nips nips2012 nips2012-91 knowledge-graph by maker-knowledge-mining

91 nips-2012-Deep Neural Networks Segment Neuronal Membranes in Electron Microscopy Images


Source: pdf

Author: Dan Ciresan, Alessandro Giusti, Luca M. Gambardella, Jürgen Schmidhuber

Abstract: We address a central problem of neuroanatomy, namely, the automatic segmentation of neuronal structures depicted in stacks of electron microscopy (EM) images. This is necessary to efficiently map 3D brain structure and connectivity. To segment biological neuron membranes, we use a special type of deep artificial neural network as a pixel classifier. The label of each pixel (membrane or nonmembrane) is predicted from raw pixel values in a square window centered on it. The input layer maps each window pixel to a neuron. It is followed by a succession of convolutional and max-pooling layers which preserve 2D information and extract features with increasing levels of abstraction. The output layer produces a calibrated probability for each class. The classifier is trained by plain gradient descent on a 512 × 512 × 30 stack with known ground truth, and tested on a stack of the same size (ground truth unknown to the authors) by the organizers of the ISBI 2012 EM Segmentation Challenge. Even without problem-specific postprocessing, our approach outperforms competing techniques by a large margin in all three considered metrics, i.e. rand error, warping error and pixel error. For pixel error, our approach is the only one outperforming a second human observer. 1

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 ch Abstract We address a central problem of neuroanatomy, namely, the automatic segmentation of neuronal structures depicted in stacks of electron microscopy (EM) images. [sent-7, score-0.81]

2 To segment biological neuron membranes, we use a special type of deep artificial neural network as a pixel classifier. [sent-9, score-0.499]

3 The label of each pixel (membrane or nonmembrane) is predicted from raw pixel values in a square window centered on it. [sent-10, score-0.688]

4 The input layer maps each window pixel to a neuron. [sent-11, score-0.55]

5 It is followed by a succession of convolutional and max-pooling layers which preserve 2D information and extract features with increasing levels of abstraction. [sent-12, score-0.307]

6 The output layer produces a calibrated probability for each class. [sent-13, score-0.106]

7 The classifier is trained by plain gradient descent on a 512 × 512 × 30 stack with known ground truth, and tested on a stack of the same size (ground truth unknown to the authors) by the organizers of the ISBI 2012 EM Segmentation Challenge. [sent-14, score-0.585]

8 For pixel error, our approach is the only one outperforming a second human observer. [sent-18, score-0.243]

9 The main tool for studying connections at the neuron level is serial-section Transmitted Electron Microscopy (ssTEM), resolving individual neurons and their shapes. [sent-21, score-0.227]

10 After preparation, a sample of neural tissue is typically sectioned into 50-nanometer slices; each slice is then recorded as a 2D grayscale image with a pixel size of about 4 × 4 nanometers (see Figure 1). [sent-22, score-0.459]

11 The visual complexity of the resulting stacks makes them hard to handle. [sent-23, score-0.113]

12 Reliable automated segmentation of neuronal structures in ssTEM stacks so far has been infeasible. [sent-24, score-0.689]

13 ch/˜ciresan 1 Figure 1: Left: the training stack (one slice shown). [sent-28, score-0.256]

14 Right: corresponding ground truth; black lines denote neuron membranes. [sent-29, score-0.182]

15 Our solution is based on a Deep Neural Network (DNN) [12, 13] used as a pixel classifier. [sent-32, score-0.243]

16 The network computes the probability of a pixel being a membrane, using as input the image intensities in a square window centered on the pixel itself. [sent-33, score-0.879]

17 An image is then segmented by classifying all of its pixels. [sent-34, score-0.115]

18 The DNN is trained on a different stack with similar characteristics, in which membranes were manually annotated. [sent-35, score-0.353]

19 DNN are inspired by convolutional neural networks introduced in 1980 [16], improved in the 1990s [25], refined and simplified in the 2000s [5, 33], and brought to their full potential by making them both large and deep [12, 13]. [sent-36, score-0.351]

20 Lately, DNN proved their efficiency on data sets extending from handwritten digits (MNIST) [10, 12], handwritten characters [11] to 3D toys (NORB) [13] and faces [35]. [sent-37, score-0.088]

21 Many other types of learning classifiers have been applied to segmentation of TEM images, where different structures are not easily characterized by intensity differences, and structure boundaries are not correlated with high image gradients, due to noise and many confounding micro-structures. [sent-40, score-0.522]

22 In most binary segmentation problems, classifiers are used to compute one or both of the following probabilities: (a) probability of a pixel belonging to each class; (b) probability of a boundary dividing two adjacent pixels. [sent-41, score-0.594]

23 Some use an additional term to account for the expected geometry of neuron membranes[23]. [sent-43, score-0.107]

24 We compute pixel probabilities only (point (a) above), and directly obtain a segmentation by mild smoothing and thresholding, without using graph cuts. [sent-44, score-0.594]

25 Others have used off-the-shelf random forest classifiers to compute unary terms of neuron membranes [22], or SVMs to compute both unary and binary terms for segmenting mitochondria [28, 27]. [sent-46, score-0.387]

26 The former approach uses haar-like features and texture histograms computed on a small region around the pixel of interest, whereas the latter uses sophisticated rotational [17] and ray [34] features computed on superpixels [3]. [sent-47, score-0.282]

27 Feature selection mirrors the researcher’s expectation of which characteristics of the image are relevant for classification, and has a large impact on classification accuracy. [sent-48, score-0.115]

28 In our approach, we bypass such problems, using raw pixel values as inputs. [sent-49, score-0.291]

29 Due to their convolutional structure, the first layers of the network automatically learn to compute meaningful features during training. [sent-50, score-0.346]

30 The main contribution of the paper is a practical state-of-the-art segmentation method for neuron membranes in ssTEM data, described in Section 2. [sent-51, score-0.602]

31 The contribution is particularly meaningful because our approach does not rely on problem-specific postprocessing: fruitful application to different biomedical segmentation problems is therefore likely. [sent-53, score-0.351]

32 2 Methods For each pixel we consider two possible classes, membrane and non-membrane. [sent-55, score-0.451]

33 1) computes the probability of a pixel p being of the former class, using as input the raw intensity values of a square window centered on p with an edge of w pixels—w being an odd number to enforce symmetry. [sent-57, score-0.445]

34 When a pixel is close to the image border, its window will include pixels outside the image boundaries; such pixels are synthesized by mirroring the pixels in the actual image across the boundary (see Figure 2). [sent-58, score-1.111]

35 The classifier is first trained using the provided training images (Section 2. [sent-59, score-0.14]

36 After training, to segment a test image, the classifier is applied to all of its pixels, thus generating a map of membrane probabilities—i. [sent-61, score-0.208]

37 , a new real-valued image the size of the input image. [sent-63, score-0.115]

38 Binary membrane segmentation is obtained by mild postprocessing techniques discussed in Section 2. [sent-64, score-0.646]

39 1 DNN architecture A DNN [13] consists of a succession of convolutional, max-pooling and fully connected layers. [sent-67, score-0.083]

40 It is a general, hierarchical feature extractor that maps raw pixel intensities of the input image into a feature vector to be classified by several fully connected layers. [sent-68, score-0.579]

41 Each convolutional layer performs a 2D convolution of its input maps with a square filter. [sent-70, score-0.361]

42 The activations of the output maps are obtained by summing the convolutional responses which are passed through a nonlinear activation function. [sent-71, score-0.339]

43 Their outputs are given by the maximum activation over non-overlapping square regions. [sent-73, score-0.139]

44 The DNN also have many more maps per layer, and thus many more connections and weights. [sent-75, score-0.089]

45 After 1 to 4 stages of convolutional and max-pooling layers several fully connected layers further combine the outputs into a 1D feature vector. [sent-76, score-0.46]

46 The output layer is always a fully connected layer with one neuron per class (two in our case). [sent-77, score-0.276]

47 Using a softmax activation function for the last layer guarantees that each neuron’s output activation can be interpreted as the probability of a particular input image belonging to that class. [sent-78, score-0.263]

48 2 Training To train the classifier, we use all available slices of the training stack, i. [sent-80, score-0.106]

49 For each slice, we use all membrane pixels as positive examples (on average, about 50000), and the same amount of pixels randomly sampled (without repetitions) among all nonmembrane pixels. [sent-83, score-0.47]

50 We take advantage of 3 this property, and synthetically augment the training set at the beginning of each epoch by randomly mirroring each training instance, and/or rotating it by ±90◦ . [sent-86, score-0.126]

51 3 Postprocessing of network outputs Because each class is equally represented in the training set but not in the testing data, the network outputs cannot be directly interpreted as probability values; instead, they tend to severely overestimate the membrane probability. [sent-88, score-0.603]

52 To fix this issue, a polynomial function post-processor is applied to the network outputs. [sent-89, score-0.081]

53 To compute its coefficients, a network N is trained on 20 slices of the training volume Ttrain and tested on the remaining 10 slices of the same volume (Ttest , for which ground truth is available). [sent-90, score-0.449]

54 We compare all outputs obtained on Ttest (a total of 2. [sent-91, score-0.097]

55 The same function is then used to calibrate the outputs of all trained networks. [sent-96, score-0.147]

56 After calibration (a grayscale transformation in image processing terms), network outputs are spatially smoothed by a 2-pixel-radius median filter. [sent-97, score-0.336]

57 This results in regularized of membrane boundaries after thresholding. [sent-98, score-0.208]

58 4 Foveation and nonuniform sampling We experimented with two related techniques for improving the network performance by manipulating its input data, namely foveation and nonuniform sampling (see Figure 3). [sent-100, score-0.435]

59 It imposes a spatially-variant blur on the input window pixels, such that full detail is kept in the central section (fovea), while the peripheral parts are defocused by means of a convolution with a disk kernel, to remove fine details. [sent-102, score-0.191]

60 The network, whose task is to classify the center pixel of the window, is then forced to disregard such peripheral fine details, which are most likely irrelevant, while still retaining the general structure of the window (context). [sent-103, score-0.434]

61 First row shows original window (Plain); other rows show effects of foveation (Fov), nonuniform sampling (Nu), and both (Fov+Nu). [sent-105, score-0.391]

62 The leftmost image illustrates how a checkerboard pattern is affected by such transformations. [sent-107, score-0.115]

63 Nonuniform sampling is motivated by the observation that (in this and other applications) larger window sizes w generally result in significant performance improvements. [sent-108, score-0.154]

64 With nonuniform sampling, image pixels are directly mapped to neurons only in the central part of the window; elsewhere, their source pixels are sampled with decreasing resolution as the distance from the window center increases. [sent-110, score-0.72]

65 As a result, the image in the window is deformed in a fisheye-like fashion, and covers a larger area of the input image with fewer neurons. [sent-111, score-0.384]

66 5 Averaging outputs of multiple networks We observed that large networks with different architectures often exhibit significant output differences for many image parts, despite being trained on the same data. [sent-114, score-0.447]

67 It is therefore reasonable to attempt to reduce such variance by averaging the calibrated outputs of several networks with different architectures. [sent-116, score-0.252]

68 The submissions obtained by averaging the outputs of multiple large networks scored significantly better in all metrics than the single networks. [sent-118, score-0.249]

69 For the training stack, a manually annotated ground truth segmentation is provided. [sent-126, score-0.535]

70 One is used as ground truth, the other to evaluate the performance of a second human observer and provide a meaningful comparison for the algorithms’ performance. [sent-128, score-0.122]

71 Warping error: a segmentation metric designed to account for topological disagreements [19]; it accounts for the number of neuron splits and mergers required to obtain the candidate segmentation from ground truth. [sent-130, score-0.884]

72 Pixel error: defined as 1 − Fpixel , where Fpixel represents the F1 score of pixel similarity. [sent-131, score-0.243]

73 The automated system accepts a stack of grayscale images, representing membrane probability values for each pixel; the stack is thresholded using 9 different threshold values, obtaining 9 binary stacks. [sent-132, score-0.64]

74 Pixel error is clearly not a suitable indicator of segmentation quality in this context, and is reported mostly for reference. [sent-134, score-0.351]

75 As the input window size increases, the network depth also increases because we keep the convolutional filter sizes small. [sent-142, score-0.443]

76 All nets are trained for 30 epochs, which leads to a total training time of several days. [sent-145, score-0.146]

77 However, once networks are trained, application to new images is relatively fast: classifying the 8 million pixels comprising the whole testing stack takes 10 to 30 minutes on four GPUs. [sent-146, score-0.392]

78 Table 1: 11-layer architecture for network N4, w = 95. [sent-148, score-0.081]

79 By averaging the outputs of all networks, results improve significantly. [sent-150, score-0.135]

80 The final result for one slice of the test stack is shown in Figure 5. [sent-151, score-0.217]

81 Since our pure pixel classifier method aims at minimizing pixel error, Rand and warping errors are just minimized as a side-effect, but never explicitly accounted for during segmentation. [sent-153, score-0.57]

82 In contrast, some competing segmentation approaches adopt different post-processing techniques directly opti6 Figure 5: Left: slice 16 of the test stack. [sent-154, score-0.451]

83 [18] 48 65 70 76 84 84 89 90 139 153 162 230 434 556 525 646 1602 1124 1134 1512 2641 685 1613 16156 60 83 79 65 134 157 78 100 102 88 109 150 Group mizing the rand error. [sent-169, score-0.12]

84 Preliminary results in this direction are encouraging: the problem-specific postprocessing techniques in [20] and [24], operating on our segmentation, reduce the Rand error to measure to 36·10−3 and 32·10−3 , respectively. [sent-172, score-0.087]

85 7 Our approach outperforms all other approaches in the competition, despite not even being tailored to this particular segmentation task. [sent-175, score-0.351]

86 Instead, the DNN acts as a generic image classifier, using raw pixel intensities as inputs, without ad-hoc post-processing. [sent-176, score-0.449]

87 This opens interesting perspectives on applying similar techniques to other biomedical image segmentation tasks. [sent-177, score-0.466]

88 References [1] Segmentation of neuronal structures in EM stacks challenge - ISBI 2012. [sent-179, score-0.267]

89 An integrated micro- and macroarchitectural analysis of the drosophila brain by computer-assisted serial section electron microscopy. [sent-224, score-0.135]

90 [10] Dan Claudiu Ciresan, Ueli Meier, Luca Maria Gambardella, and J¨ rgen Schmidhuber. [sent-226, score-0.073]

91 Deep, big, simple u neural nets for handwritten digit recognition. [sent-227, score-0.101]

92 Convolutional u neural network committees for handwritten character classification. [sent-230, score-0.125]

93 u Flexible, high performance convolutional neural networks for image classification. [sent-233, score-0.398]

94 Neocognitron: A self-organizing neural network for a mechanism of pattern recognition unaffected by shift in position. [sent-255, score-0.119]

95 Segmentation of EM images of neuronal structures using CellProfiler. [sent-285, score-0.205]

96 Geometrical consistent 3D tracing of neuronal processes in ssTEM data. [sent-296, score-0.098]

97 Neuron geometry extraction by perceptual grouping in sstem images. [sent-303, score-0.12]

98 A fully automated approach to segmentation of irregularly shaped cellular structures in EM images. [sent-335, score-0.478]

99 Evaluation of pooling operations in convolutional u architectures for object recognition. [sent-348, score-0.298]

100 Best practices for convolutional neural networks applied to visual document analysis. [sent-358, score-0.283]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('segmentation', 0.351), ('pixel', 0.243), ('dnn', 0.212), ('isbi', 0.212), ('membrane', 0.208), ('convolutional', 0.208), ('stack', 0.159), ('window', 0.154), ('membranes', 0.144), ('em', 0.129), ('rand', 0.12), ('neurons', 0.12), ('foveation', 0.12), ('sstem', 0.12), ('nonuniform', 0.117), ('image', 0.115), ('stacks', 0.113), ('neuron', 0.107), ('pixels', 0.107), ('ciresan', 0.105), ('neuronal', 0.098), ('outputs', 0.097), ('electron', 0.096), ('lugano', 0.096), ('microscopy', 0.096), ('maps', 0.089), ('postprocessing', 0.087), ('luca', 0.087), ('ttest', 0.085), ('idsia', 0.085), ('warping', 0.084), ('network', 0.081), ('dan', 0.08), ('claudiu', 0.078), ('networks', 0.075), ('ground', 0.075), ('rgen', 0.073), ('ueli', 0.073), ('kaynig', 0.072), ('organizers', 0.072), ('automated', 0.071), ('truth', 0.07), ('gambardella', 0.07), ('deep', 0.068), ('slices', 0.067), ('layer', 0.064), ('achanta', 0.064), ('lucchi', 0.064), ('maria', 0.064), ('er', 0.059), ('classi', 0.059), ('slice', 0.058), ('layers', 0.057), ('nets', 0.057), ('structures', 0.056), ('meier', 0.056), ('pooling', 0.055), ('laptev', 0.052), ('images', 0.051), ('smith', 0.05), ('trained', 0.05), ('burget', 0.048), ('connectome', 0.048), ('fov', 0.048), ('fpixel', 0.048), ('frand', 0.048), ('iftikhar', 0.048), ('mirroring', 0.048), ('mitochondria', 0.048), ('nonmembrane', 0.048), ('stephan', 0.048), ('ttrain', 0.048), ('raw', 0.048), ('observer', 0.047), ('handwritten', 0.044), ('gpu', 0.044), ('unary', 0.044), ('grayscale', 0.043), ('intensities', 0.043), ('sven', 0.042), ('bas', 0.042), ('nu', 0.042), ('succession', 0.042), ('activation', 0.042), ('competing', 0.042), ('calibrated', 0.042), ('connected', 0.041), ('drosophila', 0.039), ('photoreceptor', 0.039), ('tem', 0.039), ('fuchs', 0.039), ('miccai', 0.039), ('rotational', 0.039), ('tomaso', 0.039), ('training', 0.039), ('metrics', 0.039), ('averaging', 0.038), ('recognition', 0.038), ('peripheral', 0.037), ('architectures', 0.035)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 1.0000008 91 nips-2012-Deep Neural Networks Segment Neuronal Membranes in Electron Microscopy Images

Author: Dan Ciresan, Alessandro Giusti, Luca M. Gambardella, Jürgen Schmidhuber

Abstract: We address a central problem of neuroanatomy, namely, the automatic segmentation of neuronal structures depicted in stacks of electron microscopy (EM) images. This is necessary to efficiently map 3D brain structure and connectivity. To segment biological neuron membranes, we use a special type of deep artificial neural network as a pixel classifier. The label of each pixel (membrane or nonmembrane) is predicted from raw pixel values in a square window centered on it. The input layer maps each window pixel to a neuron. It is followed by a succession of convolutional and max-pooling layers which preserve 2D information and extract features with increasing levels of abstraction. The output layer produces a calibrated probability for each class. The classifier is trained by plain gradient descent on a 512 × 512 × 30 stack with known ground truth, and tested on a stack of the same size (ground truth unknown to the authors) by the organizers of the ISBI 2012 EM Segmentation Challenge. Even without problem-specific postprocessing, our approach outperforms competing techniques by a large margin in all three considered metrics, i.e. rand error, warping error and pixel error. For pixel error, our approach is the only one outperforming a second human observer. 1

2 0.28813568 158 nips-2012-ImageNet Classification with Deep Convolutional Neural Networks

Author: Alex Krizhevsky, Ilya Sutskever, Geoffrey E. Hinton

Abstract: We trained a large, deep convolutional neural network to classify the 1.2 million high-resolution images in the ImageNet LSVRC-2010 contest into the 1000 different classes. On the test data, we achieved top-1 and top-5 error rates of 37.5% and 17.0% which is considerably better than the previous state-of-the-art. The neural network, which has 60 million parameters and 650,000 neurons, consists of five convolutional layers, some of which are followed by max-pooling layers, and three fully-connected layers with a final 1000-way softmax. To make training faster, we used non-saturating neurons and a very efficient GPU implementation of the convolution operation. To reduce overfitting in the fully-connected layers we employed a recently-developed regularization method called “dropout” that proved to be very effective. We also entered a variant of this model in the ILSVRC-2012 competition and achieved a winning top-5 test error rate of 15.3%, compared to 26.2% achieved by the second-best entry. 1

3 0.15774728 190 nips-2012-Learning optimal spike-based representations

Author: Ralph Bourdoukan, David Barrett, Sophie Deneve, Christian K. Machens

Abstract: How can neural networks learn to represent information optimally? We answer this question by deriving spiking dynamics and learning dynamics directly from a measure of network performance. We find that a network of integrate-and-fire neurons undergoing Hebbian plasticity can learn an optimal spike-based representation for a linear decoder. The learning rule acts to minimise the membrane potential magnitude, which can be interpreted as a representation error after learning. In this way, learning reduces the representation error and drives the network into a robust, balanced regime. The network becomes balanced because small representation errors correspond to small membrane potentials, which in turn results from a balance of excitation and inhibition. The representation is robust because neurons become self-correcting, only spiking if the representation error exceeds a threshold. Altogether, these results suggest that several observed features of cortical dynamics, such as excitatory-inhibitory balance, integrate-and-fire dynamics and Hebbian plasticity, are signatures of a robust, optimal spike-based code. A central question in neuroscience is to understand how populations of neurons represent information and how they learn to do so. Usually, learning and information representation are treated as two different functions. From the outset, this separation seems like a good idea, as it reduces the problem into two smaller, more manageable chunks. Our approach, however, is to study these together. This allows us to treat learning and information representation as two sides of a single mechanism, operating at two different timescales. Experimental work has given us several clues about the regime in which real networks operate in the brain. Some of the most prominent observations are: (a) high trial-to-trial variability—a neuron responds differently to repeated, identical inputs [1, 2]; (b) asynchronous firing at the network level—spike trains of different neurons are at most very weakly correlated [3, 4, 5]; (c) tight balance of excitation and inhibition—every excitatory input is met by an inhibitory input of equal or greater size [6, 7, 8] and (4) spike-timing-dependent plasticity (STDP)—the strength of synapses change as a function of presynaptic and postsynaptic spike times [9]. Previously, it has been shown that observations (a)–(c) can be understood as signatures of an optimal, spike-based code [10, 11]. The essential idea is to derive spiking dynamics from the assumption that neurons only fire if their spike improves information representation. Information in a network may ∗ Authors contributed equally 1 originate from several possible sources: external sensory input, external neural network input, or alternatively, it may originate within the network itself as a memory, or as a computation. Whatever the source, this initial assumption leads directly to the conclusion that a network of integrate-and-fire neurons can optimally represent a signal while exhibiting properties (a)–(c). A major problem with this framework is that network connectivity must be completely specified a priori, and requires the tuning of N 2 parameters, where N is the number of neurons in the network. Although this is feasible mathematically, it is unclear how a real network could tune itself into this optimal regime. In this work, we solve this problem using a simple synaptic learning rule. The key insight is that the plasticity rule can be derived from the same basic principle as the spiking rule in the earlier work—namely, that any change should improve information representation. Surprisingly, this can be achieved with a local, Hebbian learning rule, where synaptic plasticity is proportional to the product of presynaptic firing rates with post-synaptic membrane potentials. Spiking and synaptic plasticity then work hand in hand towards the same goal: the spiking of a neuron decreases the representation error on a fast time scale, thereby giving rise to the actual population representation; synaptic plasticity decreases the representation error on a slower time scale, thereby improving or maintaining the population representation. For a large set of initial connectivities and spiking dynamics, neural networks are driven into a balanced regime, where excitation and inhibition cancel each other and where spike trains are asynchronous and irregular. Furthermore, the learning rule that we derive reproduces the main features of STDP (property (d) above). In this way, a network can learn to represent information optimally, with synaptic, neural and network dynamics consistent with those observed experimentally. 1 Derivation of the learning rule for a single neuron We begin by deriving a learning rule for a single neuron with an autapse (a self-connection) (Fig. 1A). Our approach is to derive synaptic dynamics for the autapse and spiking dynamics for the neuron such that the neuron learns to optimally represent a time-varying input signal. We will derive a learning rule for networks of neurons later, after we have developed the fundamental concepts for the single neuron case. Our first step is to derive optimal spiking dynamics for the neuron, so that we have a target for our learning rule. We do this by making two simple assumptions [11]. First, we assume that the neuron can provide an estimate or read-out x(t) of a time-dependent signal x(t) by filtering its spike train ˆ o(t) as follows: ˙ x(t) = −ˆ(t) + Γo(t), ˆ x (1) where Γ is a fixed read-out weight, which we will refer to as the neuron’s “output kernel” and the spike train can be written as o(t) = i δ(t − ti ), where {ti } are the spike times. Next, we assume that the neuron only produces a spike if that spike improves the read-out, where we measure the read-out performance through a simple squared-error loss function: 2 L(t) = x(t) − x(t) . ˆ (2) With these two assumptions, we can now derive optimal spiking dynamics. First, we observe that if the neuron produces an additional spike at time t, the read-out increases by Γ, and the loss function becomes L(t|spike) = (x(t) − (x(t) + Γ))2 . This allows us to restate our spiking rule as follows: ˆ the neuron should only produce a spike if L(t|no spike) > L(t|spike), or (x(t) − x(t))2 > (x(t) − ˆ (x(t) + Γ))2 . Now, squaring both sides of this inequality, defining V (t) ≡ Γ(x(t) − x(t)) and ˆ ˆ defining T ≡ Γ2 /2 we find that the neuron should only spike if: V (t) > T. (3) We interpret V (t) to be the membrane potential of the neuron, and we interpret T as the spike threshold. This interpretation allows us to understand the membrane potential functionally: the voltage is proportional to a prediction error—the difference between the read-out x(t) and the actual ˆ signal x(t). A spike is an error reduction mechanism—the neuron only spikes if the error exceeds the spike threshold. This is a greedy minimisation, in that the neuron fires a spike whenever that action decreases L(t) without considering the future impact of that spike. Importantly, the neuron does not require direct access to the loss function L(t). 2 To determine the membrane potential dynamics, we take the derivative of the voltage, which gives ˙ ˙ us V = Γ(x − x). (Here, and in the following, we will drop the time index for notational brevity.) ˙ ˆ ˙ Now, using Eqn. (1) we obtain V = Γx − Γ(−x + Γo) = −Γ(x − x) + Γ(x + x) − Γ2 o, so that: ˙ ˆ ˆ ˙ ˙ V = −V + Γc − Γ2 o, (4) where c = x + x is the neural input. This corresponds exactly to the dynamics of a leaky integrate˙ and-fire neuron with an inhibitory autapse1 of strength Γ2 , and a feedforward connection strength Γ. The dynamics and connectivity guarantee that a neuron spikes at just the right times to optimise the loss function (Fig. 1B). In addition, it is especially robust to noise of different forms, because of its error-correcting nature. If x is constant in time, the voltage will rise up to the threshold T at which point a spike is fired, adding a delta function to the spike train o at time t, thereby producing a read-out x that is closer to x and causing an instantaneous drop in the voltage through the autapse, ˆ by an amount Γ2 = 2T , effectively resetting the voltage to V = −T . We now have a target for learning—we know the connection strength that a neuron must have at the end of learning if it is to represent information optimally, for a linear read-out. We can use this target to derive synaptic dynamics that can learn an optimal representation from experience. Specifically, we consider an integrate-and-fire neuron with some arbitrary autapse strength ω. The dynamics of this neuron are given by ˙ V = −V + Γc − ωo. (5) This neuron will not produce the correct spike train for representing x through a linear read-out (Eqn. (1)) unless ω = Γ2 . Our goal is to derive a dynamical equation for the synapse ω so that the spike train becomes optimal. We do this by quantifying the loss that we are incurring by using the suboptimal strength, and then deriving a learning rule that minimises this loss with respect to ω. The loss function underlying the spiking dynamics determined by Eqn. (5) can be found by reversing the previous membrane potential analysis. First, we integrate the differential equation for V , assuming that ω changes on time scales much slower than the membrane potential. We obtain the following (formal) solution: V = Γx − ω¯, o (6) ˙ where o is determined by o = −¯ + o. The solution to this latter equation is o = h ∗ o, a convolution ¯ ¯ o ¯ of the spike train with the exponential kernel h(τ ) = θ(τ ) exp(−τ ). As such, it is analogous to the instantaneous firing rate of the neuron. Now, using Eqn. (6), and rewriting the read-out as x = Γ¯, we obtain the loss incurred by the ˆ o sub-optimal neuron, L = (x − x)2 = ˆ 1 V 2 + 2(ω − Γ2 )¯ + (ω − Γ2 )2 o2 . o ¯ Γ2 (7) We observe that the last two terms of Eqn. (7) will vanish whenever ω = Γ2 , i.e., when the optimal reset has been found. We can therefore simplify the problem by defining an alternative loss function, 1 2 V , (8) 2 which has the same minimum as the original loss (V = 0 or x = x, compare Eqn. (2)), but yields a ˆ simpler learning algorithm. We can now calculate how changes to ω affect LV : LV = ∂LV ∂V ∂o ¯ =V = −V o − V ω ¯ . (9) ∂ω ∂ω ∂ω We can ignore the last term in this equation (as we will show below). Finally, using simple gradient descent, we obtain a simple Hebbian-like synaptic plasticity rule: τω = − ˙ ∂LV = V o, ¯ ∂ω (10) where τ is the learning time constant. 1 This contribution of the autapse can also be interpreted as the reset of an integrate-and-fire neuron. Later, when we generalise to networks of neurons, we shall employ this interpretation. 3 This synaptic learning rule is capable of learning the synaptic weight ω that minimises the difference between x and x (Fig. 1B). During learning, the synaptic weight changes in proportion to the postˆ synaptic voltage V and the pre-synaptic firing rate o (Fig. 1C). As such, this is a Hebbian learning ¯ rule. Of course, in this single neuron case, the pre-synaptic neuron and post-synaptic neuron are the same neuron. The synaptic weight gradually approaches its optimal value Γ2 . However, it never completely stabilises, because learning never stops as long as neurons are spiking. Instead, the synapse oscillates closely about the optimal value (Fig. 1D). This is also a “greedy” learning rule, similar to the spiking rule, in that it seeks to minimise the error at each instant in time, without regard for the future impact of those changes. To demonstrate that the second term in Eqn. (5) can be neglected we note that the equations for V , o, and ω define a system ¯ of coupled differential equations that can be solved analytically by integrating between spikes. This results in a simple recurrence relation for changes in ω from the ith to the (i + 1)th spike, ωi+1 = ωi + ωi (ωi − 2T ) . τ (T − Γc − ωi ) (11) This iterative equation has a single stable fixed point at ω = 2T = Γ2 , proving that the neuron’s autaptic weight or reset will approach the optimal solution. 2 Learning in a homogeneous network We now generalise our learning rule derivation to a network of N identical, homogeneously connected neurons. This generalisation is reasonably straightforward because many characteristics of the single neuron case are shared by a network of identical neurons. We will return to the more general case of heterogeneously connected neurons in the next section. We begin by deriving optimal spiking dynamics, as in the single neuron case. This provides a target for learning, which we can then use to derive synaptic dynamics. As before, we want our network to produce spikes that optimally represent a variable x for a linear read-out. We assume that the read-out x is provided by summing and filtering the spike trains of all the neurons in the network: ˆ ˙ x = −ˆ + Γo, ˆ x (12) 2 where the row vector Γ = (Γ, . . . , Γ) contains the read-out weights of the neurons and the column vector o = (o1 , . . . , oN ) their spike trains. Here, we have used identical read-out weights for each neuron, because this indirectly leads to homogeneous connectivity, as we will demonstrate. Next, we assume that a neuron only spikes if that spike reduces a loss-function. This spiking rule is similar to the single neuron spiking rule except that this time there is some ambiguity about which neuron should spike to represent a signal. Indeed, there are many different spike patterns that provide exactly the same estimate x. For example, one neuron could fire regularly at a high rate (exactly like ˆ our previous single neuron example) while all others are silent. To avoid this firing rate ambiguity, we use a modified loss function, that selects amongst all equivalent solutions, those with the smallest neural firing rates. We do this by adding a ‘metabolic cost’ term to our loss function, so that high firing rates are penalised: ¯ L = (x − x)2 + µ o 2 , ˆ (13) where µ is a small positive constant that controls the cost-accuracy trade-off, akin to a regularisation parameter. Each neuron in the optimal network will seek to reduce this loss function by firing a spike. Specifically, the ith neuron will spike whenever L(no spike in i) > L(spike in i). This leads to the following spiking rule for the ith neuron: Vi > Ti (14) where Vi ≡ Γ(x − x) − µoi and Ti ≡ Γ2 /2 + µ/2. We can naturally interpret Vi as the membrane ˆ potential of the ith neuron and Ti as the spiking threshold of that neuron. As before, we can now derive membrane potential dynamics: ˙ V = −V + ΓT c − (ΓT Γ + µI)o, 2 (15) The read-out weights must scale as Γ ∼ 1/N so that firing rates are not unrealistically small in large networks. We can see this by calculating the average firing rate N oi /N ≈ x/(ΓN ) ∼ O(N/N ) ∼ O(1). i=1 ¯ 4 where I is the identity matrix and ΓT Γ + µI is the network connectivity. We can interpret the selfconnection terms {Γ2 +µ} as voltage resets that decrease the voltage of any neuron that spikes. This optimal network is equivalent to a network of identical integrate-and-fire neurons with homogeneous inhibitory connectivity. The network has some interesting dynamical properties. The voltages of all the neurons are largely synchronous, all increasing to the spiking threshold at about the same time3 (Fig. 1F). Nonetheless, neural spiking is asynchronous. The first neuron to spike will reset itself by Γ2 + µ, and it will inhibit all the other neurons in the network by Γ2 . This mechanism prevents neurons from spik- x 3 The first neuron to spike will be random if there is some membrane potential noise. V (A) (B) x x ˆ x 10 1 0.1 0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400 1 D 0.5 V V 0 ˆ x V ˆ x (C) 1 0 1 2 0 0.625 25 25.625 (D) start of learning 1 V 50 200.625 400 400.625 1 2.4 O 1.78 ω 1.77 25 neuron$ 0 1 2 !me$ 3 4 25 1 5 V 400.625 !me$ (F) 25 1 2.35 1.05 1.049 400 25.625 !me$ (E) neuron$ 100.625 200 end of learning 1.4 1.35 ω 100 !me$ 1 V 1 O 50.625 0 1 2 !me$ 3 4 5 V !me$ !me$ Figure 1: Learning in a single neuron and a homogeneous network. (A) A single neuron represents an input signal x by producing an output x. (B) During learning, the single neuron output x (solid red ˆ ˆ line, top panel) converges towards the input x (blue). Similarly, for a homogeneous network the output x (dashed red line, top panel) converges towards x. Connectivity also converges towards optimal ˆ connectivity in both the single neuron case (solid black line, middle panel) and the homogeneous net2 2 work case (dashed black line, middle panel), as quantified by D = maxi,j ( Ωij − Ωopt / Ωopt ) ij ij at each point in time. Consequently, the membrane potential reset (bottom panel) converges towards the optimal reset (green line, bottom panel). Spikes are indicated by blue vertical marks, and are produced when the membrane potential reaches threshold (bottom panel). Here, we have rescaled time, as indicated, for clarity. (C) Our learning rule dictates that the autapse ω in our single neuron (bottom panel) changes in proportion to the membrane potential (top panel) and the firing rate (middle panel). (D) At the end of learning, the reset ω fluctuates weakly about the optimal value. (E) For a homogeneous network, neurons spike regularly at the start of learning, as shown in this raster plot. Membrane potentials of different neurons are weakly correlated. (F) At the end of learning, spiking is very irregular and membrane potentials become more synchronous. 5 ing synchronously. The population as a whole acts similarly to the single neuron in our previous example. Each neuron fires regularly, even if a different neuron fires in every integration cycle. The design of this optimal network requires the tuning of N (N − 1) synaptic parameters. How can an arbitrary network of integrate-and-fire neurons learn this optimum? As before, we address this question by using the optimal network as a target for learning. We start with an arbitrarily connected network of integrate-and-fire neurons: ˙ V = −V + ΓT c − Ωo, (16) where Ω is a matrix of connectivity weights, which includes the resets of the individual neurons. Assuming that learning occurs on a slow time scale, we can rewrite this equation as V = ΓT x − Ω¯ . o (17) Now, repeating the arguments from the single neuron derivation, we modify the loss function to obtain an online learning rule. Specifically, we set LV = V 2 /2, and calculate the gradient: ∂LV = ∂Ωij Vk k ∂Vk =− ∂Ωij Vk δki oj − ¯ k Vk Ωkl kl ∂ ol ¯ . ∂Ωij (18) We can simplify this equation considerably by observing that the contribution of the second summation is largely averaged out under a wide variety of realistic conditions4 . Therefore, it can be neglected, and we obtain the following local learning rule: ∂LV ˙ = V i oj . ¯ τ Ωij = − ∂Ωij (19) This is a Hebbian plasticity rule, whereby connectivity changes in proportion to the presynaptic firing rate oj and post-synaptic membrane potential Vi . We assume that the neural thresholds are set ¯ to a constant T and that the neural resets are set to their optimal values −T . In the previous section we demonstrated that these resets can be obtained by a Hebbian plasticity rule (Eqn. (10)). This learning rule minimises the difference between the read-out and the signal, by approaching the optimal recurrent connection strengths for the network (Fig. 1B). As in the single neuron case, learning does not stop, so the connection strengths fluctuate close to their optimal value. During learning, network activity becomes progressively more asynchronous as it progresses towards optimal connectivity (Fig. 1E, F). 3 Learning in the general case Now that we have developed the fundamental concepts underlying our learning rule, we can derive a learning rule for the more general case of a network of N arbitrarily connected leaky integrateand-fire neurons. Our goal is to understand how such networks can learn to optimally represent a ˙ J-dimensional signal x = (x1 , . . . , xJ ), using the read-out equation x = −x + Γo. We consider a network with the following membrane potential dynamics: ˙ V = −V + ΓT c − Ωo, (20) where c is a J-dimensional input. We assume that this input is related to the signal according to ˙ c = x + x. This assumption can be relaxed by treating the input as the control for an arbitrary linear dynamical system, in which case the signal represented by the network is the output of such a computation [11]. However, this further generalisation is beyond the scope of this work. As before, we need to identify the optimal recurrent connectivity so that we have a target for learning. Most generally, the optimal recurrent connectivity is Ωopt ≡ ΓT Γ + µI. The output kernels of the individual neurons, Γi , are given by the rows of Γ, and their spiking thresholds by Ti ≡ Γi 2 /2 + 4 From the definition of the membrane potential we can see that Vk ∼ O(1/N ) because Γ ∼ 1/N . Therefore, the size of the first term in Eqn. (18) is k Vk δki oj = Vi oj ∼ O(1/N ). Therefore, the second term can ¯ ¯ be ignored if kl Vk Ωkl ∂ ol /∂Ωij ¯ O(1/N ). This happens if Ωkl O(1/N 2 ) as at the start of learning. It also happens towards the end of learning if the terms {Ωkl ∂ ol /∂Ωij } are weakly correlated with zero mean, ¯ or if the membrane potentials {Vi } are weakly correlated with zero mean. 6 µ/2. With these connections and thresholds, we find that a network of integrate-and-fire neurons ˆ ¯ will produce spike trains in such a way that the loss function L = x − x 2 + µ o 2 is minimised, ˆ where the read-out is given by x = Γ¯ . We can show this by prescribing a greedy5 spike rule: o a spike is fired by neuron i whenever L(no spike in i) > L(spike in i) [11]. The resulting spike generation rule is Vi > Ti , (21) ˆ where Vi ≡ ΓT (x − x) − µ¯i is interpreted as the membrane potential. o i 5 Despite being greedy, this spiking rule can generate firing rates that are practically identical to the optimal solutions: we checked this numerically in a large ensemble of networks with randomly chosen kernels. (A) x1 … x … 1 1 (B) xJJ x 10 L 10 T T 10 4 6 8 1 Viii V D ˆˆ ˆˆ x11 xJJ x x F 0.5 0 0.4 … … 0.2 0 0 2000 4000 !me   (C) x V V 1 x 10 x 3 ˆ x 8 0 x 10 1 2 3 !me   4 5 4 0 1 4 0 1 8 V (F) Ρ(Δt)   E-­‐I  input   0.4 ˆ x 0 3 0 1 x 10 1.3 0.95 x 10 ˆ x 4 V (E) 1 x 0 end of learning 50 neuron neuron 50 !me   2 0 ˆ x 0 0.5 ISI  Δt     1 2 !me   4 5 4 1.5 1.32 3 2 0.1 Ρ(Δt)   x E-­‐I  input   (D) start of learning 0 2 !me   0 0 0.5 ISI  Δt   1 Figure 2: Learning in a heterogeneous network. (A) A network of neurons represents an input ˆ signal x by producing an output x. (B) During learning, the loss L decreases (top panel). The difference between the connection strengths and the optimal strengths also decreases (middle panel), as 2 2 quantified by the mean difference (solid line), given by D = Ω − Ωopt / Ωopt and the maxi2 2 mum difference (dashed line), given by maxi,j ( Ωij − Ωopt / Ωopt ). The mean population firing ij ij rate (solid line, bottom panel) also converges towards the optimal firing rate (dashed line, bottom panel). (C, E) Before learning, a raster plot of population spiking shows that neurons produce bursts ˆ of spikes (upper panel). The network output x (red line, middle panel) fails to represent x (blue line, middle panel). The excitatory input (red, bottom left panel) and inhibitory input (green, bottom left panel) to a randomly selected neuron is not tightly balanced. Furthermore, a histogram of interspike intervals shows that spiking activity is not Poisson, as indicated by the red line that represents a best-fit exponential distribution. (D, F) At the end of learning, spiking activity is irregular and ˆ Poisson-like, excitatory and inhibitory input is tightly balanced and x matches x. 7 How can we learn this optimal connection matrix? As before, we can derive a learning rule by minimising the cost function LV = V 2 /2. This leads to a Hebbian learning rule with the same form as before: ˙ τ Ωij = Vi oj . ¯ (22) Again, we assume that the neural resets are given by −Ti . Furthermore, in order for this learning rule to work, we must assume that the network input explores all possible directions in the J-dimensional input space (since the kernels Γi can point in any of these directions). The learning performance does not critically depend on how the input variable space is sampled as long as the exploration is extensive. In our simulations, we randomly sample the input c from a Gaussian white noise distribution at every time step for the entire duration of the learning. We find that this learning rule decreases the loss function L, thereby approaching optimal network connectivity and producing optimal firing rates for our linear decoder (Fig. 2B). In this example, we have chosen connectivity that is initially much too weak at the start of learning. Consequently, the initial network behaviour is similar to a collection of unconnected single neurons that ignore each other. Spike trains are not Poisson-like, firing rates are excessively large, excitatory and inhibitory ˆ input is unbalanced and the decoded variable x is highly unreliable (Fig. 2C, E). As a result of learning, the network becomes tightly balanced and the spike trains become asynchronous, irregular and Poisson-like with much lower rates (Fig. 2D, F). However, despite this apparent variability, the population representation is extremely precise, only limited by the the metabolic cost and the discrete nature of a spike. This learnt representation is far more precise than a rate code with independent Poisson spike trains [11]. In particular, shuffling the spike trains in response to identical inputs drastically degrades this precision. 4 Conclusions and Discussion In population coding, large trial-to-trial spike train variability is usually interpreted as noise [2]. We show here that a deterministic network of leaky integrate-and-fire neurons with a simple Hebbian plasticity rule can self-organise into a regime where information is represented far more precisely than in noisy rate codes, while appearing to have noisy Poisson-like spiking dynamics. Our learning rule (Eqn. (22)) has the basic properties of STDP. Specifically, a presynaptic spike occurring immediately before a post-synaptic spike will potentiate a synapse, because membrane potentials are positive immediately before a postsynaptic spike. Furthermore, a presynaptic spike occurring immediately after a post-synaptic spike will depress a synapse, because membrane potentials are always negative immediately after a postsynaptic spike. This is similar in spirit to the STDP rule proposed in [12], but different to classical STDP, which depends on post-synaptic spike times [9]. This learning rule can also be understood as a mechanism for generating a tight balance between excitatory and inhibitory input. We can see this by observing that membrane potentials after learning can be interpreted as representation errors (projected onto the read-out kernels). Therefore, learning acts to minimise the magnitude of membrane potentials. Excitatory and inhibitory input must be balanced if membrane potentials are small, so we can equate balance with optimal information representation. Previous work has shown that the balanced regime produces (quasi-)chaotic network dynamics, thereby accounting for much observed cortical spike train variability [13, 14, 4]. Moreover, the STDP rule has been known to produce a balanced regime [16, 17]. Additionally, recent theoretical studies have suggested that the balanced regime plays an integral role in network computation [15, 13]. In this work, we have connected these mechanisms and functions, to conclude that learning this balance is equivalent to the development of an optimal spike-based population code, and that this learning can be achieved using a simple Hebbian learning rule. Acknowledgements We are grateful for generous funding from the Emmy-Noether grant of the Deutsche Forschungsgemeinschaft (CKM) and the Chaire d’excellence of the Agence National de la Recherche (CKM, DB), as well as a James Mcdonnell Foundation Award (SD) and EU grants BACS FP6-IST-027140, BIND MECT-CT-20095-024831, and ERC FP7-PREDSPIKE (SD). 8 References [1] Tolhurst D, Movshon J, and Dean A (1982) The statistical reliability of signals in single neurons in cat and monkey visual cortex. Vision Res 23: 775–785. [2] Shadlen MN, Newsome WT (1998) The variable discharge of cortical neurons: implications for connectivity, computation, and information coding. J Neurosci 18(10): 3870–3896. [3] Zohary E, Newsome WT (1994) Correlated neuronal discharge rate and its implication for psychophysical performance. Nature 370: 140–143. [4] Renart A, de la Rocha J, Bartho P, Hollender L, Parga N, Reyes A, & Harris, KD (2010) The asynchronous state in cortical circuits. Science 327, 587–590. [5] Ecker AS, Berens P, Keliris GA, Bethge M, Logothetis NK, Tolias AS (2010) Decorrelated neuronal firing in cortical microcircuits. Science 327: 584–587. [6] Okun M, Lampl I (2008) Instantaneous correlation of excitation and inhibition during ongoing and sensory-evoked activities. Nat Neurosci 11, 535–537. [7] Shu Y, Hasenstaub A, McCormick DA (2003) Turning on and off recurrent balanced cortical activity. Nature 423, 288–293. [8] Gentet LJ, Avermann M, Matyas F, Staiger JF, Petersen CCH (2010) Membrane potential dynamics of GABAergic neurons in the barrel cortex of behaving mice. Neuron 65: 422–435. [9] Caporale N, Dan Y (2008) Spike-timing-dependent plasticity: a Hebbian learning rule. Annu Rev Neurosci 31: 25–46. [10] Boerlin M, Deneve S (2011) Spike-based population coding and working memory. PLoS Comput Biol 7, e1001080. [11] Boerlin M, Machens CK, Deneve S (2012) Predictive coding of dynamic variables in balanced spiking networks. under review. [12] Clopath C, B¨ sing L, Vasilaki E, Gerstner W (2010) Connectivity reflects coding: a model of u voltage-based STDP with homeostasis. Nat Neurosci 13(3): 344–352. [13] van Vreeswijk C, Sompolinsky H (1998) Chaotic balanced state in a model of cortical circuits. Neural Comput 10(6): 1321–1371. [14] Brunel N (2000) Dynamics of sparsely connected networks of excitatory and inhibitory neurons. J Comput Neurosci 8, 183–208. [15] Vogels TP, Rajan K, Abbott LF (2005) Neural network dynamics. Annu Rev Neurosci 28: 357–376. [16] Vogels TP, Sprekeler H, Zenke F, Clopath C, Gerstner W. (2011) Inhibitory plasticity balances excitation and inhibition in sensory pathways and memory networks. Science 334(6062):1569– 73. [17] Song S, Miller KD, Abbott LF (2000) Competitive Hebbian learning through spike-timingdependent synaptic plasticity. Nat Neurosci 3(9): 919–926. 9

4 0.15415002 72 nips-2012-Cocktail Party Processing via Structured Prediction

Author: Yuxuan Wang, Deliang Wang

Abstract: While human listeners excel at selectively attending to a conversation in a cocktail party, machine performance is still far inferior by comparison. We show that the cocktail party problem, or the speech separation problem, can be effectively approached via structured prediction. To account for temporal dynamics in speech, we employ conditional random fields (CRFs) to classify speech dominance within each time-frequency unit for a sound mixture. To capture complex, nonlinear relationship between input and output, both state and transition feature functions in CRFs are learned by deep neural networks. The formulation of the problem as classification allows us to directly optimize a measure that is well correlated with human speech intelligibility. The proposed system substantially outperforms existing ones in a variety of noises.

5 0.14170952 8 nips-2012-A Generative Model for Parts-based Object Segmentation

Author: S. Eslami, Christopher Williams

Abstract: The Shape Boltzmann Machine (SBM) [1] has recently been introduced as a stateof-the-art model of foreground/background object shape. We extend the SBM to account for the foreground object’s parts. Our new model, the Multinomial SBM (MSBM), can capture both local and global statistics of part shapes accurately. We combine the MSBM with an appearance model to form a fully generative model of images of objects. Parts-based object segmentations are obtained simply by performing probabilistic inference in the model. We apply the model to two challenging datasets which exhibit significant shape and appearance variability, and find that it obtains results that are comparable to the state-of-the-art. There has been significant focus in computer vision on object recognition and detection e.g. [2], but a strong desire remains to obtain richer descriptions of objects than just their bounding boxes. One such description is a parts-based object segmentation, in which an image is partitioned into multiple sets of pixels, each belonging to either a part of the object of interest, or its background. The significance of parts in computer vision has been recognized since the earliest days of the field (e.g. [3, 4, 5]), and there exists a rich history of work on probabilistic models for parts-based segmentation e.g. [6, 7]. Many such models only consider local neighborhood statistics, however several models have recently been proposed that aim to increase the accuracy of segmentations by also incorporating prior knowledge about the foreground object’s shape [8, 9, 10, 11]. In such cases, probabilistic techniques often mainly differ in how accurately they represent and learn about the variability exhibited by the shapes of the object’s parts. Accurate models of the shapes and appearances of parts can be necessary to perform inference in datasets that exhibit large amounts of variability. In general, the stronger the models of these two components, the more performance is improved. A generative model has the added benefit of being able to generate samples, which allows us to visually inspect the quality of its understanding of the data and the problem. Recently, a generative probabilistic model known as the Shape Boltzmann Machine (SBM) has been used to model binary object shapes [1]. The SBM has been shown to constitute the state-of-the-art and it possesses several highly desirable characteristics: samples from the model look realistic, and it generalizes to generate samples that differ from the limited number of examples it is trained on. The main contributions of this paper are as follows: 1) In order to account for object parts we extend the SBM to use multinomial visible units instead of binary ones, resulting in the Multinomial Shape Boltzmann Machine (MSBM), and we demonstrate that the MSBM constitutes a strong model of parts-based object shape. 2) We combine the MSBM with an appearance model to form a fully generative model of images of objects (see Fig. 1). We show how parts-based object segmentations can be obtained simply by performing probabilistic inference in the model. We apply our model to two challenging datasets and find that in addition to being principled and fully generative, the model’s performance is comparable to the state-of-the-art. 1 Train labels Train images Test image Appearance model Joint Model Shape model Parsing Figure 1: Overview. Using annotated images separate models of shape and appearance are trained. Given an unseen test image, its parsing is obtained via inference in the proposed joint model. In Secs. 1 and 2 we present the model and propose efficient inference and learning schemes. In Sec. 3 we compare and contrast the resulting joint model with existing work in the literature. We describe our experimental results in Sec. 4 and conclude with a discussion in Sec. 5. 1 Model We consider datasets of cropped images of an object class. We assume that the images are constructed through some combination of a fixed number of parts. Given a dataset D = {Xd }, d = 1...n of such images X, each consisting of P pixels {xi }, i = 1...P , we wish to infer a segmentation S for the image. S consists of a labeling si for every pixel, where si is a 1-of-(L+1) encoded variable, and L is the fixed number of parts that combine to generate the foreground. In other words, si = (sli ), P l = 0...L, sli 2 {0, 1} and l sli = 1. Note that the background is also treated as a ‘part’ (l = 0). Accurate inference of S is driven by models for 1) part shapes and 2) part appearances. Part shapes: Several types of models can be used to define probabilistic distributions over segmentations S. The simplest approach is to model each pixel si independently with categorical variables whose parameters are specified by the object’s mean shape (Fig. 2(a)). Markov Random Fields (MRFs, Fig. 2(b)) additionally model interactions between nearby pixels using pairwise potential functions that efficiently capture local properties of images like smoothness and continuity. Restricted Boltzmann Machines (RBMs) and their multi-layered counterparts Deep Boltzmann Machines (DBMs, Fig. 2(c)) make heavy use of hidden variables to efficiently define higher-order potentials that take into account the configuration of larger groups of image pixels. The introduction of such hidden variables provides a way to efficiently capture complex, global properties of image pixels. RBMs and DBMs are powerful generative models, but they also have many parameters. Segmented images, however, are expensive to obtain and datasets are typically small (hundreds of examples). In order to learn a model that accurately captures the properties of part shapes we use DBMs but also impose carefully chosen connectivity and capacity constraints, following the structure of the Shape Boltzmann Machine (SBM) [1]. We further extend the model to account for multi-part shapes to obtain the Multinomial Shape Boltzmann Machine (MSBM). The MSBM has two layers of latent variables: h1 and h2 (collectively H = {h1 , h2 }), and defines a P Boltzmann distribution over segmentations p(S) = h1 ,h2 exp{ E(S, h1 , h2 |✓s )}/Z(✓s ) where X X X X X 1 2 E(S, h1 , h2 |✓s ) = bli sli + wlij sli h1 + c 1 h1 + wjk h1 h2 + c2 h2 , (1) j j j j k k k i,l j i,j,l j,k k where j and k range over the first and second layer hidden variables, and ✓s = {W 1 , W 2 , b, c1 , c2 } are the shape model parameters. In the first layer, local receptive fields are enforced by connecting each hidden unit in h1 only to a subset of the visible units, corresponding to one of four patches, as shown in Fig. 2(d,e). Each patch overlaps its neighbor by b pixels, which allows boundary continuity to be learned at the lowest layer. We share weights between the four sets of first-layer hidden units and patches, and purposely restrict the number of units in h2 . These modifications significantly reduce the number of parameters whilst taking into account an important property of shapes, namely that the strongest dependencies between pixels are typically local. 2 h2 1 1 h S S (a) Mean h S (b) MRF h2 h2 h1 S S (c) DBM b (d) SBM (e) 2D SBM Figure 2: Models of shape. Object shape is modeled with undirected graphical models. (a) 1D slice of a mean model. (b) Markov Random Field in 1D. (c) Deep Boltzmann Machine in 1D. (d) 1D slice of a Shape Boltzmann Machine. (e) Shape Boltzmann Machine in 2D. In all models latent units h are binary and visible units S are multinomial random variables. Based on Fig. 2 of [1]. k=1 k=2 k=3 k=1 k=2 k=3 k=1 k=2 k=3 ⇡ l=0 l=1 l=2 Figure 3: A model of appearances. Left: An exemplar dataset. Here we assume one background (l = 0) and two foreground (l = 1, non-body; l = 2, body) parts. Right: The corresponding appearance model. In this example, L = 2, K = 3 and W = 6. Best viewed in color. Part appearances: Pixels in a given image are assumed to have been generated by W fixed Gaussians in RGB space. During pre-training, the means {µw } and covariances {⌃w } of these Gaussians are extracted by training a mixture model with W components on every pixel in the dataset, ignoring image and part structure. It is also assumed that each of the L parts can have different appearances in different images, and that these appearances can be clustered into K classes. The classes differ in how likely they are to use each of the W components when ‘coloring in’ the part. The generative process is as follows. For part l in an image, one of the K classes is chosen (represented by a 1-of-K indicator variable al ). Given al , the probability distribution defined on pixels associated with part l is given by a Gaussian mixture model with means {µw } and covariances {⌃w } and mixing proportions { lkw }. The prior on A = {al } specifies the probability ⇡lk of appearance class k being chosen for part l. Therefore appearance parameters ✓a = {⇡lk , lkw } (see Fig. 3) and: a p(xi |A, si , ✓ ) = p(A|✓a ) = Y l Y l a sli p(xi |al , ✓ ) p(al |✓a ) = = Y Y X YY l l k w lkw N (xi |µw , ⌃w ) !alk !sli (⇡lk )alk . , (2) (3) k Combining shapes and appearances: To summarize, the latent variables for X are A, S, H, and the model’s active parameters ✓ include shape parameters ✓s and appearance parameters ✓a , so that p(X, A, S, H|✓) = Y 1 p(A|✓a )p(S, H|✓s ) p(xi |A, si , ✓a ) , Z( ) i (4) where the parameter adjusts the relative contributions of the shape and appearance components. See Fig. 4 for an illustration of the complete graphical model. During learning, we find the values of ✓ that maximize the likelihood of the training data D, and segmentation is performed on a previously-unseen image by querying the marginal distribution p(S|Xtest , ✓). Note that Z( ) is constant throughout the execution of the algorithms. We set via trial and error in our experiments. 3 n H ✓a si al H xi L+1 ✓s S X A P Figure 4: A model of shape and appearance. Left: The joint model. Pixels xi are modeled via appearance variables al . The model’s belief about each layer’s shape is captured by shape variables H. Segmentation variables si assign each pixel to a layer. Right: Schematic for an image X. 2 Inference and learning Inference: We approximate p(A, S, H|X, ✓) by drawing samples of A, S and H using block-Gibbs Markov Chain Monte Carlo (MCMC). The desired distribution p(S|X, ✓) can then be obtained by considering only the samples for S (see Algorithm 1). In order to sample p(A|S, H, X, ✓) we consider the conditional distribution of appearance class k being chosen for part l which is given by: Q P ·s ⇡lk i ( w lkw N (xi |µw , ⌃w )) li h Q P i. p(alk = 1|S, X, ✓) = P (5) K ·sli r=1 ⇡lr i( w lrw N (xi |µw , ⌃w )) Since the MSBM only has edges between each pair of adjacent layers, all hidden units within a layer are conditionally independent given the units in the other two layers. This property can be exploited to make inference in the shape model exact and efficient. The conditional probabilities are: X X 1 2 p(h1 = 1|s, h2 , ✓) = ( wlij sli + wjk h2 + c1 ), (6) j k j i,l p(h2 k 1 = 1|h , ✓) = ( X k 2 wjk h1 j + c2 ), j (7) j where (y) = 1/(1 + exp( y)) is the sigmoid function. To sample from p(H|S, X, ✓) we iterate between Eqns. 6 and 7 multiple times and keep only the final values of h1 and h2 . Finally, we draw samples for the pixels in p(S|A, H, X, ✓) independently: P 1 exp( j wlij h1 + bli ) p(xi |A, sli = 1, ✓) j p(sli = 1|A, H, X, ✓) = PL . (8) P 1 1 m=1 exp( j wmij hj + bmi ) p(xi |A, smi = 1, ✓) Seeding: Since the latent-space is extremely high-dimensional, in practice we find it helpful to run several inference chains, each initializing S(1) to a different value. The ‘best’ inference is retained and the others are discarded. The computation of the likelihood p(X|✓) of image X is intractable, so we approximate the quality of each inference using a scoring function: 1X Score(X|✓) = p(X, A(t) , S(t) , H(t) |✓), (9) T t where {A(t) , S(t) , H(t) }, t = 1...T are the samples obtained from the posterior p(A, S, H|X, ✓). If the samples were drawn from the prior p(A, S, H|✓) the scoring function would be an unbiased estimator of p(X|✓), but would be wildly inaccurate due to the high probability of missing the important regions of latent space (see e.g. [12, p. 107-109] for further discussion of this issue). Learning: Learning of the model involves maximizing the log likelihood log p(D|✓a , ✓s ) of the training dataset D with respect to the model parameters ✓a and ✓s . Since training is partially supervised, in that for each image X its corresponding segmentation S is also given, we can learn the parameters of the shape and appearance components separately. For appearances, the learning of the mixing coefficients and the histogram parameters decomposes into standard mixture updates independently for each part. For shapes, we follow the standard deep 4 Algorithm 1 MCMC inference algorithm. 1: procedure I NFER(X, ✓) 2: Initialize S(1) , H(1) 3: for t 2 : chain length do 4: A(t) ⇠ p(A|S(t 1) , H(t 1) , X, ✓) 5: S(t) ⇠ p(S|A(t) , H(t 1) , X, ✓) 6: H(t) ⇠ p(H|S(t) , ✓) 7: return {S(t) }t=burnin:chain length learning literature closely [13, 1]. In the pre-training phase we greedily train the model bottom up, one layer at a time. We begin by training an RBM on the observed data using stochastic maximum likelihood learning (SML; also referred to as ‘persistent CD’; [14, 13]). Once this RBM is trained, we infer the conditional mean of the hidden units for each training image. The resulting vectors then serve as the training data for a second RBM which is again trained using SML. We use the parameters of these two RBMs to initialize the parameters of the full MSBM model. In the second phase we perform approximate stochastic gradient ascent in the likelihood of the full model to finetune the parameters in an EM-like scheme as described in [13]. 3 Related work Existing probabilistic models of images can be categorized by the amount of variability they expect to encounter in the data and by how they model this variability. A significant portion of the literature models images using only two parts: a foreground object and its background e.g. [15, 16, 17, 18, 19]. Models that account for the parts within the foreground object mainly differ in how accurately they learn about and represent the variability of the shapes of the object’s parts. In Probabilistic Index Maps (PIMs) [8] a mean partitioning is learned, and the deformable PIM [9] additionally allows for local deformations of this mean partitioning. Stel Component Analysis [10] accounts for larger amounts of shape variability by learning a number of different template means for the object that are blended together on a pixel-by-pixel basis. Factored Shapes and Appearances [11] models global properties of shape using a factor analysis-like model, and ‘masked’ RBMs have been used to model more local properties of shape [20]. However, none of these models constitute a strong model of shape in terms of realism of samples and generalization capabilities [1]. We demonstrate in Sec. 4 that, like the SBM, the MSBM does in fact possess these properties. The closest works to ours in terms of ability to deal with datasets that exhibit significant variability in both shape and appearance are the works of Bo and Fowlkes [21] and Thomas et al. [22]. Bo and Fowlkes [21] present an algorithm for pedestrian segmentation that models the shapes of the parts using several template means. The different parts are composed using hand coded geometric constraints, which means that the model cannot be automatically extended to other application domains. The Implicit Shape Model (ISM) used in [22] is reliant on interest point detectors and defines distributions over segmentations only in the posterior, and therefore is not fully generative. The model presented here is entirely learned from data and fully generative, therefore it can be applied to new datasets and diagnosed with relative ease. Due to its modular structure, we also expect it to rapidly absorb future developments in shape and appearance models. 4 Experiments Penn-Fudan pedestrians: The first dataset that we considered is Penn-Fudan pedestrians [23], consisting of 169 images of pedestrians (Fig. 6(a)). The images are annotated with ground-truth segmentations for L = 7 different parts (hair, face, upper and lower clothes, shoes, legs, arms; Fig. 6(d)). We compare the performance of the model with the algorithm of Bo and Fowlkes [21]. For the shape component, we trained an MSBM on the 684 images of a labeled version of the HumanEva dataset [24] (at 48 ⇥ 24 pixels; also flipped horizontally) with overlap b = 4, and 400 and 50 hidden units in the first and second layers respectively. Each layer was pre-trained for 3000 epochs (iterations). After pre-training, joint training was performed for 1000 epochs. 5 (c) Completion (a) Sampling (b) Diffs ! ! ! Figure 5: Learned shape model. (a) A chain of samples (1000 samples between frames). The apparent ‘blurriness’ of samples is not due to averaging or resizing. We display the probability of each pixel belonging to different parts. If, for example, there is a 50-50 chance that a pixel belongs to the red or blue parts, we display that pixel in purple. (b) Differences between the samples and their most similar counterparts in the training dataset. (c) Completion of occlusions (pink). To assess the realism and generalization characteristics of the learned MSBM we sample from it. In Fig. 5(a) we show a chain of unconstrained samples from an MSBM generated via block-Gibbs MCMC (1000 samples between frames). The model captures highly non-linear correlations in the data whilst preserving the object’s details (e.g. face and arms). To demonstrate that the model has not simply memorized the training data, in Fig. 5(b) we show the difference between the sampled shapes in Fig. 5(a) and their closest images in the training set (based on per-pixel label agreement). We see that the model generalizes in non-trivial ways to generate realistic shapes that it had not encountered during training. In Fig. 5(c) we show how the MSBM completes rectangular occlusions. The samples highlight the variability in possible completions captured by the model. Note how, e.g. the length of the person’s trousers on one leg affects the model’s predictions for the other, demonstrating the model’s knowledge about long-range dependencies. An interactive M ATLAB GUI for sampling from this MSBM has been included in the supplementary material. The Penn-Fudan dataset (at 200 ⇥ 100 pixels) was then split into 10 train/test cross-validation splits without replacement. We used the training images in each split to train the appearance component with a vocabulary of size W = 50 and K = 100 mixture components1 . We additionally constrained the model by sharing the appearance models for the arms and legs with that of the face. We assess the quality of the appearance model by performing the following experiment: for each test image, we used the scoring function described in Eq. 9 to evaluate a number of different proposal segmentations for that image. We considered 10 randomly chosen segmentations from the training dataset as well as the ground-truth segmentation for the test image, and found that the appearance model correctly assigns the highest score to the ground-truth 95% of the time. During inference, the shape and appearance models (which are defined on images of different sizes), were combined at 200 ⇥ 100 pixels via M ATLAB’s imresize function, and we set = 0.8 (Eq. 8) via trial and error. Inference chains were seeded at 100 exemplar segmentations from the HumanEva dataset (obtained using the K-medoids algorithm with K = 100), and were run for 20 Gibbs iterations each (with 5 iterations of Eqs. 6 and 7 per Gibbs iteration). Our unoptimized M ATLAB implementation completed inference for each chain in around 7 seconds. We compute the conditional probability of each pixel belonging to different parts given the last set of samples obtained from the highest scoring chain, assign each pixel independently to the most likely part at that pixel, and report the percentage of correctly labeled pixels (see Table 1). We find that accuracy can be improved using superpixels (SP) computed on X (pixels within a superpixel are all assigned the most common label within it; as with [21] we use gPb-OWT-UCM [25]). We also report the accuracy obtained, had the top scoring seed segmentation been returned as the final segmentation for each image. Here the quality of the seed is determined solely by the appearance model. We observe that the model has comparable performance to the state-of-the-art but pedestrianspecific algorithm of [21], and that inference in the model significantly improves the accuracy of the segmentations over the baseline (top seed+SP). Qualitative results can be seen in Fig. 6(c). 1 We obtained the best quantitative results with these settings. The appearances exhibited by the parts in the dataset are highly varied, and the complexity of the appearance model reflects this fact. 6 Table 1: Penn-Fudan pedestrians. We report the percentage of correctly labeled pixels. The final column is an average of the background, upper and lower body scores (as reported in [21]). FG BG Upper Body Lower Body Head Average Bo and Fowlkes [21] 73.3% 81.1% 73.6% 71.6% 51.8% 69.5% MSBM MSBM + SP 70.7% 71.6% 72.8% 73.8% 68.6% 69.9% 66.7% 68.5% 53.0% 54.1% 65.3% 66.6% Top seed Top seed + SP 59.0% 61.6% 61.8% 67.3% 56.8% 60.8% 49.8% 54.1% 45.5% 43.5% 53.5% 56.4% Table 2: ETHZ cars. We report the percentage of pixels belonging to each part that are labeled correctly. The final column is an average weighted by the frequency of occurrence of each label. BG Body Wheel Window Bumper License Light Average ISM [22] 93.2% 72.2% 63.6% 80.5% 73.8% 56.2% 34.8% 86.8% MSBM 94.6% 72.7% 36.8% 74.4% 64.9% 17.9% 19.9% 86.0% Top seed 92.2% 68.4% 28.3% 63.8% 45.4% 11.2% 15.1% 81.8% ETHZ cars: The second dataset that we considered is the ETHZ labeled cars dataset [22], which itself is a subset of the LabelMe dataset [23], consisting of 139 images of cars, all in the same semiprofile view (Fig. 7(a)). The images are annotated with ground-truth segmentations for L = 6 parts (body, wheel, window, bumper, license plate, headlight; Fig. 7(d)). We compare the performance of the model with the ISM of Thomas et al. [22], who also report their results on this dataset. The dataset was split into 10 train/test cross-validation splits without replacement. We used the training images in each split to train both the shape and appearance components. For the shape component, we trained an MSBM at 50 ⇥ 50 pixels with overlap b = 4, and 2000 and 100 hidden units in the first and second layers respectively. Each layer was pre-trained for 3000 epochs and joint training was performed for 1000 epochs. The appearance model was trained with a vocabulary of size W = 50 and K = 100 mixture components and we set = 0.7. Inference chains were seeded at 50 exemplar segmentations (obtained using K-medoids). We find that the use of superpixels does not help with this dataset (due to the poor quality of superpixels obtained for these images). Qualitative and quantitative results that show the performance of model to be comparable to the state-of-the-art ISM can be seen in Fig. 7(c) and Table 2. We believe the discrepancy in accuracy between the MSBM and ISM on the ‘license’ and ‘light’ labels to mainly be due to ISM’s use of interest-points, as they are able to locate such fine structures accurately. By incorporating better models of part appearance into the generative model, we expect to see this discrepancy decrease. 5 Conclusions and future work In this paper we have shown how the SBM can be extended to obtain the MSBM, and presented a principled probabilistic model of images of objects that exploits the MSBM as its model for part shapes. We demonstrated how object segmentations can be obtained simply by performing MCMC inference in the model. The model can also be treated as a probabilistic evaluator of segmentations: given a proposal segmentation it can be used to estimate its likelihood. This leads us to believe that the combination of a generative model such as ours, with a discriminative, bottom-up segmentation algorithm could be highly effective. We are currently investigating how textured appearance models, which take into account the spatial structure of pixels, affect the learning and inference algorithms and the performance of the model. Acknowledgments Thanks to Charless Fowlkes and Vittorio Ferrari for access to datasets, and to Pushmeet Kohli and John Winn for valuable discussions. AE has received funding from the Carnegie Trust, the SORSAS scheme, and the IST Programme under the PASCAL2 Network of Excellence (IST-2007-216886). 7 (a) Test (c) MSBM (b) Bo and Fowlkes (d) Ground truth Background Hair Face Upper Shoes Legs Lower Arms (d) Ground truth (c) MSBM (b) Thomas et al. (a) Test Figure 6: Penn-Fudan pedestrians. (a) Test images. (b) Results reported by Bo and Fowlkes [21]. (c) Output of the joint model. (d) Ground-truth images. Images shown are those selected by [21]. Background Body Wheel Window Bumper License Headlight Figure 7: ETHZ cars. (a) Test images. (b) Results reported by Thomas et al. [22]. (c) Output of the joint model. (d) Ground-truth images. Images shown are those selected by [22]. 8 References [1] S. M. Ali Eslami, Nicolas Heess, and John Winn. The Shape Boltzmann Machine: a Strong Model of Object Shape. In IEEE CVPR, 2012. [2] Mark Everingham, Luc Van Gool, Christopher K. I. Williams, John Winn, and Andrew Zisserman. The PASCAL Visual Object Classes (VOC) Challenge. International Journal of Computer Vision, 88:303–338, 2010. [3] Martin Fischler and Robert Elschlager. The Representation and Matching of Pictorial Structures. IEEE Transactions on Computers, 22(1):67–92, 1973. [4] David Marr. Vision: A Computational Investigation into the Human Representation and Processing of Visual Information. Freeman, 1982. [5] Irving Biederman. Recognition-by-components: A theory of human image understanding. Psychological Review, 94:115–147, 1987. [6] Ashish Kapoor and John Winn. Located Hidden Random Fields: Learning Discriminative Parts for Object Detection. In ECCV, pages 302–315, 2006. [7] John Winn and Jamie Shotton. The Layout Consistent Random Field for Recognizing and Segmenting Partially Occluded Objects. In IEEE CVPR, pages 37–44, 2006. [8] Nebojsa Jojic and Yaron Caspi. Capturing Image Structure with Probabilistic Index Maps. In IEEE CVPR, pages 212–219, 2004. [9] John Winn and Nebojsa Jojic. LOCUS: Learning object classes with unsupervised segmentation. In ICCV, pages 756–763, 2005. [10] Nebojsa Jojic, Alessandro Perina, Marco Cristani, Vittorio Murino, and Brendan Frey. Stel component analysis. In IEEE CVPR, pages 2044–2051, 2009. [11] S. M. Ali Eslami and Christopher K. I. Williams. Factored Shapes and Appearances for Partsbased Object Understanding. In BMVC, pages 18.1–18.12, 2011. [12] Nicolas Heess. Learning generative models of mid-level structure in natural images. PhD thesis, University of Edinburgh, 2011. [13] Ruslan Salakhutdinov and Geoffrey Hinton. Deep Boltzmann Machines. In AISTATS, volume 5, pages 448–455, 2009. [14] Tijmen Tieleman. Training restricted Boltzmann machines using approximations to the likelihood gradient. In ICML, pages 1064–1071, 2008. [15] Carsten Rother, Vladimir Kolmogorov, and Andrew Blake. “GrabCut”: interactive foreground extraction using iterated graph cuts. ACM SIGGRAPH, 23:309–314, 2004. [16] Eran Borenstein, Eitan Sharon, and Shimon Ullman. Combining Top-Down and Bottom-Up Segmentation. In CVPR Workshop on Perceptual Organization in Computer Vision, 2004. [17] Himanshu Arora, Nicolas Loeff, David Forsyth, and Narendra Ahuja. Unsupervised Segmentation of Objects using Efficient Learning. IEEE CVPR, pages 1–7, 2007. [18] Bogdan Alexe, Thomas Deselaers, and Vittorio Ferrari. ClassCut for unsupervised class segmentation. In ECCV, pages 380–393, 2010. [19] Nicolas Heess, Nicolas Le Roux, and John Winn. Weakly Supervised Learning of ForegroundBackground Segmentation using Masked RBMs. In ICANN, 2011. [20] Nicolas Le Roux, Nicolas Heess, Jamie Shotton, and John Winn. Learning a Generative Model of Images by Factoring Appearance and Shape. Neural Computation, 23(3):593–650, 2011. [21] Yihang Bo and Charless Fowlkes. Shape-based Pedestrian Parsing. In IEEE CVPR, 2011. [22] Alexander Thomas, Vittorio Ferrari, Bastian Leibe, Tinne Tuytelaars, and Luc Van Gool. Using Recognition and Annotation to Guide a Robot’s Attention. IJRR, 28(8):976–998, 2009. [23] Bryan Russell, Antonio Torralba, Kevin Murphy, and William Freeman. LabelMe: A Database and Tool for Image Annotation. International Journal of Computer Vision, 77:157–173, 2008. [24] Leonid Sigal, Alexandru Balan, and Michael Black. HumanEva. International Journal of Computer Vision, 87(1-2):4–27, 2010. [25] Pablo Arbelaez, Michael Maire, Charless C. Fowlkes, and Jitendra Malik. From Contours to Regions: An Empirical Evaluation. In IEEE CVPR, 2009. 9

6 0.13842641 303 nips-2012-Searching for objects driven by context

7 0.11605328 195 nips-2012-Learning visual motion in recurrent neural networks

8 0.11243061 87 nips-2012-Convolutional-Recursive Deep Learning for 3D Object Classification

9 0.10404886 92 nips-2012-Deep Representations and Codes for Image Auto-Annotation

10 0.10191248 193 nips-2012-Learning to Align from Scratch

11 0.10188857 197 nips-2012-Learning with Recursive Perceptual Representations

12 0.095380515 230 nips-2012-Multiple Choice Learning: Learning to Produce Multiple Structured Outputs

13 0.095356554 90 nips-2012-Deep Learning of Invariant Features via Simulated Fixations in Video

14 0.088550612 93 nips-2012-Deep Spatio-Temporal Architectures and Learning for Protein Structure Prediction

15 0.087048523 137 nips-2012-From Deformations to Parts: Motion-based Segmentation of 3D Objects

16 0.082284011 159 nips-2012-Image Denoising and Inpainting with Deep Neural Networks

17 0.079989716 202 nips-2012-Locally Uniform Comparison Image Descriptor

18 0.077608325 200 nips-2012-Local Supervised Learning through Space Partitioning

19 0.077512644 62 nips-2012-Burn-in, bias, and the rationality of anchoring

20 0.077512644 116 nips-2012-Emergence of Object-Selective Features in Unsupervised Feature Learning


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, 0.18), (1, 0.064), (2, -0.256), (3, 0.075), (4, 0.079), (5, 0.056), (6, 0.001), (7, -0.028), (8, -0.034), (9, 0.022), (10, -0.009), (11, 0.038), (12, 0.031), (13, 0.102), (14, -0.026), (15, -0.029), (16, -0.047), (17, -0.018), (18, 0.069), (19, -0.119), (20, -0.066), (21, -0.146), (22, -0.001), (23, 0.11), (24, 0.028), (25, -0.036), (26, -0.038), (27, 0.027), (28, -0.025), (29, 0.008), (30, 0.033), (31, 0.018), (32, -0.016), (33, 0.007), (34, 0.105), (35, 0.031), (36, -0.015), (37, 0.078), (38, 0.005), (39, 0.013), (40, -0.05), (41, -0.199), (42, -0.137), (43, -0.027), (44, 0.09), (45, 0.118), (46, 0.051), (47, -0.065), (48, 0.076), (49, -0.03)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.95709568 91 nips-2012-Deep Neural Networks Segment Neuronal Membranes in Electron Microscopy Images

Author: Dan Ciresan, Alessandro Giusti, Luca M. Gambardella, Jürgen Schmidhuber

Abstract: We address a central problem of neuroanatomy, namely, the automatic segmentation of neuronal structures depicted in stacks of electron microscopy (EM) images. This is necessary to efficiently map 3D brain structure and connectivity. To segment biological neuron membranes, we use a special type of deep artificial neural network as a pixel classifier. The label of each pixel (membrane or nonmembrane) is predicted from raw pixel values in a square window centered on it. The input layer maps each window pixel to a neuron. It is followed by a succession of convolutional and max-pooling layers which preserve 2D information and extract features with increasing levels of abstraction. The output layer produces a calibrated probability for each class. The classifier is trained by plain gradient descent on a 512 × 512 × 30 stack with known ground truth, and tested on a stack of the same size (ground truth unknown to the authors) by the organizers of the ISBI 2012 EM Segmentation Challenge. Even without problem-specific postprocessing, our approach outperforms competing techniques by a large margin in all three considered metrics, i.e. rand error, warping error and pixel error. For pixel error, our approach is the only one outperforming a second human observer. 1

2 0.82290912 158 nips-2012-ImageNet Classification with Deep Convolutional Neural Networks

Author: Alex Krizhevsky, Ilya Sutskever, Geoffrey E. Hinton

Abstract: We trained a large, deep convolutional neural network to classify the 1.2 million high-resolution images in the ImageNet LSVRC-2010 contest into the 1000 different classes. On the test data, we achieved top-1 and top-5 error rates of 37.5% and 17.0% which is considerably better than the previous state-of-the-art. The neural network, which has 60 million parameters and 650,000 neurons, consists of five convolutional layers, some of which are followed by max-pooling layers, and three fully-connected layers with a final 1000-way softmax. To make training faster, we used non-saturating neurons and a very efficient GPU implementation of the convolution operation. To reduce overfitting in the fully-connected layers we employed a recently-developed regularization method called “dropout” that proved to be very effective. We also entered a variant of this model in the ILSVRC-2012 competition and achieved a winning top-5 test error rate of 15.3%, compared to 26.2% achieved by the second-best entry. 1

3 0.71383572 93 nips-2012-Deep Spatio-Temporal Architectures and Learning for Protein Structure Prediction

Author: Pietro D. Lena, Ken Nagata, Pierre F. Baldi

Abstract: Residue-residue contact prediction is a fundamental problem in protein structure prediction. Hower, despite considerable research efforts, contact prediction methods are still largely unreliable. Here we introduce a novel deep machine-learning architecture which consists of a multidimensional stack of learning modules. For contact prediction, the idea is implemented as a three-dimensional stack of Neural Networks NNk , where i and j index the spatial coordinates of the contact ij map and k indexes “time”. The temporal dimension is introduced to capture the fact that protein folding is not an instantaneous process, but rather a progressive refinement. Networks at level k in the stack can be trained in supervised fashion to refine the predictions produced by the previous level, hence addressing the problem of vanishing gradients, typical of deep architectures. Increased accuracy and generalization capabilities of this approach are established by rigorous comparison with other classical machine learning approaches for contact prediction. The deep approach leads to an accuracy for difficult long-range contacts of about 30%, roughly 10% above the state-of-the-art. Many variations in the architectures and the training algorithms are possible, leaving room for further improvements. Furthermore, the approach is applicable to other problems with strong underlying spatial and temporal components. 1

4 0.67825729 87 nips-2012-Convolutional-Recursive Deep Learning for 3D Object Classification

Author: Richard Socher, Brody Huval, Bharath Bath, Christopher D. Manning, Andrew Y. Ng

Abstract: Recent advances in 3D sensing technologies make it possible to easily record color and depth images which together can improve object recognition. Most current methods rely on very well-designed features for this new 3D modality. We introduce a model based on a combination of convolutional and recursive neural networks (CNN and RNN) for learning features and classifying RGB-D images. The CNN layer learns low-level translationally invariant features which are then given as inputs to multiple, fixed-tree RNNs in order to compose higher order features. RNNs can be seen as combining convolution and pooling into one efficient, hierarchical operation. Our main result is that even RNNs with random weights compose powerful features. Our model obtains state of the art performance on a standard RGB-D object dataset while being more accurate and faster during training and testing than comparable architectures such as two-layer CNNs. 1

5 0.62831295 159 nips-2012-Image Denoising and Inpainting with Deep Neural Networks

Author: Junyuan Xie, Linli Xu, Enhong Chen

Abstract: We present a novel approach to low-level vision problems that combines sparse coding and deep networks pre-trained with denoising auto-encoder (DA). We propose an alternative training scheme that successfully adapts DA, originally designed for unsupervised feature learning, to the tasks of image denoising and blind inpainting. Our method’s performance in the image denoising task is comparable to that of KSVD which is a widely used sparse coding technique. More importantly, in blind image inpainting task, the proposed method provides solutions to some complex problems that have not been tackled before. Specifically, we can automatically remove complex patterns like superimposed text from an image, rather than simple patterns like pixels missing at random. Moreover, the proposed method does not need the information regarding the region that requires inpainting to be given a priori. Experimental results demonstrate the effectiveness of the proposed method in the tasks of image denoising and blind inpainting. We also show that our new training scheme for DA is more effective and can improve the performance of unsupervised feature learning. 1

6 0.62544763 193 nips-2012-Learning to Align from Scratch

7 0.59905934 92 nips-2012-Deep Representations and Codes for Image Auto-Annotation

8 0.59245068 8 nips-2012-A Generative Model for Parts-based Object Segmentation

9 0.56694227 210 nips-2012-Memorability of Image Regions

10 0.56269693 90 nips-2012-Deep Learning of Invariant Features via Simulated Fixations in Video

11 0.5620802 170 nips-2012-Large Scale Distributed Deep Networks

12 0.54057872 146 nips-2012-Graphical Gaussian Vector for Image Categorization

13 0.53698182 39 nips-2012-Analog readout for optical reservoir computers

14 0.53687936 202 nips-2012-Locally Uniform Comparison Image Descriptor

15 0.51260734 101 nips-2012-Discriminatively Trained Sparse Code Gradients for Contour Detection

16 0.51002187 230 nips-2012-Multiple Choice Learning: Learning to Produce Multiple Structured Outputs

17 0.50022012 176 nips-2012-Learning Image Descriptors with the Boosting-Trick

18 0.49369368 72 nips-2012-Cocktail Party Processing via Structured Prediction

19 0.49261612 303 nips-2012-Searching for objects driven by context

20 0.48408619 238 nips-2012-Neurally Plausible Reinforcement Learning of Working Memory Tasks


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(0, 0.023), (17, 0.011), (21, 0.049), (38, 0.064), (42, 0.023), (54, 0.023), (55, 0.08), (74, 0.087), (75, 0.307), (76, 0.109), (80, 0.083), (92, 0.045)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.75473303 91 nips-2012-Deep Neural Networks Segment Neuronal Membranes in Electron Microscopy Images

Author: Dan Ciresan, Alessandro Giusti, Luca M. Gambardella, Jürgen Schmidhuber

Abstract: We address a central problem of neuroanatomy, namely, the automatic segmentation of neuronal structures depicted in stacks of electron microscopy (EM) images. This is necessary to efficiently map 3D brain structure and connectivity. To segment biological neuron membranes, we use a special type of deep artificial neural network as a pixel classifier. The label of each pixel (membrane or nonmembrane) is predicted from raw pixel values in a square window centered on it. The input layer maps each window pixel to a neuron. It is followed by a succession of convolutional and max-pooling layers which preserve 2D information and extract features with increasing levels of abstraction. The output layer produces a calibrated probability for each class. The classifier is trained by plain gradient descent on a 512 × 512 × 30 stack with known ground truth, and tested on a stack of the same size (ground truth unknown to the authors) by the organizers of the ISBI 2012 EM Segmentation Challenge. Even without problem-specific postprocessing, our approach outperforms competing techniques by a large margin in all three considered metrics, i.e. rand error, warping error and pixel error. For pixel error, our approach is the only one outperforming a second human observer. 1

2 0.51806474 158 nips-2012-ImageNet Classification with Deep Convolutional Neural Networks

Author: Alex Krizhevsky, Ilya Sutskever, Geoffrey E. Hinton

Abstract: We trained a large, deep convolutional neural network to classify the 1.2 million high-resolution images in the ImageNet LSVRC-2010 contest into the 1000 different classes. On the test data, we achieved top-1 and top-5 error rates of 37.5% and 17.0% which is considerably better than the previous state-of-the-art. The neural network, which has 60 million parameters and 650,000 neurons, consists of five convolutional layers, some of which are followed by max-pooling layers, and three fully-connected layers with a final 1000-way softmax. To make training faster, we used non-saturating neurons and a very efficient GPU implementation of the convolution operation. To reduce overfitting in the fully-connected layers we employed a recently-developed regularization method called “dropout” that proved to be very effective. We also entered a variant of this model in the ILSVRC-2012 competition and achieved a winning top-5 test error rate of 15.3%, compared to 26.2% achieved by the second-best entry. 1

3 0.51136804 210 nips-2012-Memorability of Image Regions

Author: Aditya Khosla, Jianxiong Xiao, Antonio Torralba, Aude Oliva

Abstract: While long term human visual memory can store a remarkable amount of visual information, it tends to degrade over time. Recent works have shown that image memorability is an intrinsic property of an image that can be reliably estimated using state-of-the-art image features and machine learning algorithms. However, the class of features and image information that is forgotten has not been explored yet. In this work, we propose a probabilistic framework that models how and which local regions from an image may be forgotten using a data-driven approach that combines local and global images features. The model automatically discovers memorability maps of individual images without any human annotation. We incorporate multiple image region attributes in our algorithm, leading to improved memorability prediction of images as compared to previous works. 1

4 0.51128376 193 nips-2012-Learning to Align from Scratch

Author: Gary Huang, Marwan Mattar, Honglak Lee, Erik G. Learned-miller

Abstract: Unsupervised joint alignment of images has been demonstrated to improve performance on recognition tasks such as face verification. Such alignment reduces undesired variability due to factors such as pose, while only requiring weak supervision in the form of poorly aligned examples. However, prior work on unsupervised alignment of complex, real-world images has required the careful selection of feature representation based on hand-crafted image descriptors, in order to achieve an appropriate, smooth optimization landscape. In this paper, we instead propose a novel combination of unsupervised joint alignment with unsupervised feature learning. Specifically, we incorporate deep learning into the congealing alignment framework. Through deep learning, we obtain features that can represent the image at differing resolutions based on network depth, and that are tuned to the statistics of the specific data being aligned. In addition, we modify the learning algorithm for the restricted Boltzmann machine by incorporating a group sparsity penalty, leading to a topographic organization of the learned filters and improving subsequent alignment results. We apply our method to the Labeled Faces in the Wild database (LFW). Using the aligned images produced by our proposed unsupervised algorithm, we achieve higher accuracy in face verification compared to prior work in both unsupervised and supervised alignment. We also match the accuracy for the best available commercial method. 1

5 0.50763202 52 nips-2012-Bayesian Nonparametric Modeling of Suicide Attempts

Author: Francisco Ruiz, Isabel Valera, Carlos Blanco, Fernando Pérez-Cruz

Abstract: The National Epidemiologic Survey on Alcohol and Related Conditions (NESARC) database contains a large amount of information, regarding the way of life, medical conditions, etc., of a representative sample of the U.S. population. In this paper, we are interested in seeking the hidden causes behind the suicide attempts, for which we propose to model the subjects using a nonparametric latent model based on the Indian Buffet Process (IBP). Due to the nature of the data, we need to adapt the observation model for discrete random variables. We propose a generative model in which the observations are drawn from a multinomial-logit distribution given the IBP matrix. The implementation of an efficient Gibbs sampler is accomplished using the Laplace approximation, which allows integrating out the weighting factors of the multinomial-logit likelihood model. Finally, the experiments over the NESARC database show that our model properly captures some of the hidden causes that model suicide attempts. 1

6 0.50655013 215 nips-2012-Minimizing Uncertainty in Pipelines

7 0.50445151 90 nips-2012-Deep Learning of Invariant Features via Simulated Fixations in Video

8 0.50388885 168 nips-2012-Kernel Latent SVM for Visual Recognition

9 0.5031234 274 nips-2012-Priors for Diversity in Generative Latent Variable Models

10 0.50137138 306 nips-2012-Semantic Kernel Forests from Multiple Taxonomies

11 0.49725768 188 nips-2012-Learning from Distributions via Support Measure Machines

12 0.49716121 197 nips-2012-Learning with Recursive Perceptual Representations

13 0.49695694 229 nips-2012-Multimodal Learning with Deep Boltzmann Machines

14 0.49628988 201 nips-2012-Localizing 3D cuboids in single-view images

15 0.4961147 211 nips-2012-Meta-Gaussian Information Bottleneck

16 0.49470294 3 nips-2012-A Bayesian Approach for Policy Learning from Trajectory Preference Queries

17 0.49406177 303 nips-2012-Searching for objects driven by context

18 0.49340865 8 nips-2012-A Generative Model for Parts-based Object Segmentation

19 0.49305362 101 nips-2012-Discriminatively Trained Sparse Code Gradients for Contour Detection

20 0.49229974 92 nips-2012-Deep Representations and Codes for Image Auto-Annotation