comparison writeup/aistats2011_submission.tex @ 600:1f5d2d01b84d

draft submission to AISTATS 2011
author Yoshua Bengio <bengioy@iro.umontreal.ca>
date Sun, 31 Oct 2010 09:11:47 -0400
parents
children 203c6071e104
comparison
equal deleted inserted replaced
598:a0fdc1f134da 600:1f5d2d01b84d
1 %\documentclass[twoside,11pt]{article} % For LaTeX2e
2 \documentclass{article} % For LaTeX2e
3 \usepackage{aistats2e_2011}
4 \usepackage{times}
5 \usepackage{wrapfig}
6 \usepackage{amsthm}
7 \usepackage{amsmath}
8 \usepackage{bbm}
9 \usepackage[utf8]{inputenc}
10 \usepackage[psamsfonts]{amssymb}
11 %\usepackage{algorithm,algorithmic} % not used after all
12 \usepackage{graphicx,subfigure}
13 \usepackage[numbers]{natbib}
14
15 \addtolength{\textwidth}{10mm}
16 \addtolength{\evensidemargin}{-5mm}
17 \addtolength{\oddsidemargin}{-5mm}
18
19 %\setlength\parindent{0mm}
20
21 \begin{document}
22
23 \title{Deeper Learners Benefit More from Multi-Task and Perturbed Examples}
24 \author{
25 Yoshua Bengio \and
26 Frédéric Bastien \and
27 Arnaud Bergeron \and
28 Nicolas Boulanger-Lewandowski \and
29 Thomas Breuel \and
30 Youssouf Chherawala \and
31 Moustapha Cisse \and
32 Myriam Côté \and
33 Dumitru Erhan \and
34 Jeremy Eustache \and
35 Xavier Glorot \and
36 Xavier Muller \and
37 Sylvain Pannetier Lebeuf \and
38 Razvan Pascanu \and
39 Salah Rifai \and
40 Francois Savard \and
41 Guillaume Sicard
42 }
43 \date{{\tt bengioy@iro.umontreal.ca}, Dept. IRO, U. Montreal, P.O. Box 6128, Centre-Ville branch, H3C 3J7, Montreal (Qc), Canada}
44 %\jmlrheading{}{2010}{}{10/2010}{XX/2011}{Yoshua Bengio et al}
45 %\editor{}
46
47 %\makeanontitle
48 \maketitle
49
50 %{\bf Running title: Deep Self-Taught Learning}
51
52 \vspace*{-2mm}
53 \begin{abstract}
54 Recent theoretical and empirical work in statistical machine learning has demonstrated the potential of learning algorithms for deep architectures, i.e., function classes obtained by composing multiple levels of representation. The hypothesis evaluated here is that intermediate levels of representation, because
55 they can be shared across tasks and examples from different but related
56 distributions, can yield even more benefits where there are more such levels of representation. The experiments are performed on a large-scale handwritten character recognition setting with 62 classes (upper case, lower case, digits). We show that a deep learner could not only {\em beat previously published results but also reach human-level performance}.
57 \end{abstract}
58 \vspace*{-3mm}
59
60 %\begin{keywords}
61 %Deep learning, self-taught learning, out-of-distribution examples, handwritten character recognition, multi-task learning
62 %\end{keywords}
63 %\keywords{self-taught learning \and multi-task learning \and out-of-distribution examples \and deep learning \and handwriting recognition}
64
65
66
67 \section{Introduction}
68 \vspace*{-1mm}
69
70 {\bf Deep Learning} has emerged as a promising new area of research in
71 statistical machine learning~\citep{Hinton06,ranzato-07-small,Bengio-nips-2006,VincentPLarochelleH2008,ranzato-08,TaylorHintonICML2009,Larochelle-jmlr-2009,Salakhutdinov+Hinton-2009,HonglakL2009,HonglakLNIPS2009,Jarrett-ICCV2009,Taylor-cvpr-2010}. See \citet{Bengio-2009} for a review.
72 Learning algorithms for deep architectures are centered on the learning
73 of useful representations of data, which are better suited to the task at hand,
74 and are organized in a hierarchy with multiple levels.
75 This is in part inspired by observations of the mammalian visual cortex,
76 which consists of a chain of processing elements, each of which is associated with a
77 different representation of the raw visual input. In fact,
78 it was found recently that the features learnt in deep architectures resemble
79 those observed in the first two of these stages (in areas V1 and V2
80 of visual cortex) \citep{HonglakL2008}, and that they become more and
81 more invariant to factors of variation (such as camera movement) in
82 higher layers~\citep{Goodfellow2009}.
83 Learning a hierarchy of features increases the
84 ease and practicality of developing representations that are at once
85 tailored to specific tasks, yet are able to borrow statistical strength
86 from other related tasks (e.g., modeling different kinds of objects). Finally, learning the
87 feature representation can lead to higher-level (more abstract, more
88 general) features that are more robust to unanticipated sources of
89 variance extant in real data.
90
91 Whereas a deep architecture can in principle be more powerful than a
92 shallow one in terms of representation, depth appears to render the
93 training problem more difficult in terms of optimization and local minima.
94 It is also only recently that successful algorithms were proposed to
95 overcome some of these difficulties. All are based on unsupervised
96 learning, often in an greedy layer-wise ``unsupervised pre-training''
97 stage~\citep{Bengio-2009}.
98 The principle is that each layer starting from
99 the bottom is trained to represent its input (the output of the previous
100 layer). After this
101 unsupervised initialization, the stack of layers can be
102 converted into a deep supervised feedforward neural network and fine-tuned by
103 stochastic gradient descent.
104 One of these layer initialization techniques,
105 applied here, is the Denoising
106 Auto-encoder~(DAE)~\citep{VincentPLarochelleH2008-very-small} (see
107 Figure~\ref{fig:da}), which performed similarly or
108 better~\citep{VincentPLarchelleH2008-very-small} than previously
109 proposed Restricted Boltzmann Machines (RBM)~\citep{Hinton06}
110 in terms of unsupervised extraction
111 of a hierarchy of features useful for classification. Each layer is trained
112 to denoise its input, creating a layer of features that can be used as
113 input for the next layer. Note that training a Denoising Auto-Encoder
114 can actually been seen as training a particular RBM by an inductive
115 principle different from maximum likelihood~\cite{Vincent-SM-2010}, namely by
116 Score Matching~\citep{Hyvarinen-2005,HyvarinenA2008}.
117
118 Previous comparative experimental results with stacking of RBMs and DAEs
119 to build deep supervised predictors had shown that they could outperform
120 shallow architectures in a variety of settings (see~\citet{Bengio-2009}
121 for a review), especially
122 when the data involves complex interactions between many factors of
123 variation~\citep{LarochelleH2007}. Other experiments have suggested
124 that the unsupervised layer-wise pre-training acted as a useful
125 prior~\citep{Erhan+al-2010} that allows one to initialize a deep
126 neural network in a relatively much smaller region of parameter space,
127 corresponding to better generalization.
128
129 To further the understanding of the reasons for the good performance
130 observed with deep learners, we focus here on the following {\em hypothesis}:
131 intermediate levels of representation, especially when there are
132 more such levels, can be exploited to {\bf share
133 statistical strength across different but related types of examples},
134 such as examples coming from other tasks than the task of interest
135 (the multi-task setting), or examples coming from an overlapping
136 but different distribution (images with different kinds of perturbations
137 and noises, here). This is consistent with the hypotheses discussed
138 at length in~\citet{Bengio-2009} regarding the potential advantage
139 of deep learning and the idea that more levels of representation can
140 give rise to more abstract, more general features of the raw input.
141
142 This hypothesis is related to a learning setting called
143 {\bf self-taught learning}~\citep{RainaR2007}, which combines principles
144 of semi-supervised and multi-task learning: the learner can exploit examples
145 that are unlabeled and possibly come from a distribution different from the target
146 distribution, e.g., from other classes than those of interest.
147 It has already been shown that deep learners can clearly take advantage of
148 unsupervised learning and unlabeled examples~\citep{Bengio-2009,WestonJ2008-small},
149 but more needed to be done to explore the impact
150 of {\em out-of-distribution} examples and of the {\em multi-task} setting
151 (one exception is~\citep{CollobertR2008}, which shares and uses unsupervised
152 pre-training only with the first layer). In particular the {\em relative
153 advantage of deep learning} for these settings has not been evaluated.
154
155
156 %
157 The {\bf main claim} of this paper is that deep learners (with several levels of representation) can
158 {\bf benefit more from self-taught learning than shallow learners} (with a single
159 level), both in the context of the multi-task setting and from {\em
160 out-of-distribution examples} in general. Because we are able to improve on state-of-the-art
161 performance and reach human-level performance
162 on a large-scale task, we consider that this paper is also a contribution
163 to advance the application of machine learning to handwritten character recognition.
164 More precisely, we ask and answer the following questions:
165
166 %\begin{enumerate}
167 $\bullet$ %\item
168 Do the good results previously obtained with deep architectures on the
169 MNIST digit images generalize to the setting of a similar but much larger and richer
170 dataset, the NIST special database 19, with 62 classes and around 800k examples?
171
172 $\bullet$ %\item
173 To what extent does the perturbation of input images (e.g. adding
174 noise, affine transformations, background images) make the resulting
175 classifiers better not only on similarly perturbed images but also on
176 the {\em original clean examples}? We study this question in the
177 context of the 62-class and 10-class tasks of the NIST special database 19.
178
179 $\bullet$ %\item
180 Do deep architectures {\em benefit {\bf more} from such out-of-distribution}
181 examples, in particular do they benefit more from
182 examples that are perturbed versions of the examples from the task of interest?
183
184 $\bullet$ %\item
185 Similarly, does the feature learning step in deep learning algorithms benefit {\bf more}
186 from training with moderately {\em different classes} (i.e. a multi-task learning scenario) than
187 a corresponding shallow and purely supervised architecture?
188 We train on 62 classes and test on 10 (digits) or 26 (upper case or lower case)
189 to answer this question.
190 %\end{enumerate}
191
192 Our experimental results provide positive evidence towards all of these questions,
193 as well as {\em classifiers that reach human-level performance on 62-class isolated character
194 recognition and beat previously published results on the NIST dataset (special database 19)}.
195 To achieve these results, we introduce in the next section a sophisticated system
196 for stochastically transforming character images and then explain the methodology,
197 which is based on training with or without these transformed images and testing on
198 clean ones. We measure the relative advantage of out-of-distribution examples
199 (perturbed or out-of-class)
200 for a deep learner vs a supervised shallow one.
201 Code for generating these transformations as well as for the deep learning
202 algorithms are made available at {\tt http://anonymous.url.net}.%{\tt http://hg.assembla.com/ift6266}.
203 We also estimate the relative advantage for deep learners of training with
204 other classes than those of interest, by comparing learners trained with
205 62 classes with learners trained with only a subset (on which they
206 are then tested).
207 The conclusion discusses
208 the more general question of why deep learners may benefit so much from
209 the self-taught learning framework. Since out-of-distribution data
210 (perturbed or from other related classes) is very common, this conclusion
211 is of practical importance.
212
213 \vspace*{-3mm}
214 %\newpage
215 \section{Perturbed and Transformed Character Images}
216 \label{s:perturbations}
217 \vspace*{-2mm}
218
219 Figure~\ref{fig:transform} shows the different transformations we used to stochastically
220 transform $32 \times 32$ source images (such as the one in Fig.\ref{fig:torig})
221 in order to obtain data from a larger distribution which
222 covers a domain substantially larger than the clean characters distribution from
223 which we start.
224 Although character transformations have been used before to
225 improve character recognizers, this effort is on a large scale both
226 in number of classes and in the complexity of the transformations, hence
227 in the complexity of the learning task.
228 The code for these transformations (mostly python) is available at
229 {\tt http://anonymous.url.net}. All the modules in the pipeline share
230 a global control parameter ($0 \le complexity \le 1$) that allows one to modulate the
231 amount of deformation or noise introduced.
232 There are two main parts in the pipeline. The first one,
233 from slant to pinch below, performs transformations. The second
234 part, from blur to contrast, adds different kinds of noise.
235 More details can be found in~\citep{ift6266-tr-anonymous}.
236
237 \begin{figure}[ht]
238 \centering
239 \subfigure[Original]{\includegraphics[scale=0.6]{images/Original.png}\label{fig:torig}}
240 \subfigure[Thickness]{\includegraphics[scale=0.6]{images/Thick_only.png}}
241 \subfigure[Slant]{\includegraphics[scale=0.6]{images/Slant_only.png}}
242 \subfigure[Affine Transformation]{\includegraphics[scale=0.6]{images/Affine_only.png}}
243 \subfigure[Local Elastic Deformation]{\includegraphics[scale=0.6]{images/Localelasticdistorsions_only.png}}
244 \subfigure[Pinch]{\includegraphics[scale=0.6]{images/Pinch_only.png}}
245 %Noise
246 \subfigure[Motion Blur]{\includegraphics[scale=0.6]{images/Motionblur_only.png}}
247 \subfigure[Occlusion]{\includegraphics[scale=0.6]{images/occlusion_only.png}}
248 \subfigure[Gaussian Smoothing]{\includegraphics[scale=0.6]{images/Bruitgauss_only.png}}
249 \subfigure[Pixels Permutation]{\includegraphics[scale=0.6]{images/Permutpixel_only.png}}
250 \subfigure[Gaussian Noise]{\includegraphics[scale=0.6]{images/Distorsiongauss_only.png}}
251 \subfigure[Background Image Addition]{\includegraphics[scale=0.6]{images/background_other_only.png}}
252 \subfigure[Salt \& Pepper]{\includegraphics[scale=0.6]{images/Poivresel_only.png}}
253 \subfigure[Scratches]{\includegraphics[scale=0.6]{images/Rature_only.png}}
254 \subfigure[Grey Level \& Contrast]{\includegraphics[scale=0.6]{images/Contrast_only.png}}
255 \caption{Top left (a): example original image. Others (b-o): examples of the effect
256 of each transformation module taken separately. Actual perturbed examples are obtained by
257 a pipeline of these, with random choices about which module to apply and how much perturbation
258 to apply.}
259 \label{fig:transform}
260 \vspace*{-2mm}
261 \end{figure}
262
263 \vspace*{-3mm}
264 \section{Experimental Setup}
265 \vspace*{-1mm}
266
267 Much previous work on deep learning had been performed on
268 the MNIST digits task~\citep{Hinton06,ranzato-07-small,Bengio-nips-2006,Salakhutdinov+Hinton-2009},
269 with 60~000 examples, and variants involving 10~000
270 examples~\citep{Larochelle-jmlr-toappear-2008,VincentPLarochelleH2008}.
271 The focus here is on much larger training sets, from 10 times to
272 to 1000 times larger, and 62 classes.
273
274 The first step in constructing the larger datasets (called NISTP and P07) is to sample from
275 a {\em data source}: {\bf NIST} (NIST database 19), {\bf Fonts}, {\bf Captchas},
276 and {\bf OCR data} (scanned machine printed characters). Once a character
277 is sampled from one of these sources (chosen randomly), the second step is to
278 apply a pipeline of transformations and/or noise processes outlined in section \ref{s:perturbations}.
279
280 To provide a baseline of error rate comparison we also estimate human performance
281 on both the 62-class task and the 10-class digits task.
282 We compare the best Multi-Layer Perceptrons (MLP) against
283 the best Stacked Denoising Auto-encoders (SDA), when
284 both models' hyper-parameters are selected to minimize the validation set error.
285 We also provide a comparison against a precise estimate
286 of human performance obtained via Amazon's Mechanical Turk (AMT)
287 service ({\tt http://mturk.com}).
288 AMT users are paid small amounts
289 of money to perform tasks for which human intelligence is required.
290 Mechanical Turk has been used extensively in natural language processing and vision.
291 %processing \citep{SnowEtAl2008} and vision
292 %\citep{SorokinAndForsyth2008,whitehill09}.
293 AMT users were presented
294 with 10 character images (from a test set) and asked to choose 10 corresponding ASCII
295 characters. They were forced to choose a single character class (either among the
296 62 or 10 character classes) for each image.
297 80 subjects classified 2500 images per (dataset,task) pair.
298 Different humans labelers sometimes provided a different label for the same
299 example, and we were able to estimate the error variance due to this effect
300 because each image was classified by 3 different persons.
301 The average error of humans on the 62-class task NIST test set
302 is 18.2\%, with a standard error of 0.1\%.
303
304 \vspace*{-3mm}
305 \subsection{Data Sources}
306 \vspace*{-2mm}
307
308 %\begin{itemize}
309 %\item
310 {\bf NIST.}
311 Our main source of characters is the NIST Special Database 19~\citep{Grother-1995},
312 widely used for training and testing character
313 recognition systems~\citep{Granger+al-2007,Cortes+al-2000,Oliveira+al-2002-short,Milgram+al-2005}.
314 The dataset is composed of 814255 digits and characters (upper and lower cases), with hand checked classifications,
315 extracted from handwritten sample forms of 3600 writers. The characters are labelled by one of the 62 classes
316 corresponding to ``0''-``9'',``A''-``Z'' and ``a''-``z''. The dataset contains 8 parts (partitions) of varying complexity.
317 The fourth partition (called $hsf_4$, 82587 examples),
318 experimentally recognized to be the most difficult one, is the one recommended
319 by NIST as a testing set and is used in our work as well as some previous work~\citep{Granger+al-2007,Cortes+al-2000,Oliveira+al-2002-short,Milgram+al-2005}
320 for that purpose. We randomly split the remainder (731668 examples) into a training set and a validation set for
321 model selection.
322 The performances reported by previous work on that dataset mostly use only the digits.
323 Here we use all the classes both in the training and testing phase. This is especially
324 useful to estimate the effect of a multi-task setting.
325 The distribution of the classes in the NIST training and test sets differs
326 substantially, with relatively many more digits in the test set, and a more uniform distribution
327 of letters in the test set (whereas in the training set they are distributed
328 more like in natural text).
329 %\vspace*{-1mm}
330
331 %\item
332 {\bf Fonts.}
333 In order to have a good variety of sources we downloaded an important number of free fonts from:
334 {\tt http://cg.scs.carleton.ca/\textasciitilde luc/freefonts.html}.
335 % TODO: pointless to anonymize, it's not pointing to our work
336 Including the operating system's (Windows 7) fonts, there is a total of $9817$ different fonts that we can choose uniformly from.
337 The chosen {\tt ttf} file is either used as input of the Captcha generator (see next item) or, by producing a corresponding image,
338 directly as input to our models.
339 \vspace*{-1mm}
340
341 %\item
342 {\bf Captchas.}
343 The Captcha data source is an adaptation of the \emph{pycaptcha} library (a python based captcha generator library) for
344 generating characters of the same format as the NIST dataset. This software is based on
345 a random character class generator and various kinds of transformations similar to those described in the previous sections.
346 In order to increase the variability of the data generated, many different fonts are used for generating the characters.
347 Transformations (slant, distortions, rotation, translation) are applied to each randomly generated character with a complexity
348 depending on the value of the complexity parameter provided by the user of the data source.
349 %Two levels of complexity are allowed and can be controlled via an easy to use facade class. %TODO: what's a facade class?
350 \vspace*{-1mm}
351
352 %\item
353 {\bf OCR data.}
354 A large set (2 million) of scanned, OCRed and manually verified machine-printed
355 characters where included as an
356 additional source. This set is part of a larger corpus being collected by the Image Understanding
357 Pattern Recognition Research group led by Thomas Breuel at University of Kaiserslautern
358 ({\tt http://www.iupr.com}), and which will be publicly released.
359 %TODO: let's hope that Thomas is not a reviewer! :) Seriously though, maybe we should anonymize this
360 %\end{itemize}
361
362 \vspace*{-3mm}
363 \subsection{Data Sets}
364 \vspace*{-2mm}
365
366 All data sets contain 32$\times$32 grey-level images (values in $[0,1]$) associated with a label
367 from one of the 62 character classes.
368 %\begin{itemize}
369 \vspace*{-1mm}
370
371 %\item
372 {\bf NIST.} This is the raw NIST special database 19~\citep{Grother-1995}. It has
373 \{651668 / 80000 / 82587\} \{training / validation / test\} examples.
374 \vspace*{-1mm}
375
376 %\item
377 {\bf P07.} This dataset is obtained by taking raw characters from all four of the above sources
378 and sending them through the transformation pipeline described in section \ref{s:perturbations}.
379 For each new example to generate, a data source is selected with probability $10\%$ from the fonts,
380 $25\%$ from the captchas, $25\%$ from the OCR data and $40\%$ from NIST. We apply all the transformations in the
381 order given above, and for each of them we sample uniformly a \emph{complexity} in the range $[0,0.7]$.
382 It has \{81920000 / 80000 / 20000\} \{training / validation / test\} examples.
383 \vspace*{-1mm}
384
385 %\item
386 {\bf NISTP.} This one is equivalent to P07 (complexity parameter of $0.7$ with the same proportions of data sources)
387 except that we only apply
388 transformations from slant to pinch. Therefore, the character is
389 transformed but no additional noise is added to the image, giving images
390 closer to the NIST dataset.
391 It has \{81920000 / 80000 / 20000\} \{training / validation / test\} examples.
392 %\end{itemize}
393
394 \vspace*{-3mm}
395 \subsection{Models and their Hyperparameters}
396 \vspace*{-2mm}
397
398 The experiments are performed using MLPs (with a single
399 hidden layer) and SDAs.
400 \emph{Hyper-parameters are selected based on the {\bf NISTP} validation set error.}
401
402 {\bf Multi-Layer Perceptrons (MLP).}
403 Whereas previous work had compared deep architectures to both shallow MLPs and
404 SVMs, we only compared to MLPs here because of the very large datasets used
405 (making the use of SVMs computationally challenging because of their quadratic
406 scaling behavior). Preliminary experiments on training SVMs (libSVM) with subsets of the training
407 set allowing the program to fit in memory yielded substantially worse results
408 than those obtained with MLPs. For training on nearly a billion examples
409 (with the perturbed data), the MLPs and SDA are much more convenient than
410 classifiers based on kernel methods.
411 The MLP has a single hidden layer with $\tanh$ activation functions, and softmax (normalized
412 exponentials) on the output layer for estimating $P(class | image)$.
413 The number of hidden units is taken in $\{300,500,800,1000,1500\}$.
414 Training examples are presented in minibatches of size 20. A constant learning
415 rate was chosen among $\{0.001, 0.01, 0.025, 0.075, 0.1, 0.5\}$.
416 %through preliminary experiments (measuring performance on a validation set),
417 %and $0.1$ (which was found to work best) was then selected for optimizing on
418 %the whole training sets.
419 \vspace*{-1mm}
420
421
422 {\bf Stacked Denoising Auto-Encoders (SDA).}
423 Various auto-encoder variants and Restricted Boltzmann Machines (RBMs)
424 can be used to initialize the weights of each layer of a deep MLP (with many hidden
425 layers)~\citep{Hinton06,ranzato-07-small,Bengio-nips-2006},
426 apparently setting parameters in the
427 basin of attraction of supervised gradient descent yielding better
428 generalization~\citep{Erhan+al-2010}. This initial {\em unsupervised
429 pre-training phase} uses all of the training images but not the training labels.
430 Each layer is trained in turn to produce a new representation of its input
431 (starting from the raw pixels).
432 It is hypothesized that the
433 advantage brought by this procedure stems from a better prior,
434 on the one hand taking advantage of the link between the input
435 distribution $P(x)$ and the conditional distribution of interest
436 $P(y|x)$ (like in semi-supervised learning), and on the other hand
437 taking advantage of the expressive power and bias implicit in the
438 deep architecture (whereby complex concepts are expressed as
439 compositions of simpler ones through a deep hierarchy).
440
441 \begin{figure}[ht]
442 \vspace*{-2mm}
443 \centerline{\resizebox{0.8\textwidth}{!}{\includegraphics{images/denoising_autoencoder_small.pdf}}}
444 \vspace*{-2mm}
445 \caption{Illustration of the computations and training criterion for the denoising
446 auto-encoder used to pre-train each layer of the deep architecture. Input $x$ of
447 the layer (i.e. raw input or output of previous layer)
448 s corrupted into $\tilde{x}$ and encoded into code $y$ by the encoder $f_\theta(\cdot)$.
449 The decoder $g_{\theta'}(\cdot)$ maps $y$ to reconstruction $z$, which
450 is compared to the uncorrupted input $x$ through the loss function
451 $L_H(x,z)$, whose expected value is approximately minimized during training
452 by tuning $\theta$ and $\theta'$.}
453 \label{fig:da}
454 \vspace*{-2mm}
455 \end{figure}
456
457 Here we chose to use the Denoising
458 Auto-encoder~\citep{VincentPLarochelleH2008} as the building block for
459 these deep hierarchies of features, as it is simple to train and
460 explain (see Figure~\ref{fig:da}, as well as
461 tutorial and code there: {\tt http://deeplearning.net/tutorial}),
462 provides efficient inference, and yielded results
463 comparable or better than RBMs in series of experiments
464 \citep{VincentPLarochelleH2008}. During training, a Denoising
465 Auto-encoder is presented with a stochastically corrupted version
466 of the input and trained to reconstruct the uncorrupted input,
467 forcing the hidden units to represent the leading regularities in
468 the data. Here we use the random binary masking corruption
469 (which sets to 0 a random subset of the inputs).
470 Once it is trained, in a purely unsupervised way,
471 its hidden units' activations can
472 be used as inputs for training a second one, etc.
473 After this unsupervised pre-training stage, the parameters
474 are used to initialize a deep MLP, which is fine-tuned by
475 the same standard procedure used to train them (see previous section).
476 The SDA hyper-parameters are the same as for the MLP, with the addition of the
477 amount of corruption noise (we used the masking noise process, whereby a
478 fixed proportion of the input values, randomly selected, are zeroed), and a
479 separate learning rate for the unsupervised pre-training stage (selected
480 from the same above set). The fraction of inputs corrupted was selected
481 among $\{10\%, 20\%, 50\%\}$. Another hyper-parameter is the number
482 of hidden layers but it was fixed to 3 based on previous work with
483 SDAs on MNIST~\citep{VincentPLarochelleH2008}. The size of the hidden
484 layers was kept constant across hidden layers, and the best results
485 were obtained with the largest values that we could experiment
486 with given our patience, with 1000 hidden units.
487
488 \vspace*{-1mm}
489
490 \begin{figure}[ht]
491 %\vspace*{-2mm}
492 \centerline{\resizebox{.99\textwidth}{!}{\includegraphics{images/error_rates_charts.pdf}}}
493 %\vspace*{-3mm}
494 \caption{SDAx are the {\bf deep} models. Error bars indicate a 95\% confidence interval. 0 indicates that the model was trained
495 on NIST, 1 on NISTP, and 2 on P07. Left: overall results
496 of all models, on NIST and NISTP test sets.
497 Right: error rates on NIST test digits only, along with the previous results from
498 literature~\citep{Granger+al-2007,Cortes+al-2000,Oliveira+al-2002-short,Milgram+al-2005}
499 respectively based on ART, nearest neighbors, MLPs, and SVMs.}
500 \label{fig:error-rates-charts}
501 \vspace*{-2mm}
502 \end{figure}
503
504
505 \begin{figure}[ht]
506 \vspace*{-3mm}
507 \centerline{\resizebox{.99\textwidth}{!}{\includegraphics{images/improvements_charts.pdf}}}
508 \vspace*{-3mm}
509 \caption{Relative improvement in error rate due to self-taught learning.
510 Left: Improvement (or loss, when negative)
511 induced by out-of-distribution examples (perturbed data).
512 Right: Improvement (or loss, when negative) induced by multi-task
513 learning (training on all classes and testing only on either digits,
514 upper case, or lower-case). The deep learner (SDA) benefits more from
515 both self-taught learning scenarios, compared to the shallow MLP.}
516 \label{fig:improvements-charts}
517 \vspace*{-2mm}
518 \end{figure}
519
520 \section{Experimental Results}
521 \vspace*{-2mm}
522
523 %%\vspace*{-1mm}
524 %\subsection{SDA vs MLP vs Humans}
525 %%\vspace*{-1mm}
526 The models are either trained on NIST (MLP0 and SDA0),
527 NISTP (MLP1 and SDA1), or P07 (MLP2 and SDA2), and tested
528 on either NIST, NISTP or P07, either on the 62-class task
529 or on the 10-digits task. Training (including about half
530 for unsupervised pre-training, for DAs) on the larger
531 datasets takes around one day on a GPU-285.
532 Figure~\ref{fig:error-rates-charts} summarizes the results obtained,
533 comparing humans, the three MLPs (MLP0, MLP1, MLP2) and the three SDAs (SDA0, SDA1,
534 SDA2), along with the previous results on the digits NIST special database
535 19 test set from the literature, respectively based on ARTMAP neural
536 networks ~\citep{Granger+al-2007}, fast nearest-neighbor search
537 ~\citep{Cortes+al-2000}, MLPs ~\citep{Oliveira+al-2002-short}, and SVMs
538 ~\citep{Milgram+al-2005}.% More detailed and complete numerical results
539 %(figures and tables, including standard errors on the error rates) can be
540 %found in Appendix.
541 The deep learner not only outperformed the shallow ones and
542 previously published performance (in a statistically and qualitatively
543 significant way) but when trained with perturbed data
544 reaches human performance on both the 62-class task
545 and the 10-class (digits) task.
546 17\% error (SDA1) or 18\% error (humans) may seem large but a large
547 majority of the errors from humans and from SDA1 are from out-of-context
548 confusions (e.g. a vertical bar can be a ``1'', an ``l'' or an ``L'', and a
549 ``c'' and a ``C'' are often indistinguishible).
550
551 In addition, as shown in the left of
552 Figure~\ref{fig:improvements-charts}, the relative improvement in error
553 rate brought by self-taught learning is greater for the SDA, and these
554 differences with the MLP are statistically and qualitatively
555 significant.
556 The left side of the figure shows the improvement to the clean
557 NIST test set error brought by the use of out-of-distribution examples
558 (i.e. the perturbed examples examples from NISTP or P07).
559 Relative percent change is measured by taking
560 $100 \% \times$ (original model's error / perturbed-data model's error - 1).
561 The right side of
562 Figure~\ref{fig:improvements-charts} shows the relative improvement
563 brought by the use of a multi-task setting, in which the same model is
564 trained for more classes than the target classes of interest (i.e. training
565 with all 62 classes when the target classes are respectively the digits,
566 lower-case, or upper-case characters). Again, whereas the gain from the
567 multi-task setting is marginal or negative for the MLP, it is substantial
568 for the SDA. Note that to simplify these multi-task experiments, only the original
569 NIST dataset is used. For example, the MLP-digits bar shows the relative
570 percent improvement in MLP error rate on the NIST digits test set
571 is $100\% \times$ (single-task
572 model's error / multi-task model's error - 1). The single-task model is
573 trained with only 10 outputs (one per digit), seeing only digit examples,
574 whereas the multi-task model is trained with 62 outputs, with all 62
575 character classes as examples. Hence the hidden units are shared across
576 all tasks. For the multi-task model, the digit error rate is measured by
577 comparing the correct digit class with the output class associated with the
578 maximum conditional probability among only the digit classes outputs. The
579 setting is similar for the other two target classes (lower case characters
580 and upper case characters).
581 %%\vspace*{-1mm}
582 %\subsection{Perturbed Training Data More Helpful for SDA}
583 %%\vspace*{-1mm}
584
585 %%\vspace*{-1mm}
586 %\subsection{Multi-Task Learning Effects}
587 %%\vspace*{-1mm}
588
589 \iffalse
590 As previously seen, the SDA is better able to benefit from the
591 transformations applied to the data than the MLP. In this experiment we
592 define three tasks: recognizing digits (knowing that the input is a digit),
593 recognizing upper case characters (knowing that the input is one), and
594 recognizing lower case characters (knowing that the input is one). We
595 consider the digit classification task as the target task and we want to
596 evaluate whether training with the other tasks can help or hurt, and
597 whether the effect is different for MLPs versus SDAs. The goal is to find
598 out if deep learning can benefit more (or less) from multiple related tasks
599 (i.e. the multi-task setting) compared to a corresponding purely supervised
600 shallow learner.
601
602 We use a single hidden layer MLP with 1000 hidden units, and a SDA
603 with 3 hidden layers (1000 hidden units per layer), pre-trained and
604 fine-tuned on NIST.
605
606 Our results show that the MLP benefits marginally from the multi-task setting
607 in the case of digits (5\% relative improvement) but is actually hurt in the case
608 of characters (respectively 3\% and 4\% worse for lower and upper class characters).
609 On the other hand the SDA benefited from the multi-task setting, with relative
610 error rate improvements of 27\%, 15\% and 13\% respectively for digits,
611 lower and upper case characters, as shown in Table~\ref{tab:multi-task}.
612 \fi
613
614
615 \vspace*{-2mm}
616 \section{Conclusions and Discussion}
617 \vspace*{-2mm}
618
619 We have found that the self-taught learning framework is more beneficial
620 to a deep learner than to a traditional shallow and purely
621 supervised learner. More precisely,
622 the answers are positive for all the questions asked in the introduction.
623 %\begin{itemize}
624
625 $\bullet$ %\item
626 {\bf Do the good results previously obtained with deep architectures on the
627 MNIST digits generalize to a much larger and richer (but similar)
628 dataset, the NIST special database 19, with 62 classes and around 800k examples}?
629 Yes, the SDA {\em systematically outperformed the MLP and all the previously
630 published results on this dataset} (the ones that we are aware of), {\em in fact reaching human-level
631 performance} at around 17\% error on the 62-class task and 1.4\% on the digits,
632 and beating previously published results on the same data.
633
634 $\bullet$ %\item
635 {\bf To what extent do self-taught learning scenarios help deep learners,
636 and do they help them more than shallow supervised ones}?
637 We found that distorted training examples not only made the resulting
638 classifier better on similarly perturbed images but also on
639 the {\em original clean examples}, and more importantly and more novel,
640 that deep architectures benefit more from such {\em out-of-distribution}
641 examples. MLPs were helped by perturbed training examples when tested on perturbed input
642 images (65\% relative improvement on NISTP)
643 but only marginally helped (5\% relative improvement on all classes)
644 or even hurt (10\% relative loss on digits)
645 with respect to clean examples . On the other hand, the deep SDAs
646 were significantly boosted by these out-of-distribution examples.
647 Similarly, whereas the improvement due to the multi-task setting was marginal or
648 negative for the MLP (from +5.6\% to -3.6\% relative change),
649 it was quite significant for the SDA (from +13\% to +27\% relative change),
650 which may be explained by the arguments below.
651 %\end{itemize}
652
653 In the original self-taught learning framework~\citep{RainaR2007}, the
654 out-of-sample examples were used as a source of unsupervised data, and
655 experiments showed its positive effects in a \emph{limited labeled data}
656 scenario. However, many of the results by \citet{RainaR2007} (who used a
657 shallow, sparse coding approach) suggest that the {\em relative gain of self-taught
658 learning vs ordinary supervised learning} diminishes as the number of labeled examples increases.
659 We note instead that, for deep
660 architectures, our experiments show that such a positive effect is accomplished
661 even in a scenario with a \emph{large number of labeled examples},
662 i.e., here, the relative gain of self-taught learning is probably preserved
663 in the asymptotic regime.
664
665 {\bf Why would deep learners benefit more from the self-taught learning framework}?
666 The key idea is that the lower layers of the predictor compute a hierarchy
667 of features that can be shared across tasks or across variants of the
668 input distribution. A theoretical analysis of generalization improvements
669 due to sharing of intermediate features across tasks already points
670 towards that explanation~\cite{baxter95a}.
671 Intermediate features that can be used in different
672 contexts can be estimated in a way that allows to share statistical
673 strength. Features extracted through many levels are more likely to
674 be more abstract and more invariant to some of the factors of variation
675 in the underlying distribution (as the experiments in~\citet{Goodfellow2009} suggest),
676 increasing the likelihood that they would be useful for a larger array
677 of tasks and input conditions.
678 Therefore, we hypothesize that both depth and unsupervised
679 pre-training play a part in explaining the advantages observed here, and future
680 experiments could attempt at teasing apart these factors.
681 And why would deep learners benefit from the self-taught learning
682 scenarios even when the number of labeled examples is very large?
683 We hypothesize that this is related to the hypotheses studied
684 in~\citet{Erhan+al-2010}. In~\citet{Erhan+al-2010}
685 it was found that online learning on a huge dataset did not make the
686 advantage of the deep learning bias vanish, and a similar phenomenon
687 may be happening here. We hypothesize that unsupervised pre-training
688 of a deep hierarchy with self-taught learning initializes the
689 model in the basin of attraction of supervised gradient descent
690 that corresponds to better generalization. Furthermore, such good
691 basins of attraction are not discovered by pure supervised learning
692 (with or without self-taught settings) from random initialization, and more labeled examples
693 does not allow the shallow or purely supervised models to discover
694 the kind of better basins associated
695 with deep learning and self-taught learning.
696
697 A Flash demo of the recognizer (where both the MLP and the SDA can be compared)
698 can be executed on-line at {\tt http://deep.host22.com}.
699
700 \iffalse
701 \section*{Appendix I: Detailed Numerical Results}
702
703 These tables correspond to Figures 2 and 3 and contain the raw error rates for each model and dataset considered.
704 They also contain additional data such as test errors on P07 and standard errors.
705
706 \begin{table}[ht]
707 \caption{Overall comparison of error rates ($\pm$ std.err.) on 62 character classes (10 digits +
708 26 lower + 26 upper), except for last columns -- digits only, between deep architecture with pre-training
709 (SDA=Stacked Denoising Autoencoder) and ordinary shallow architecture
710 (MLP=Multi-Layer Perceptron). The models shown are all trained using perturbed data (NISTP or P07)
711 and using a validation set to select hyper-parameters and other training choices.
712 \{SDA,MLP\}0 are trained on NIST,
713 \{SDA,MLP\}1 are trained on NISTP, and \{SDA,MLP\}2 are trained on P07.
714 The human error rate on digits is a lower bound because it does not count digits that were
715 recognized as letters. For comparison, the results found in the literature
716 on NIST digits classification using the same test set are included.}
717 \label{tab:sda-vs-mlp-vs-humans}
718 \begin{center}
719 \begin{tabular}{|l|r|r|r|r|} \hline
720 & NIST test & NISTP test & P07 test & NIST test digits \\ \hline
721 Humans& 18.2\% $\pm$.1\% & 39.4\%$\pm$.1\% & 46.9\%$\pm$.1\% & $1.4\%$ \\ \hline
722 SDA0 & 23.7\% $\pm$.14\% & 65.2\%$\pm$.34\% & 97.45\%$\pm$.06\% & 2.7\% $\pm$.14\%\\ \hline
723 SDA1 & 17.1\% $\pm$.13\% & 29.7\%$\pm$.3\% & 29.7\%$\pm$.3\% & 1.4\% $\pm$.1\%\\ \hline
724 SDA2 & 18.7\% $\pm$.13\% & 33.6\%$\pm$.3\% & 39.9\%$\pm$.17\% & 1.7\% $\pm$.1\%\\ \hline
725 MLP0 & 24.2\% $\pm$.15\% & 68.8\%$\pm$.33\% & 78.70\%$\pm$.14\% & 3.45\% $\pm$.15\% \\ \hline
726 MLP1 & 23.0\% $\pm$.15\% & 41.8\%$\pm$.35\% & 90.4\%$\pm$.1\% & 3.85\% $\pm$.16\% \\ \hline
727 MLP2 & 24.3\% $\pm$.15\% & 46.0\%$\pm$.35\% & 54.7\%$\pm$.17\% & 4.85\% $\pm$.18\% \\ \hline
728 \citep{Granger+al-2007} & & & & 4.95\% $\pm$.18\% \\ \hline
729 \citep{Cortes+al-2000} & & & & 3.71\% $\pm$.16\% \\ \hline
730 \citep{Oliveira+al-2002} & & & & 2.4\% $\pm$.13\% \\ \hline
731 \citep{Milgram+al-2005} & & & & 2.1\% $\pm$.12\% \\ \hline
732 \end{tabular}
733 \end{center}
734 \end{table}
735
736 \begin{table}[ht]
737 \caption{Relative change in error rates due to the use of perturbed training data,
738 either using NISTP, for the MLP1/SDA1 models, or using P07, for the MLP2/SDA2 models.
739 A positive value indicates that training on the perturbed data helped for the
740 given test set (the first 3 columns on the 62-class tasks and the last one is
741 on the clean 10-class digits). Clearly, the deep learning models did benefit more
742 from perturbed training data, even when testing on clean data, whereas the MLP
743 trained on perturbed data performed worse on the clean digits and about the same
744 on the clean characters. }
745 \label{tab:perturbation-effect}
746 \begin{center}
747 \begin{tabular}{|l|r|r|r|r|} \hline
748 & NIST test & NISTP test & P07 test & NIST test digits \\ \hline
749 SDA0/SDA1-1 & 38\% & 84\% & 228\% & 93\% \\ \hline
750 SDA0/SDA2-1 & 27\% & 94\% & 144\% & 59\% \\ \hline
751 MLP0/MLP1-1 & 5.2\% & 65\% & -13\% & -10\% \\ \hline
752 MLP0/MLP2-1 & -0.4\% & 49\% & 44\% & -29\% \\ \hline
753 \end{tabular}
754 \end{center}
755 \end{table}
756
757 \begin{table}[ht]
758 \caption{Test error rates and relative change in error rates due to the use of
759 a multi-task setting, i.e., training on each task in isolation vs training
760 for all three tasks together, for MLPs vs SDAs. The SDA benefits much
761 more from the multi-task setting. All experiments on only on the
762 unperturbed NIST data, using validation error for model selection.
763 Relative improvement is 1 - single-task error / multi-task error.}
764 \label{tab:multi-task}
765 \begin{center}
766 \begin{tabular}{|l|r|r|r|} \hline
767 & single-task & multi-task & relative \\
768 & setting & setting & improvement \\ \hline
769 MLP-digits & 3.77\% & 3.99\% & 5.6\% \\ \hline
770 MLP-lower & 17.4\% & 16.8\% & -4.1\% \\ \hline
771 MLP-upper & 7.84\% & 7.54\% & -3.6\% \\ \hline
772 SDA-digits & 2.6\% & 3.56\% & 27\% \\ \hline
773 SDA-lower & 12.3\% & 14.4\% & 15\% \\ \hline
774 SDA-upper & 5.93\% & 6.78\% & 13\% \\ \hline
775 \end{tabular}
776 \end{center}
777 \end{table}
778
779 \fi
780
781 %\afterpage{\clearpage}
782 %\clearpage
783 {
784 %\bibliographystyle{spbasic} % basic style, author-year citations
785 \bibliographystyle{plainnat}
786 \bibliography{strings,strings-short,strings-shorter,ift6266_ml,specials,aigaion-shorter}
787 %\bibliographystyle{unsrtnat}
788 %\bibliographystyle{apalike}
789 }
790
791
792 \end{document}