Gonzalo R. Arce.

Non-negative multilinear principal component analysis of auditory temporal

modulations for music genre classification. IEEE Transaction on Audio,

Speech and Language Processing, 2010.

[84] Roy D. Patterson, Mike H. Allerhand, and Christian Giguere. Time-domain

modelling of peripheral auditory processing: A modular architecture and

software platform. Journal of the Acoustical Society of America, 1995.

[85] Graham E. Poliner and D. Ellis. A discriminative model for polyphonic piano

transcription. EURASIP Journal on Advances in Signal Processing, 2007.

[86] Graham E. Poliner and D. Ellis. Improving generalization for

classification-based polyphonic piano transcription. In Proceedings of IEEE

Workshop on Applications of Signal Processing to Audio and Acoustics, 2007.

BIBLIOGRAPHY 107

[87] Matti Ryynnen and Anssi Klapuri. Polyphonic music transcription using note

event modeling. In Proceedings of IEEE Workshop on Applications of Signal

Processing to Audio and Acoustics, 2005.

[88] Ruslan Salakhutdinov and Geoffrey Hinton. Semantic hashing. International

Journal of Approximate Reasoning, 2009.

[89] Andrew I. Schein, Alexandrin Popescul, Lyle H. Ungar, and David M.

Pennock. Methods and metrics for cold-start recommendations. In

Proceedings of the 25th annual international ACM SIGIR conference on

Research and development in information retrieval, SIGIR ’02, pages 253–260,

New York, NY, USA, 2002. ACM.

[90] Jan Schluter and Christian Osendorfer. Music Similarity Estimation with the

Mean-Covariance Restricted Boltzmann Machine. In Proceedings of the 10th

International Conference on Machine Learning and Applications, 2011.

[91] Erik M. Schmidt and Youngmoo E. Kim. Learning emotion-based acoustic

features with deep belief networks. In Proceedings of the 2011 IEEE Workshop

on Applications of Signal Processing to Audio and Acoustics (WASPAA),

2011.

[92] Erik M. Schmidt, Jeffrey Scott, and Youngmoo E. Kim. Feature learning in

dynamic environments:modeling the acoustic structure of musical emotion. In

Proceedings of the 13th International Conference on Music Information

Retrieval (ISMIR), 2012.

[93] Malcolm Slaney. Auditory toolbox-version 2, 1998. online web resource.

[94] Malcolm Slaney. Web-scale multimedia analysis: Does content matter? IEEE

Multimedia, 18, 2011.

[95] Paris Smaragdis and Judy C. Brown. Non-negative matrix factorization for

polyphonic music transcription. In Proceedings of IEEE Workshop on

Applications of Signal Processing to Audio and Acoustics, 2003.

BIBLIOGRAPHY 108

[96] Evan Smith and Michael S. Lewicki. Efficient auditory coding. Nature, 2006.

[97] Julius Smith. Spectral Audio Signal Processing. W3K Publishing, 2011.

[98] Paul Smolensky. Information processing in dynamical systems:foundation of

harmony theory. In Parallel Distributed Processing: explorations in the

microstructure of cognition, vol. 1, pages 194–281. MIT Press, Cambridge,

1986.

[99] Stanley Smith Stevens, John Volkman, and Edwin Newman. A scale for the

measurement of the psychological magnitude pitch. Journal of the Acoustical

Society of America, 1937.

[100] K. Trohidis, G. Tsoumakas, G. Kalliris, and I. Vlahavas. Multilabel

classification of music into emotions. In Proceedings of the 9th International

Conference on Music Information Retrieval (ISMIR), 2008.

[101] Douglas Turnbull, Luke Barrington, and Gert Lanckriet. Modeling music and

words using a multi-class naive bayes approach. In Proceedings of the 7th

International Conference on Music Information Retrieval (ISMIR), 2006.

[102] Douglas Turnbull, Luke Barrington, David Torres, and Gert Lanckriet.

Towards musical query-by-semantic description using the CAL500 data set. In

ACM Special Interest Group on Information Retrieval Conference, 2007.

[103] Douglas Turnbull, Luke Barrington, David Torres, and Gert R. G. Lanckriet.

Semantic annotation and retrieval of music and sound effects. IEEE

Transactions on Audio, Speech, and Language Processing, 2008.

[104] George Tzanetakis and Perry Cook. Musical genre classification of audio

signals. IEEE Transaction on Speech and Audio Processing, 2002.

[105] George Tzanetakis, Randy Jones, and Kirk McNally. Stereo panning features

for classifying recording production style. In Proceedings of the 8th

International Conference on Music Information Retrieval (ISMIR), 2007.

BIBLIOGRAPHY 109

[106] C. G. v. d. Boogaart and R. Lienhart. Note onset detection for the

transcription of polyphonic piano music. In Preceedings of IEEE International

Conference on Multimedia and Expo, 2009.

[107] Emmanuel Vincent, Nancy Bertin, and Roland Badeau. Adaptive harmonic

spectral decomposition for multiple pitch estimation. IEEE Transaction on

Audio, Speech and Language Processing, 2010.

[108] Gregory H. Wakefield. Mathematical representation of joint time-chroma

distributions. In SPIE, Denver, Colorado, 1999.

[109] Erling Wold, Thom Blum, Douglas Keislar, and James Wheaton.

Content-based classification, search, and retrieval of audio. IEEE MultiMedia,

3(3):27–36, 1996.

[110] Tong Tong Wu and Kenneth Lange. Coordinate descent algorithms for lasso

penalized regression. Annals of Applied Statistics, 2008.

[111] Jan Wulng and Martin Riedmiller. Unsupervised learning of local features for

music classification. In Proceedings of the 13th International Conference on

Music Information Retrieval (ISMIR), 2012.

[112] Bo Xie, Wei Bian, Dacheng Tao, and Parag Chordia. Music tagging with

regularized logistic regression. In Proceedings of the 12th International

Conference on Music Information Retrieval (ISMIR), 2011.

[113] Eberhard Zwicker. A scale for the measurement of the psychological

magnitude pitch. Journal of the Acoustical Society of America, 1961.