comparison writeup/nips2010_cameraready.tex @ 604:51213beaed8b

draft of NIPS 2010 workshop camera-ready version
author Yoshua Bengio <bengioy@iro.umontreal.ca>
date Mon, 22 Nov 2010 14:52:33 -0500
parents
children 63f838479510
comparison
equal deleted inserted replaced
603:eb6244c6d861 604:51213beaed8b
1 \documentclass{article} % For LaTeX2e
2 \usepackage{nips10submit_e,times}
3 \usepackage{wrapfig}
4 \usepackage{amsthm,amsmath,bbm}
5 \usepackage[psamsfonts]{amssymb}
6 \usepackage{algorithm,algorithmic}
7 \usepackage[utf8]{inputenc}
8 \usepackage{graphicx,subfigure}
9 \usepackage[numbers]{natbib}
10
11 \addtolength{\textwidth}{20mm}
12 \addtolength{\textheight}{20mm}
13 \addtolength{\topmargin}{-10mm}
14 \addtolength{\evensidemargin}{-10mm}
15 \addtolength{\oddsidemargin}{-10mm}
16
17 %\setlength\parindent{0mm}
18
19 \title{Deep Self-Taught Learning for Handwritten Character Recognition}
20 \author{
21 Frédéric Bastien,
22 Yoshua Bengio,
23 Arnaud Bergeron,
24 Nicolas Boulanger-Lewandowski,
25 Thomas Breuel,\\
26 {\bf Youssouf Chherawala,
27 Moustapha Cisse,
28 Myriam Côté,
29 Dumitru Erhan,
30 Jeremy Eustache,}\\
31 {\bf Xavier Glorot,
32 Xavier Muller,
33 Sylvain Pannetier Lebeuf,
34 Razvan Pascanu,} \\
35 {\bf Salah Rifai,
36 Francois Savard,
37 Guillaume Sicard}\\
38 Dept. IRO, U. Montreal
39 }
40
41 \begin{document}
42
43 %\makeanontitle
44 \maketitle
45
46 \vspace*{-2mm}
47 \begin{abstract}
48 Recent theoretical and empirical work in statistical machine learning has
49 demonstrated the importance of learning algorithms for deep
50 architectures, i.e., function classes obtained by composing multiple
51 non-linear transformations. Self-taught learning (exploiting unlabeled
52 examples or examples from other distributions) has already been applied
53 to deep learners, but mostly to show the advantage of unlabeled
54 examples. Here we explore the advantage brought by {\em out-of-distribution examples}.
55 For this purpose we
56 developed a powerful generator of stochastic variations and noise
57 processes for character images, including not only affine transformations
58 but also slant, local elastic deformations, changes in thickness,
59 background images, grey level changes, contrast, occlusion, and various
60 types of noise. The out-of-distribution examples are obtained from these
61 highly distorted images or by including examples of object classes
62 different from those in the target test set.
63 We show that {\em deep learners benefit
64 more from them than a corresponding shallow learner}, at least in the area of
65 handwritten character recognition. In fact, we show that they reach
66 human-level performance on both handwritten digit classification and
67 62-class handwritten character recognition.
68 \end{abstract}
69 \vspace*{-3mm}
70
71 \section{Introduction}
72 \vspace*{-1mm}
73
74 {\bf Deep Learning} has emerged as a promising new area of research in
75 statistical machine learning~\citep{Hinton06,ranzato-07-small,Bengio-nips-2006,VincentPLarochelleH2008-very-small,ranzato-08,TaylorHintonICML2009,Larochelle-jmlr-2009,Salakhutdinov+Hinton-2009,HonglakL2009,HonglakLNIPS2009,Jarrett-ICCV2009,Taylor-cvpr-2010}. See \citet{Bengio-2009} for a review.
76 Learning algorithms for deep architectures are centered on the learning
77 of useful representations of data, which are better suited to the task at hand,
78 and are organized in a hierarchy with multiple levels.
79 This is in part inspired by observations of the mammalian visual cortex,
80 which consists of a chain of processing elements, each of which is associated with a
81 different representation of the raw visual input. In fact,
82 it was found recently that the features learnt in deep architectures resemble
83 those observed in the first two of these stages (in areas V1 and V2
84 of visual cortex)~\citep{HonglakL2008}, and that they become more and
85 more invariant to factors of variation (such as camera movement) in
86 higher layers~\citep{Goodfellow2009}.
87 It has been hypothesized that learning a hierarchy of features increases the
88 ease and practicality of developing representations that are at once
89 tailored to specific tasks, yet are able to borrow statistical strength
90 from other related tasks (e.g., modeling different kinds of objects). Finally, learning the
91 feature representation can lead to higher-level (more abstract, more
92 general) features that are more robust to unanticipated sources of
93 variance extant in real data.
94
95 {\bf Self-taught learning}~\citep{RainaR2007} is a paradigm that combines principles
96 of semi-supervised and multi-task learning: the learner can exploit examples
97 that are unlabeled and possibly come from a distribution different from the target
98 distribution, e.g., from other classes than those of interest.
99 It has already been shown that deep learners can clearly take advantage of
100 unsupervised learning and unlabeled examples~\citep{Bengio-2009,WestonJ2008-small},
101 but more needs to be done to explore the impact
102 of {\em out-of-distribution} examples and of the multi-task setting
103 (one exception is~\citep{CollobertR2008}, which uses a different kind
104 of learning algorithm). In particular the {\em relative
105 advantage} of deep learning for these settings has not been evaluated.
106 The hypothesis discussed in the conclusion is that a deep hierarchy of features
107 may be better able to provide sharing of statistical strength
108 between different regions in input space or different tasks.
109
110 \iffalse
111 Whereas a deep architecture can in principle be more powerful than a
112 shallow one in terms of representation, depth appears to render the
113 training problem more difficult in terms of optimization and local minima.
114 It is also only recently that successful algorithms were proposed to
115 overcome some of these difficulties. All are based on unsupervised
116 learning, often in an greedy layer-wise ``unsupervised pre-training''
117 stage~\citep{Bengio-2009}.
118 The principle is that each layer starting from
119 the bottom is trained to represent its input (the output of the previous
120 layer). After this
121 unsupervised initialization, the stack of layers can be
122 converted into a deep supervised feedforward neural network and fine-tuned by
123 stochastic gradient descent.
124 One of these layer initialization techniques,
125 applied here, is the Denoising
126 Auto-encoder~(DA)~\citep{VincentPLarochelleH2008-very-small} (see
127 Figure~\ref{fig:da}), which performed similarly or
128 better~\citep{VincentPLarochelleH2008-very-small} than previously
129 proposed Restricted Boltzmann Machines (RBM)~\citep{Hinton06}
130 in terms of unsupervised extraction
131 of a hierarchy of features useful for classification. Each layer is trained
132 to denoise its input, creating a layer of features that can be used as
133 input for the next layer, forming a Stacked Denoising Auto-encoder (SDA).
134 Note that training a Denoising Auto-encoder
135 can actually been seen as training a particular RBM by an inductive
136 principle different from maximum likelihood~\citep{Vincent-SM-2010},
137 namely by Score Matching~\citep{Hyvarinen-2005,HyvarinenA2008}.
138 \fi
139
140 Previous comparative experimental results with stacking of RBMs and DAs
141 to build deep supervised predictors had shown that they could outperform
142 shallow architectures in a variety of settings, especially
143 when the data involves complex interactions between many factors of
144 variation~\citep{LarochelleH2007,Bengio-2009}. Other experiments have suggested
145 that the unsupervised layer-wise pre-training acted as a useful
146 prior~\citep{Erhan+al-2010} that allows one to initialize a deep
147 neural network in a relatively much smaller region of parameter space,
148 corresponding to better generalization.
149
150 To further the understanding of the reasons for the good performance
151 observed with deep learners, we focus here on the following {\em hypothesis}:
152 intermediate levels of representation, especially when there are
153 more such levels, can be exploited to {\bf share
154 statistical strength across different but related types of examples},
155 such as examples coming from other tasks than the task of interest
156 (the multi-task setting), or examples coming from an overlapping
157 but different distribution (images with different kinds of perturbations
158 and noises, here). This is consistent with the hypotheses discussed
159 in~\citet{Bengio-2009} regarding the potential advantage
160 of deep learning and the idea that more levels of representation can
161 give rise to more abstract, more general features of the raw input.
162
163 This hypothesis is related to a learning setting called
164 {\bf self-taught learning}~\citep{RainaR2007}, which combines principles
165 of semi-supervised and multi-task learning: the learner can exploit examples
166 that are unlabeled and possibly come from a distribution different from the target
167 distribution, e.g., from other classes than those of interest.
168 It has already been shown that deep learners can clearly take advantage of
169 unsupervised learning and unlabeled examples~\citep{Bengio-2009,WestonJ2008-small},
170 but more needed to be done to explore the impact
171 of {\em out-of-distribution} examples and of the {\em multi-task} setting
172 (one exception is~\citep{CollobertR2008}, which shares and uses unsupervised
173 pre-training only with the first layer). In particular the {\em relative
174 advantage of deep learning} for these settings has not been evaluated.
175
176
177 %
178 The {\bf main claim} of this paper is that deep learners (with several levels of representation) can
179 {\bf benefit more from out-of-distribution examples than shallow learners} (with a single
180 level), both in the context of the multi-task setting and from
181 perturbed examples. Because we are able to improve on state-of-the-art
182 performance and reach human-level performance
183 on a large-scale task, we consider that this paper is also a contribution
184 to advance the application of machine learning to handwritten character recognition.
185 More precisely, we ask and answer the following questions:
186
187 %\begin{enumerate}
188 $\bullet$ %\item
189 Do the good results previously obtained with deep architectures on the
190 MNIST digit images generalize to the setting of a similar but much larger and richer
191 dataset, the NIST special database 19, with 62 classes and around 800k examples?
192
193 $\bullet$ %\item
194 To what extent does the perturbation of input images (e.g. adding
195 noise, affine transformations, background images) make the resulting
196 classifiers better not only on similarly perturbed images but also on
197 the {\em original clean examples}? We study this question in the
198 context of the 62-class and 10-class tasks of the NIST special database 19.
199
200 $\bullet$ %\item
201 Do deep architectures {\em benefit {\bf more} from such out-of-distribution}
202 examples, in particular do they benefit more from
203 examples that are perturbed versions of the examples from the task of interest?
204
205 $\bullet$ %\item
206 Similarly, does the feature learning step in deep learning algorithms benefit {\bf more}
207 from training with moderately {\em different classes} (i.e. a multi-task learning scenario) than
208 a corresponding shallow and purely supervised architecture?
209 We train on 62 classes and test on 10 (digits) or 26 (upper case or lower case)
210 to answer this question.
211 %\end{enumerate}
212
213 Our experimental results provide positive evidence towards all of these questions,
214 as well as {\bf classifiers that reach human-level performance on 62-class isolated character
215 recognition and beat previously published results on the NIST dataset (special database 19)}.
216 To achieve these results, we introduce in the next section a sophisticated system
217 for stochastically transforming character images and then explain the methodology,
218 which is based on training with or without these transformed images and testing on
219 clean ones.
220 Code for generating these transformations as well as for the deep learning
221 algorithms are made available at {\tt http://hg.assembla.com/ift6266}.
222
223 \vspace*{-3mm}
224 %\newpage
225 \section{Perturbed and Transformed Character Images}
226 \label{s:perturbations}
227 \vspace*{-2mm}
228
229 \begin{minipage}[h]{\linewidth}
230 \begin{wrapfigure}[8]{l}{0.15\textwidth}
231 %\begin{minipage}[b]{0.14\linewidth}
232 \vspace*{-5mm}
233 \begin{center}
234 \includegraphics[scale=.4]{images/Original.png}\\
235 {\bf Original}
236 \end{center}
237 \end{wrapfigure}
238 %\vspace{0.7cm}
239 %\end{minipage}%
240 %\hspace{0.3cm}\begin{minipage}[b]{0.86\linewidth}
241 This section describes the different transformations we used to stochastically
242 transform $32 \times 32$ source images (such as the one on the left)
243 in order to obtain data from a larger distribution which
244 covers a domain substantially larger than the clean characters distribution from
245 which we start.
246 Although character transformations have been used before to
247 improve character recognizers, this effort is on a large scale both
248 in number of classes and in the complexity of the transformations, hence
249 in the complexity of the learning task.
250 More details can
251 be found in this technical report~\citep{ARXIV-2010}.
252 The code for these transformations (mostly python) is available at
253 {\tt http://hg.assembla.com/ift6266}. All the modules in the pipeline share
254 a global control parameter ($0 \le complexity \le 1$) that allows one to modulate the
255 amount of deformation or noise introduced.
256 There are two main parts in the pipeline. The first one,
257 from thickness to pinch, performs transformations. The second
258 part, from blur to contrast, adds different kinds of noise.
259 \end{minipage}
260
261 \vspace*{1mm}
262 %\subsection{Transformations}
263 {\large\bf 2.1 Transformations}
264 \vspace*{1mm}
265
266
267 \begin{minipage}[h]{\linewidth}
268 \begin{wrapfigure}[7]{l}{0.15\textwidth}
269 %\begin{minipage}[b]{0.14\linewidth}
270 %\centering
271 \begin{center}
272 \vspace*{-5mm}
273 \includegraphics[scale=.4]{images/Thick_only.png}\\
274 {\bf Thickness}
275 \end{center}
276 %\vspace{.6cm}
277 %\end{minipage}%
278 %\hspace{0.3cm}\begin{minipage}[b]{0.86\linewidth}
279 \end{wrapfigure}
280 To change character {\bf thickness}, morphological operators of dilation and erosion~\citep{Haralick87,Serra82}
281 are applied. The neighborhood of each pixel is multiplied
282 element-wise with a {\em structuring element} matrix.
283 The pixel value is replaced by the maximum or the minimum of the resulting
284 matrix, respectively for dilation or erosion. Ten different structural elements with
285 increasing dimensions (largest is $5\times5$) were used. For each image,
286 randomly sample the operator type (dilation or erosion) with equal probability and one structural
287 element from a subset of the $n=round(m \times complexity)$ smallest structuring elements
288 where $m=10$ for dilation and $m=6$ for erosion (to avoid completely erasing thin characters).
289 A neutral element (no transformation)
290 is always present in the set.
291 %\vspace{.4cm}
292 \end{minipage}
293 \vspace*{3mm}
294
295 \begin{minipage}[b]{0.14\linewidth}
296 \centering
297 \includegraphics[scale=.4]{images/Slant_only.png}\\
298 {\bf Slant}
299 \end{minipage}%
300 \hspace{0.3cm}
301 \begin{minipage}[b]{0.83\linewidth}
302 %\centering
303 To produce {\bf slant}, each row of the image is shifted
304 proportionally to its height: $shift = round(slant \times height)$.
305 $slant \sim U[-complexity,complexity]$.
306 The shift is randomly chosen to be either to the left or to the right.
307 \vspace{8mm}
308 \end{minipage}
309 \vspace*{3mm}
310
311 \begin{minipage}[h]{\linewidth}
312 \begin{minipage}[b]{0.14\linewidth}
313 %\centering
314 \begin{wrapfigure}[8]{l}{0.15\textwidth}
315 \vspace*{-6mm}
316 \begin{center}
317 \includegraphics[scale=.4]{images/Affine_only.png}\\
318 {\small {\bf Affine \mbox{Transformation}}}
319 \end{center}
320 \end{wrapfigure}
321 %\end{minipage}%
322 %\hspace{0.3cm}\begin{minipage}[b]{0.86\linewidth}
323 A $2 \times 3$ {\bf affine transform} matrix (with
324 parameters $(a,b,c,d,e,f)$) is sampled according to the $complexity$.
325 Output pixel $(x,y)$ takes the value of input pixel
326 nearest to $(ax+by+c,dx+ey+f)$,
327 producing scaling, translation, rotation and shearing.
328 Marginal distributions of $(a,b,c,d,e,f)$ have been tuned to
329 forbid large rotations (to avoid confusing classes) but to give good
330 variability of the transformation: $a$ and $d$ $\sim U[1-3
331 complexity,1+3\,complexity]$, $b$ and $e$ $\sim U[-3 \,complexity,3\,
332 complexity]$, and $c$ and $f \sim U[-4 \,complexity, 4 \,
333 complexity]$.\\
334 \end{minipage}
335 \end{minipage}
336
337 \iffalse
338 \vspace*{-4.5mm}
339
340 \begin{minipage}[h]{\linewidth}
341 \begin{wrapfigure}[7]{l}{0.15\textwidth}
342 %\hspace*{-8mm}\begin{minipage}[b]{0.25\linewidth}
343 %\centering
344 \begin{center}
345 \vspace*{-4mm}
346 \includegraphics[scale=.4]{images/Localelasticdistorsions_only.png}\\
347 {\bf Local Elastic Deformation}
348 \end{center}
349 \end{wrapfigure}
350 %\end{minipage}%
351 %\hspace{-3mm}\begin{minipage}[b]{0.85\linewidth}
352 %\vspace*{-20mm}
353 The {\bf local elastic deformation}
354 module induces a ``wiggly'' effect in the image, following~\citet{SimardSP03-short},
355 which provides more details.
356 The intensity of the displacement fields is given by
357 $\alpha = \sqrt[3]{complexity} \times 10.0$, which are
358 convolved with a Gaussian 2D kernel (resulting in a blur) of
359 standard deviation $\sigma = 10 - 7 \times\sqrt[3]{complexity}$.
360 %\vspace{.9cm}
361 \end{minipage}
362
363 \vspace*{7mm}
364
365 %\begin{minipage}[b]{0.14\linewidth}
366 %\centering
367 \begin{minipage}[h]{\linewidth}
368 \begin{wrapfigure}[7]{l}{0.15\textwidth}
369 \vspace*{-5mm}
370 \begin{center}
371 \includegraphics[scale=.4]{images/Pinch_only.png}\\
372 {\bf Pinch}
373 \end{center}
374 \end{wrapfigure}
375 %\vspace{.6cm}
376 %\end{minipage}%
377 %\hspace{0.3cm}\begin{minipage}[b]{0.86\linewidth}
378 The {\bf pinch} module applies the ``Whirl and pinch'' GIMP filter with whirl set to 0.
379 A pinch is ``similar to projecting the image onto an elastic
380 surface and pressing or pulling on the center of the surface'' (GIMP documentation manual).
381 For a square input image, draw a radius-$r$ disk
382 around its center $C$. Any pixel $P$ belonging to
383 that disk has its value replaced by
384 the value of a ``source'' pixel in the original image,
385 on the line that goes through $C$ and $P$, but
386 at some other distance $d_2$. Define $d_1=distance(P,C)$
387 and $d_2 = sin(\frac{\pi{}d_1}{2r})^{-pinch} \times
388 d_1$, where $pinch$ is a parameter of the filter.
389 The actual value is given by bilinear interpolation considering the pixels
390 around the (non-integer) source position thus found.
391 Here $pinch \sim U[-complexity, 0.7 \times complexity]$.
392 %\vspace{1.5cm}
393 \end{minipage}
394
395 \vspace{1mm}
396
397 {\large\bf 2.2 Injecting Noise}
398 %\subsection{Injecting Noise}
399 \vspace{2mm}
400
401 \begin{minipage}[h]{\linewidth}
402 %\vspace*{-.2cm}
403 \begin{minipage}[t]{0.14\linewidth}
404 \centering
405 \vspace*{-2mm}
406 \includegraphics[scale=.4]{images/Motionblur_only.png}\\
407 {\bf Motion Blur}
408 \end{minipage}%
409 \hspace{0.3cm}\begin{minipage}[t]{0.83\linewidth}
410 %\vspace*{.5mm}
411 The {\bf motion blur} module is GIMP's ``linear motion blur'', which
412 has parameters $length$ and $angle$. The value of
413 a pixel in the final image is approximately the mean of the first $length$ pixels
414 found by moving in the $angle$ direction,
415 $angle \sim U[0,360]$ degrees, and $length \sim {\rm Normal}(0,(3 \times complexity)^2)$.
416 \vspace{5mm}
417 \end{minipage}
418 \end{minipage}
419
420 \vspace*{1mm}
421
422 \begin{minipage}[h]{\linewidth}
423 \begin{minipage}[t]{0.14\linewidth}
424 \centering
425 \includegraphics[scale=.4]{images/occlusion_only.png}\\
426 {\bf Occlusion}
427 %\vspace{.5cm}
428 \end{minipage}%
429 \hspace{0.3cm}\begin{minipage}[t]{0.83\linewidth}
430 \vspace*{-18mm}
431 The {\bf occlusion} module selects a random rectangle from an {\em occluder} character
432 image and places it over the original {\em occluded}
433 image. Pixels are combined by taking the max(occluder, occluded),
434 i.e. keeping the lighter ones.
435 The rectangle corners
436 are sampled so that larger complexity gives larger rectangles.
437 The destination position in the occluded image are also sampled
438 according to a normal distribution (more details in~\citet{ift6266-tr-anonymous}).
439 This module is skipped with probability 60\%.
440 %\vspace{7mm}
441 \end{minipage}
442 \end{minipage}
443
444 \vspace*{1mm}
445
446 \begin{wrapfigure}[8]{l}{0.15\textwidth}
447 \vspace*{-6mm}
448 \begin{center}
449 %\begin{minipage}[t]{0.14\linewidth}
450 %\centering
451 \includegraphics[scale=.4]{images/Bruitgauss_only.png}\\
452 {\bf Gaussian Smoothing}
453 \end{center}
454 \end{wrapfigure}
455 %\vspace{.5cm}
456 %\end{minipage}%
457 %\hspace{0.3cm}\begin{minipage}[t]{0.86\linewidth}
458 With the {\bf Gaussian smoothing} module,
459 different regions of the image are spatially smoothed.
460 This is achieved by first convolving
461 the image with an isotropic Gaussian kernel of
462 size and variance chosen uniformly in the ranges $[12,12 + 20 \times
463 complexity]$ and $[2,2 + 6 \times complexity]$. This filtered image is normalized
464 between $0$ and $1$. We also create an isotropic weighted averaging window, of the
465 kernel size, with maximum value at the center. For each image we sample
466 uniformly from $3$ to $3 + 10 \times complexity$ pixels that will be
467 averaging centers between the original image and the filtered one. We
468 initialize to zero a mask matrix of the image size. For each selected pixel
469 we add to the mask the averaging window centered on it. The final image is
470 computed from the following element-wise operation: $\frac{image + filtered\_image
471 \times mask}{mask+1}$.
472 This module is skipped with probability 75\%.
473 %\end{minipage}
474
475 \newpage
476
477 \vspace*{-9mm}
478
479 %\hspace*{-3mm}\begin{minipage}[t]{0.18\linewidth}
480 %\centering
481 \begin{minipage}[t]{\linewidth}
482 \begin{wrapfigure}[7]{l}{0.15\textwidth}
483 \vspace*{-5mm}
484 \begin{center}
485 \includegraphics[scale=.4]{images/Permutpixel_only.png}\\
486 {\small\bf Permute Pixels}
487 \end{center}
488 \end{wrapfigure}
489 %\end{minipage}%
490 %\hspace{-0cm}\begin{minipage}[t]{0.86\linewidth}
491 %\vspace*{-20mm}
492 This module {\bf permutes neighbouring pixels}. It first selects a
493 fraction $\frac{complexity}{3}$ of pixels randomly in the image. Each
494 of these pixels is then sequentially exchanged with a random pixel
495 among its four nearest neighbors (on its left, right, top or bottom).
496 This module is skipped with probability 80\%.\\
497 \vspace*{1mm}
498 \end{minipage}
499
500 \vspace{-3mm}
501
502 \begin{minipage}[t]{\linewidth}
503 \begin{wrapfigure}[7]{l}{0.15\textwidth}
504 %\vspace*{-3mm}
505 \begin{center}
506 %\hspace*{-3mm}\begin{minipage}[t]{0.18\linewidth}
507 %\centering
508 \vspace*{-5mm}
509 \includegraphics[scale=.4]{images/Distorsiongauss_only.png}\\
510 {\small \bf Gauss. Noise}
511 \end{center}
512 \end{wrapfigure}
513 %\end{minipage}%
514 %\hspace{0.3cm}\begin{minipage}[t]{0.86\linewidth}
515 \vspace*{12mm}
516 The {\bf Gaussian noise} module simply adds, to each pixel of the image independently, a
517 noise $\sim Normal(0,(\frac{complexity}{10})^2)$.
518 This module is skipped with probability 70\%.
519 %\vspace{1.1cm}
520 \end{minipage}
521
522 \vspace*{1.2cm}
523
524 \begin{minipage}[t]{\linewidth}
525 \begin{minipage}[t]{0.14\linewidth}
526 \centering
527 \includegraphics[scale=.4]{images/background_other_only.png}\\
528 {\small \bf Bg Image}
529 \end{minipage}%
530 \hspace{0.3cm}\begin{minipage}[t]{0.83\linewidth}
531 \vspace*{-18mm}
532 Following~\citet{Larochelle-jmlr-2009}, the {\bf background image} module adds a random
533 background image behind the letter, from a randomly chosen natural image,
534 with contrast adjustments depending on $complexity$, to preserve
535 more or less of the original character image.
536 %\vspace{.8cm}
537 \end{minipage}
538 \end{minipage}
539 %\vspace{-.7cm}
540
541 \begin{minipage}[t]{0.14\linewidth}
542 \centering
543 \includegraphics[scale=.4]{images/Poivresel_only.png}\\
544 {\small \bf Salt \& Pepper}
545 \end{minipage}%
546 \hspace{0.3cm}\begin{minipage}[t]{0.83\linewidth}
547 \vspace*{-18mm}
548 The {\bf salt and pepper noise} module adds noise $\sim U[0,1]$ to random subsets of pixels.
549 The number of selected pixels is $0.2 \times complexity$.
550 This module is skipped with probability 75\%.
551 %\vspace{.9cm}
552 \end{minipage}
553 %\vspace{-.7cm}
554
555 \vspace{1mm}
556
557 \begin{minipage}[t]{\linewidth}
558 \begin{wrapfigure}[7]{l}{0.14\textwidth}
559 %\begin{minipage}[t]{0.14\linewidth}
560 %\centering
561 \begin{center}
562 \vspace*{-4mm}
563 \hspace*{-1mm}\includegraphics[scale=.4]{images/Rature_only.png}\\
564 {\bf Scratches}
565 %\end{minipage}%
566 \end{center}
567 \end{wrapfigure}
568 %\hspace{0.3cm}\begin{minipage}[t]{0.86\linewidth}
569 %\vspace{.4cm}
570 The {\bf scratches} module places line-like white patches on the image. The
571 lines are heavily transformed images of the digit ``1'' (one), chosen
572 at random among 500 such 1 images,
573 randomly cropped and rotated by an angle $\sim Normal(0,(100 \times
574 complexity)^2$ (in degrees), using bi-cubic interpolation.
575 Two passes of a grey-scale morphological erosion filter
576 are applied, reducing the width of the line
577 by an amount controlled by $complexity$.
578 This module is skipped with probability 85\%. The probabilities
579 of applying 1, 2, or 3 patches are (50\%,30\%,20\%).
580 \end{minipage}
581
582 \vspace*{1mm}
583
584 \begin{minipage}[t]{0.25\linewidth}
585 \centering
586 \hspace*{-16mm}\includegraphics[scale=.4]{images/Contrast_only.png}\\
587 {\bf Grey Level \& Contrast}
588 \end{minipage}%
589 \hspace{-12mm}\begin{minipage}[t]{0.82\linewidth}
590 \vspace*{-18mm}
591 The {\bf grey level and contrast} module changes the contrast by changing grey levels, and may invert the image polarity (white
592 to black and black to white). The contrast is $C \sim U[1-0.85 \times complexity,1]$
593 so the image is normalized into $[\frac{1-C}{2},1-\frac{1-C}{2}]$. The
594 polarity is inverted with probability 50\%.
595 %\vspace{.7cm}
596 \end{minipage}
597 \vspace{2mm}
598
599 \fi
600
601 \iffalse
602 \begin{figure}[ht]
603 \centerline{\resizebox{.9\textwidth}{!}{\includegraphics{images/example_t.png}}}\\
604 \caption{Illustration of the pipeline of stochastic
605 transformations applied to the image of a lower-case \emph{t}
606 (the upper left image). Each image in the pipeline (going from
607 left to right, first top line, then bottom line) shows the result
608 of applying one of the modules in the pipeline. The last image
609 (bottom right) is used as training example.}
610 \label{fig:pipeline}
611 \end{figure}
612 \fi
613
614 \vspace*{-3mm}
615 \section{Experimental Setup}
616 \vspace*{-1mm}
617
618 Much previous work on deep learning had been performed on
619 the MNIST digits task~\citep{Hinton06,ranzato-07-small,Bengio-nips-2006,Salakhutdinov+Hinton-2009},
620 with 60~000 examples, and variants involving 10~000
621 examples~\citep{Larochelle-jmlr-toappear-2008,VincentPLarochelleH2008}.
622 The focus here is on much larger training sets, from 10 times to
623 to 1000 times larger, and 62 classes.
624
625 The first step in constructing the larger datasets (called NISTP and P07) is to sample from
626 a {\em data source}: {\bf NIST} (NIST database 19), {\bf Fonts}, {\bf Captchas},
627 and {\bf OCR data} (scanned machine printed characters). Once a character
628 is sampled from one of these {\em data sources} (chosen randomly), the second step is to
629 apply a pipeline of transformations and/or noise processes described in section \ref{s:perturbations}.
630
631 To provide a baseline of error rate comparison we also estimate human performance
632 on both the 62-class task and the 10-class digits task.
633 We compare the best Multi-Layer Perceptrons (MLP) against
634 the best Stacked Denoising Auto-encoders (SDA), when
635 both models' hyper-parameters are selected to minimize the validation set error.
636 We also provide a comparison against a precise estimate
637 of human performance obtained via Amazon's Mechanical Turk (AMT)
638 service ({\tt http://mturk.com}).
639 AMT users are paid small amounts
640 of money to perform tasks for which human intelligence is required.
641 An incentive for them to do the job right is that payment can be denied
642 if the job is not properly done.
643 Mechanical Turk has been used extensively in natural language processing and vision.
644 %processing \citep{SnowEtAl2008} and vision
645 %\citep{SorokinAndForsyth2008,whitehill09}.
646 AMT users were presented
647 with 10 character images at a time (from a test set) and asked to choose 10 corresponding ASCII
648 characters. They were forced to choose a single character class (either among the
649 62 or 10 character classes) for each image.
650 80 subjects classified 2500 images per (dataset,task) pair.
651 Different humans labelers sometimes provided a different label for the same
652 example, and we were able to estimate the error variance due to this effect
653 because each image was classified by 3 different persons.
654 The average error of humans on the 62-class task NIST test set
655 is 18.2\%, with a standard error of 0.1\%.
656
657 \vspace*{-3mm}
658 \subsection{Data Sources}
659 \vspace*{-2mm}
660
661 %\begin{itemize}
662 %\item
663 {\bf NIST.}
664 Our main source of characters is the NIST Special Database 19~\citep{Grother-1995},
665 widely used for training and testing character
666 recognition systems~\citep{Granger+al-2007,Cortes+al-2000,Oliveira+al-2002-short,Milgram+al-2005}.
667 The dataset is composed of 814255 digits and characters (upper and lower cases), with hand checked classifications,
668 extracted from handwritten sample forms of 3600 writers. The characters are labelled by one of the 62 classes
669 corresponding to ``0''-``9'',``A''-``Z'' and ``a''-``z''. The dataset contains 8 parts (partitions) of varying complexity.
670 The fourth partition (called $hsf_4$, 82587 examples),
671 experimentally recognized to be the most difficult one, is the one recommended
672 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}
673 for that purpose. We randomly split the remainder (731,668 examples) into a training set and a validation set for
674 model selection.
675 The performances reported by previous work on that dataset mostly use only the digits.
676 Here we use all the classes both in the training and testing phase. This is especially
677 useful to estimate the effect of a multi-task setting.
678 The distribution of the classes in the NIST training and test sets differs
679 substantially, with relatively many more digits in the test set, and a more uniform distribution
680 of letters in the test set (whereas in the training set they are distributed
681 more like in natural text).
682 \vspace*{-1mm}
683
684 %\item
685 {\bf Fonts.}
686 In order to have a good variety of sources we downloaded an important number of free fonts from:
687 {\tt http://cg.scs.carleton.ca/\textasciitilde luc/freefonts.html}.
688 % TODO: pointless to anonymize, it's not pointing to our work
689 Including an operating system's (Windows 7) fonts, there is a total of $9817$ different fonts that we can choose uniformly from.
690 The chosen {\tt ttf} file is either used as input of the Captcha generator (see next item) or, by producing a corresponding image,
691 directly as input to our models.
692 \vspace*{-1mm}
693
694 %\item
695 {\bf Captchas.}
696 The Captcha data source is an adaptation of the \emph{pycaptcha} library (a Python-based captcha generator library) for
697 generating characters of the same format as the NIST dataset. This software is based on
698 a random character class generator and various kinds of transformations similar to those described in the previous sections.
699 In order to increase the variability of the data generated, many different fonts are used for generating the characters.
700 Transformations (slant, distortions, rotation, translation) are applied to each randomly generated character with a complexity
701 depending on the value of the complexity parameter provided by the user of the data source.
702 %Two levels of complexity are allowed and can be controlled via an easy to use facade class. %TODO: what's a facade class?
703 \vspace*{-1mm}
704
705 %\item
706 {\bf OCR data.}
707 A large set (2 million) of scanned, OCRed and manually verified machine-printed
708 characters where included as an
709 additional source. This set is part of a larger corpus being collected by the Image Understanding
710 Pattern Recognition Research group led by Thomas Breuel at University of Kaiserslautern
711 ({\tt http://www.iupr.com}), and which will be publicly released.
712 %TODO: let's hope that Thomas is not a reviewer! :) Seriously though, maybe we should anonymize this
713 %\end{itemize}
714
715 \vspace*{-3mm}
716 \subsection{Data Sets}
717 \vspace*{-2mm}
718
719 All data sets contain 32$\times$32 grey-level images (values in $[0,1]$) associated with a label
720 from one of the 62 character classes. They are obtained from the optional application of the
721 perturbation pipeline to iid samples from the datasources, and they are randomly split into
722 training set, validation set, and test set.
723 %\begin{itemize}
724 \vspace*{-1mm}
725
726 %\item
727 {\bf NIST.} This is the raw NIST special database 19~\citep{Grother-1995}. It has
728 \{651668 / 80000 / 82587\} \{training / validation / test\} examples, containing
729 upper case, lower case, and digits.
730 \vspace*{-1mm}
731
732 %\item
733 {\bf P07.} This dataset of upper case, lower case and digit images
734 is obtained by taking raw characters from all four of the above sources
735 and sending them through the transformation pipeline described in section \ref{s:perturbations}.
736 For each new example to generate, a data source is selected with probability $10\%$ from the fonts,
737 $25\%$ from the captchas, $25\%$ from the OCR data and $40\%$ from NIST. We apply all the transformations in the
738 order given above, and for each of them we sample uniformly a \emph{complexity} in the range $[0,0.7]$.
739 It has \{81920000 / 80000 / 20000\} \{training / validation / test\} examples.
740 \vspace*{-1mm}
741
742 %\item
743 {\bf NISTP.} This one is equivalent to P07 (complexity parameter of $0.7$ with the same proportions of data sources)
744 except that we only apply
745 transformations from slant to pinch. Therefore, the character is
746 transformed but no additional noise is added to the image, giving images
747 closer to the NIST dataset.
748 It has \{81,920,000 / 80,000 / 20,000\} \{training / validation / test\} examples
749 obtained from the corresponding NIST sets plus other sources.
750 %\end{itemize}
751
752 \vspace*{-3mm}
753 \subsection{Models and their Hyperparameters}
754 \vspace*{-2mm}
755
756 The experiments are performed using MLPs (with a single
757 hidden layer) and deep SDAs.
758 \emph{Hyper-parameters are selected based on the {\bf NISTP} validation set error.}
759
760 {\bf Multi-Layer Perceptrons (MLP).}
761 Whereas previous work had compared deep architectures to both shallow MLPs and
762 SVMs, we only compared to MLPs here because of the very large datasets used
763 (making the use of SVMs computationally challenging because of their quadratic
764 scaling behavior). Preliminary experiments on training SVMs (libSVM) with subsets of the training
765 set allowing the program to fit in memory yielded substantially worse results
766 than those obtained with MLPs. For training on nearly a hundred million examples
767 (with the perturbed data), the MLPs and SDA are much more convenient than
768 classifiers based on kernel methods.
769 The MLP has a single hidden layer with $\tanh$ activation functions, and softmax (normalized
770 exponentials) on the output layer for estimating $P(class | image)$.
771 The number of hidden units is taken in $\{300,500,800,1000,1500\}$.
772 Training examples are presented in minibatches of size 20. A constant learning
773 rate was chosen among $\{0.001, 0.01, 0.025, 0.075, 0.1, 0.5\}$.
774 %through preliminary experiments (measuring performance on a validation set),
775 %and $0.1$ (which was found to work best) was then selected for optimizing on
776 %the whole training sets.
777 \vspace*{-1mm}
778
779
780 {\bf Stacked Denoising Auto-encoders (SDA).}
781 Various auto-encoder variants and Restricted Boltzmann Machines (RBMs)
782 can be used to initialize the weights of each layer of a deep MLP (with many hidden
783 layers)~\citep{Hinton06,ranzato-07-small,Bengio-nips-2006},
784 apparently setting parameters in the
785 basin of attraction of supervised gradient descent yielding better
786 generalization~\citep{Erhan+al-2010}. This initial {\em unsupervised
787 pre-training phase} uses all of the training images but not the training labels.
788 Each layer is trained in turn to produce a new representation of its input
789 (starting from the raw pixels).
790 It is hypothesized that the
791 advantage brought by this procedure stems from a better prior,
792 on the one hand taking advantage of the link between the input
793 distribution $P(x)$ and the conditional distribution of interest
794 $P(y|x)$ (like in semi-supervised learning), and on the other hand
795 taking advantage of the expressive power and bias implicit in the
796 deep architecture (whereby complex concepts are expressed as
797 compositions of simpler ones through a deep hierarchy).
798
799 \begin{figure}[ht]
800 \vspace*{-2mm}
801 \centerline{\resizebox{0.8\textwidth}{!}{\includegraphics{images/denoising_autoencoder_small.pdf}}}
802 \vspace*{-2mm}
803 \caption{Illustration of the computations and training criterion for the denoising
804 auto-encoder used to pre-train each layer of the deep architecture. Input $x$ of
805 the layer (i.e. raw input or output of previous layer)
806 s corrupted into $\tilde{x}$ and encoded into code $y$ by the encoder $f_\theta(\cdot)$.
807 The decoder $g_{\theta'}(\cdot)$ maps $y$ to reconstruction $z$, which
808 is compared to the uncorrupted input $x$ through the loss function
809 $L_H(x,z)$, whose expected value is approximately minimized during training
810 by tuning $\theta$ and $\theta'$.}
811 \label{fig:da}
812 \vspace*{-2mm}
813 \end{figure}
814
815 Here we chose to use the Denoising
816 Auto-encoder~\citep{VincentPLarochelleH2008} as the building block for
817 these deep hierarchies of features, as it is simple to train and
818 explain (see Figure~\ref{fig:da}, as well as
819 tutorial and code there: {\tt http://deeplearning.net/tutorial}),
820 provides efficient inference, and yielded results
821 comparable or better than RBMs in series of experiments
822 \citep{VincentPLarochelleH2008-very-small}. It really corresponds to a Gaussian
823 RBM trained by a Score Matching criterion~\cite{Vincent-SM-2010}.
824 During training, a Denoising
825 Auto-encoder is presented with a stochastically corrupted version
826 of the input and trained to reconstruct the uncorrupted input,
827 forcing the hidden units to represent the leading regularities in
828 the data. Here we use the random binary masking corruption
829 (which sets to 0 a random subset of the inputs).
830 Once it is trained, in a purely unsupervised way,
831 its hidden units' activations can
832 be used as inputs for training a second one, etc.
833 After this unsupervised pre-training stage, the parameters
834 are used to initialize a deep MLP, which is fine-tuned by
835 the same standard procedure used to train them (see previous section).
836 The SDA hyper-parameters are the same as for the MLP, with the addition of the
837 amount of corruption noise (we used the masking noise process, whereby a
838 fixed proportion of the input values, randomly selected, are zeroed), and a
839 separate learning rate for the unsupervised pre-training stage (selected
840 from the same above set). The fraction of inputs corrupted was selected
841 among $\{10\%, 20\%, 50\%\}$. Another hyper-parameter is the number
842 of hidden layers but it was fixed to 3 based on previous work with
843 SDAs on MNIST~\citep{VincentPLarochelleH2008-very-small}. The size of the hidden
844 layers was kept constant across hidden layers, and the best results
845 were obtained with the largest values that we could experiment
846 with given our patience, with 1000 hidden units.
847
848 \vspace*{-1mm}
849
850 \begin{figure}[ht]
851 \vspace*{-2mm}
852 \centerline{\resizebox{.99\textwidth}{!}{\includegraphics{images/error_rates_charts.pdf}}}
853 \vspace*{-3mm}
854 \caption{SDAx are the {\bf deep} models. Error bars indicate a 95\% confidence interval. 0 indicates that the model was trained
855 on NIST, 1 on NISTP, and 2 on P07. Left: overall results
856 of all models, on NIST and NISTP test sets.
857 Right: error rates on NIST test digits only, along with the previous results from
858 literature~\citep{Granger+al-2007,Cortes+al-2000,Oliveira+al-2002-short,Milgram+al-2005}
859 respectively based on ART, nearest neighbors, MLPs, and SVMs.}
860 \label{fig:error-rates-charts}
861 \vspace*{-2mm}
862 \end{figure}
863
864
865 \begin{figure}[ht]
866 \vspace*{-3mm}
867 \centerline{\resizebox{.99\textwidth}{!}{\includegraphics{images/improvements_charts.pdf}}}
868 \vspace*{-3mm}
869 \caption{Relative improvement in error rate due to self-taught learning.
870 Left: Improvement (or loss, when negative)
871 induced by out-of-distribution examples (perturbed data).
872 Right: Improvement (or loss, when negative) induced by multi-task
873 learning (training on all classes and testing only on either digits,
874 upper case, or lower-case). The deep learner (SDA) benefits more from
875 both self-taught learning scenarios, compared to the shallow MLP.}
876 \label{fig:improvements-charts}
877 \vspace*{-2mm}
878 \end{figure}
879
880 \section{Experimental Results}
881 \vspace*{-2mm}
882
883 %\vspace*{-1mm}
884 %\subsection{SDA vs MLP vs Humans}
885 %\vspace*{-1mm}
886 The models are either trained on NIST (MLP0 and SDA0),
887 NISTP (MLP1 and SDA1), or P07 (MLP2 and SDA2), and tested
888 on either NIST, NISTP or P07 (regardless of the data set used for training),
889 either on the 62-class task
890 or on the 10-digits task. Training time (including about half
891 for unsupervised pre-training, for DAs) on the larger
892 datasets is around one day on a GPU (GTX 285).
893 Figure~\ref{fig:error-rates-charts} summarizes the results obtained,
894 comparing humans, the three MLPs (MLP0, MLP1, MLP2) and the three SDAs (SDA0, SDA1,
895 SDA2), along with the previous results on the digits NIST special database
896 19 test set from the literature, respectively based on ARTMAP neural
897 networks ~\citep{Granger+al-2007}, fast nearest-neighbor search
898 ~\citep{Cortes+al-2000}, MLPs ~\citep{Oliveira+al-2002-short}, and SVMs
899 ~\citep{Milgram+al-2005}.% More detailed and complete numerical results
900 %(figures and tables, including standard errors on the error rates) can be
901 %found in Appendix.
902 The deep learner not only outperformed the shallow ones and
903 previously published performance (in a statistically and qualitatively
904 significant way) but when trained with perturbed data
905 reaches human performance on both the 62-class task
906 and the 10-class (digits) task.
907 17\% error (SDA1) or 18\% error (humans) may seem large but a large
908 majority of the errors from humans and from SDA1 are from out-of-context
909 confusions (e.g. a vertical bar can be a ``1'', an ``l'' or an ``L'', and a
910 ``c'' and a ``C'' are often indistinguishible).
911
912 In addition, as shown in the left of
913 Figure~\ref{fig:improvements-charts}, the relative improvement in error
914 rate brought by self-taught learning is greater for the SDA, and these
915 differences with the MLP are statistically and qualitatively
916 significant.
917 The left side of the figure shows the improvement to the clean
918 NIST test set error brought by the use of out-of-distribution examples
919 (i.e. the perturbed examples examples from NISTP or P07),
920 over the models trained exclusively on NIST (respectively SDA0 and MLP0).
921 Relative percent change is measured by taking
922 $100 \% \times$ (original model's error / perturbed-data model's error - 1).
923 The right side of
924 Figure~\ref{fig:improvements-charts} shows the relative improvement
925 brought by the use of a multi-task setting, in which the same model is
926 trained for more classes than the target classes of interest (i.e. training
927 with all 62 classes when the target classes are respectively the digits,
928 lower-case, or upper-case characters). Again, whereas the gain from the
929 multi-task setting is marginal or negative for the MLP, it is substantial
930 for the SDA. Note that to simplify these multi-task experiments, only the original
931 NIST dataset is used. For example, the MLP-digits bar shows the relative
932 percent improvement in MLP error rate on the NIST digits test set
933 is $100\% \times$ (single-task
934 model's error / multi-task model's error - 1). The single-task model is
935 trained with only 10 outputs (one per digit), seeing only digit examples,
936 whereas the multi-task model is trained with 62 outputs, with all 62
937 character classes as examples. Hence the hidden units are shared across
938 all tasks. For the multi-task model, the digit error rate is measured by
939 comparing the correct digit class with the output class associated with the
940 maximum conditional probability among only the digit classes outputs. The
941 setting is similar for the other two target classes (lower case characters
942 and upper case characters). Note however that some types of perturbations
943 (NISTP) help more than others (P07) when testing on the clean images.
944 %%\vspace*{-1mm}
945 %\subsection{Perturbed Training Data More Helpful for SDA}
946 %\vspace*{-1mm}
947
948 %\vspace*{-1mm}
949 %\subsection{Multi-Task Learning Effects}
950 %\vspace*{-1mm}
951
952 \iffalse
953 As previously seen, the SDA is better able to benefit from the
954 transformations applied to the data than the MLP. In this experiment we
955 define three tasks: recognizing digits (knowing that the input is a digit),
956 recognizing upper case characters (knowing that the input is one), and
957 recognizing lower case characters (knowing that the input is one). We
958 consider the digit classification task as the target task and we want to
959 evaluate whether training with the other tasks can help or hurt, and
960 whether the effect is different for MLPs versus SDAs. The goal is to find
961 out if deep learning can benefit more (or less) from multiple related tasks
962 (i.e. the multi-task setting) compared to a corresponding purely supervised
963 shallow learner.
964
965 We use a single hidden layer MLP with 1000 hidden units, and a SDA
966 with 3 hidden layers (1000 hidden units per layer), pre-trained and
967 fine-tuned on NIST.
968
969 Our results show that the MLP benefits marginally from the multi-task setting
970 in the case of digits (5\% relative improvement) but is actually hurt in the case
971 of characters (respectively 3\% and 4\% worse for lower and upper class characters).
972 On the other hand the SDA benefited from the multi-task setting, with relative
973 error rate improvements of 27\%, 15\% and 13\% respectively for digits,
974 lower and upper case characters, as shown in Table~\ref{tab:multi-task}.
975 \fi
976
977
978 \vspace*{-2mm}
979 \section{Conclusions and Discussion}
980 \vspace*{-2mm}
981
982 We have found that the self-taught learning framework is more beneficial
983 to a deep learner than to a traditional shallow and purely
984 supervised learner. More precisely,
985 the answers are positive for all the questions asked in the introduction.
986 %\begin{itemize}
987
988 $\bullet$ %\item
989 {\bf Do the good results previously obtained with deep architectures on the
990 MNIST digits generalize to a much larger and richer (but similar)
991 dataset, the NIST special database 19, with 62 classes and around 800k examples}?
992 Yes, the SDA {\em systematically outperformed the MLP and all the previously
993 published results on this dataset} (the ones that we are aware of), {\em in fact reaching human-level
994 performance} at around 17\% error on the 62-class task and 1.4\% on the digits,
995 and beating previously published results on the same data.
996
997 $\bullet$ %\item
998 {\bf To what extent do self-taught learning scenarios help deep learners,
999 and do they help them more than shallow supervised ones}?
1000 We found that distorted training examples not only made the resulting
1001 classifier better on similarly perturbed images but also on
1002 the {\em original clean examples}, and more importantly and more novel,
1003 that deep architectures benefit more from such {\em out-of-distribution}
1004 examples. MLPs were helped by perturbed training examples when tested on perturbed input
1005 images (65\% relative improvement on NISTP)
1006 but only marginally helped (5\% relative improvement on all classes)
1007 or even hurt (10\% relative loss on digits)
1008 with respect to clean examples. On the other hand, the deep SDAs
1009 were significantly boosted by these out-of-distribution examples.
1010 Similarly, whereas the improvement due to the multi-task setting was marginal or
1011 negative for the MLP (from +5.6\% to -3.6\% relative change),
1012 it was quite significant for the SDA (from +13\% to +27\% relative change),
1013 which may be explained by the arguments below.
1014 %\end{itemize}
1015
1016 In the original self-taught learning framework~\citep{RainaR2007}, the
1017 out-of-sample examples were used as a source of unsupervised data, and
1018 experiments showed its positive effects in a \emph{limited labeled data}
1019 scenario. However, many of the results by \citet{RainaR2007} (who used a
1020 shallow, sparse coding approach) suggest that the {\em relative gain of self-taught
1021 learning vs ordinary supervised learning} diminishes as the number of labeled examples increases.
1022 We note instead that, for deep
1023 architectures, our experiments show that such a positive effect is accomplished
1024 even in a scenario with a \emph{large number of labeled examples},
1025 i.e., here, the relative gain of self-taught learning and
1026 out-of-distribution examples is probably preserved
1027 in the asymptotic regime. However, note that in our perturbation experiments
1028 (but not in our multi-task experiments),
1029 even the out-of-distribution examples are labeled, unlike in the
1030 earlier self-taught learning experiments~\citep{RainaR2007}.
1031
1032 {\bf Why would deep learners benefit more from the self-taught learning framework}?
1033 The key idea is that the lower layers of the predictor compute a hierarchy
1034 of features that can be shared across tasks or across variants of the
1035 input distribution. A theoretical analysis of generalization improvements
1036 due to sharing of intermediate features across tasks already points
1037 towards that explanation~\cite{baxter95a}.
1038 Intermediate features that can be used in different
1039 contexts can be estimated in a way that allows to share statistical
1040 strength. Features extracted through many levels are more likely to
1041 be more abstract and more invariant to some of the factors of variation
1042 in the underlying distribution (as the experiments in~\citet{Goodfellow2009} suggest),
1043 increasing the likelihood that they would be useful for a larger array
1044 of tasks and input conditions.
1045 Therefore, we hypothesize that both depth and unsupervised
1046 pre-training play a part in explaining the advantages observed here, and future
1047 experiments could attempt at teasing apart these factors.
1048 And why would deep learners benefit from the self-taught learning
1049 scenarios even when the number of labeled examples is very large?
1050 We hypothesize that this is related to the hypotheses studied
1051 in~\citet{Erhan+al-2010}. In~\citet{Erhan+al-2010}
1052 it was found that online learning on a huge dataset did not make the
1053 advantage of the deep learning bias vanish, and a similar phenomenon
1054 may be happening here. We hypothesize that unsupervised pre-training
1055 of a deep hierarchy with self-taught learning initializes the
1056 model in the basin of attraction of supervised gradient descent
1057 that corresponds to better generalization. Furthermore, such good
1058 basins of attraction are not discovered by pure supervised learning
1059 (with or without self-taught settings), and more labeled examples
1060 does not allow the model to go from the poorer basins of attraction discovered
1061 by the purely supervised shallow models to the kind of better basins associated
1062 with deep learning and self-taught learning.
1063
1064 A Flash demo of the recognizer (where both the MLP and the SDA can be compared)
1065 can be executed on-line at {\tt http://deep.host22.com}.
1066
1067 \newpage
1068 {
1069 \bibliography{strings,strings-short,strings-shorter,ift6266_ml,aigaion-shorter,specials}
1070 %\bibliographystyle{plainnat}
1071 \bibliographystyle{unsrtnat}
1072 %\bibliographystyle{apalike}
1073 }
1074
1075
1076 \end{document}