nips nips2012 nips2012-127 knowledge-graph by maker-knowledge-mining

127 nips-2012-Fast Bayesian Inference for Non-Conjugate Gaussian Process Regression


Source: pdf

Author: Emtiyaz Khan, Shakir Mohamed, Kevin P. Murphy

Abstract: We present a new variational inference algorithm for Gaussian process regression with non-conjugate likelihood functions, with application to a wide array of problems including binary and multi-class classification, and ordinal regression. Our method constructs a concave lower bound that is optimized using an efficient fixed-point updating algorithm. We show that the new algorithm has highly competitive computational complexity, matching that of alternative approximate inference methods. We also prove that the use of concave variational bounds provides stable and guaranteed convergence – a property not available to other approaches. We show empirically for both binary and multi-class classification that our new algorithm converges much faster than existing variational methods, and without any degradation in performance. 1

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 Our method constructs a concave lower bound that is optimized using an efficient fixed-point updating algorithm. [sent-3, score-0.388]

2 We also prove that the use of concave variational bounds provides stable and guaranteed convergence – a property not available to other approaches. [sent-5, score-0.548]

3 We show empirically for both binary and multi-class classification that our new algorithm converges much faster than existing variational methods, and without any degradation in performance. [sent-6, score-0.309]

4 For real-valued outputs, we can combine the GP prior with a Gaussian likelihood and perform exact posterior inference in closed form. [sent-8, score-0.241]

5 However, in other cases, such as classification, the likelihood is no longer conjugate to the GP prior, and exact inference is no longer tractable. [sent-9, score-0.145]

6 In this paper, we use a variational approach, where we compute a lower bound to the log marginal likelihood using Jensen’s inequality. [sent-19, score-0.551]

7 Unfortunately, most of these methods are slow, since they attempt to solve for the posterior covariance matrix, which has size O(N 2 ), where N is the number of data points. [sent-22, score-0.147]

8 In [19], a reparameterization was proposed that only requires computing O(N ) variational parameters. [sent-23, score-0.358]

9 In this paper, we propose a new lower bound that is concave, and derive an efficient iterative algorithm for its maximization. [sent-25, score-0.147]

10 Table 1: Gaussian process regression (top left) and its graphical model (right), along with the example likelihoods for outputs (bottom left). [sent-28, score-0.163]

11 2 Gaussian Process Regression Gaussian process (GP) regression is a powerful method for non-parametric regression that has gained a great deal of attention as a flexible and accurate modeling approach. [sent-33, score-0.153]

12 Consider N data points with the n’th observation denoted by yn , with corresponding features xn . [sent-34, score-0.13]

13 A Gaussian process model uses a non-linear latent function z(x) to obtain the distribution of the observation y using an appropriate likelihood [15, 18]. [sent-35, score-0.149]

14 A Gaussian process [20] specifies a distribution over z(x), and is a stochastic process that is characterized by a mean function µ(x) and a covariance function Σ(x, x ), which are specified using a kernel function that depends on the observed features x. [sent-38, score-0.16]

15 We consider frequently encountered data such as binary, ordinal, categorical and count observations, and describe their likelihoods in Table 1. [sent-55, score-0.194]

16 For the case of categorical observations, the latent function z is a vector whose k’th element is the latent function for k’th category. [sent-56, score-0.184]

17 In all cases, the likelihoods we consider are not conjugate to the Gaussian prior distribution and as a result, the posterior distribution is intractable. [sent-59, score-0.163]

18 Similarly, the integrations required in computing the predictive distribution and the marginal likelihood are intractable. [sent-60, score-0.121]

19 To deal with this intractability we make use of variational methods. [sent-61, score-0.213]

20 3 Variational Lower Bound to the Log Marginal Likelihood Inference and model selection are always problematic in any Gaussian process regression using nonconjugate likelihoods due to the fact that the marginal likelihood contains an intractable integral. [sent-62, score-0.284]

21 In this section, we derive a tractable variational lower bound to the marginal likelihood. [sent-63, score-0.41]

22 We show 2 that the lower bound takes a well known form and can be maximized using concave optimization. [sent-64, score-0.388]

23 Throughout the section, we assume scalar zn , with extension to the vector case being straightforward. [sent-65, score-0.164]

24 We begin with the intractable log marginal likelihood L(θ) in Eq. [sent-66, score-0.191]

25 We use a Gaussian posterior with mean m and covariance V. [sent-68, score-0.147]

26 The full set of variational parameters is thus γ = {m, V}. [sent-69, score-0.213]

27 As log is a concave function, we obtain a lower bound LJ (θ, γ) using Jensen’s inequality, given in Eq. [sent-70, score-0.458]

28 The first integral is simply the Kullback−Leibler (KL) divergence from the variational Gaussian posterior q(z|m, V) to the GP prior p(z|µ, Σ) as shown in Eq. [sent-72, score-0.309]

29 The second integral can be expressed in terms of the expectation with respect to the marginal q(zn |mn , Vnn ) as shown in the second term of Eq. [sent-75, score-0.129]

30 Here mn is the n’th element of m and Vnn is the n’th diagonal element of V, the two variables collectively denoted by γn . [sent-77, score-0.222]

31 The lower bound LJ is still intractable since the expectation of log p(yn |zn ) is not available in closed form for the distributions listed in Table 1. [sent-78, score-0.258]

32 To derive a tractable lower bound, we make use of local variational bounds (LVB) fb , defined such that E[log p(yn |zn )] ≥ fb (yn , mn , Vnn ), giving us Eq. [sent-79, score-0.973]

33 (6) n=1 We discuss the choice of LVBs in the next section, but first discuss the well-known form that the lower bound of Eq. [sent-82, score-0.147]

34 Similarly, the function with respect to V is similar to the graphical lasso [8] or covariance selection problem [7], but is different in that the argument is a covariance matrix instead of a precision matrix [8]. [sent-85, score-0.202]

35 These two objective functions are coupled through the non-linear term fb (·). [sent-86, score-0.267]

36 In our case, this term arises from the likelihood, and is smooth and concave as we discuss in next section. [sent-88, score-0.241]

37 It is straightforward to show that the variational lower bound is strictly concave with respect to γ if fb is jointly concave with respect to mn and Vnn . [sent-89, score-1.336]

38 Strict concavity of terms other than fb is well-known since both the least squares and covariance selection problems are concave. [sent-90, score-0.457]

39 Similar concavity results have been discussed by Braun and McAuliffe [5] for the discrete choice model, and more recently by Challis and Barber [6] for the Bayesian linear model, who consider concavity with respect to the Cholesky factor of V. [sent-91, score-0.254]

40 We consider concavity with respect to V instead of its Cholesky factor, which allows us to exploit the special structure of V, as explained in Section 5. [sent-92, score-0.146]

41 4 Concave Local Variational Bounds In this section, we describe concave LVBs for various likelihoods. [sent-93, score-0.241]

42 We describe the LVBs for the likelihoods given in Table 1 with z being a scalar for count, binary, and ordinal data, but a vector of length K for categorical data, K being the number of classes. [sent-95, score-0.257]

43 For the Poison distribution, the expectation is available in closed form and we do not need any bounding: E[log p(y|η)] = ym − exp(m + v/2) − log y! [sent-97, score-0.141]

44 This function is jointly concave with respect to m and v since the exponential is a convex function. [sent-99, score-0.309]

45 3 For binary data, we use the piecewise linear/quadratic bounds proposed by [16], which is a bound on the logistic-log-partition (LLP) function log(1 + exp(x)) and can be used to obtain a bound over the sigmoid function σ(x). [sent-100, score-0.401]

46 The final bound can be expressed as sum of R pieces: E(log p(y|η)) = R fb (y, m, v) = ym − r=1 fbr (m, v) where fbr is the expectation of r’th quadratic piece. [sent-101, score-0.593]

47 The function fbr is jointly concave with respect to m, v and their gradients are available in closed-form. [sent-102, score-0.424]

48 An important property of the piecewise bound is that its maximum error is bounded and can be driven to zero by increasing the number of pieces. [sent-103, score-0.17]

49 For this reason, this bound always performs better than other existing bounds, such as Jaakola’s bound [12], given that the number of pieces is chosen appropriately. [sent-106, score-0.247]

50 Finally, the cumulative logit likeilhood for ordinal observations depends on σ(x) and its expectation can be bounded using piecewise bounds in a similar way. [sent-107, score-0.387]

51 For the multinomial logit distribution, we can use the bounds proposed by [3] and [4], both leading to concave LVBs. [sent-108, score-0.487]

52 The first bound takes the form fb (y, m, V) = yT m − lse(m + v/2) with y represented using a 1-of-K encoding. [sent-109, score-0.358]

53 This function is jointly concave with respect to m and v, which can be shown by noting the fact that the log-sum-exp function is convex. [sent-110, score-0.309]

54 The second bound is the product of sigmoids bound proposed by [4] which bounds the likelihood with product of sigmoids (see Eq. [sent-111, score-0.398]

55 We can also use piecewise linear/quadratic bound to bound each sigmoid. [sent-113, score-0.261]

56 Alternatively, we can use the recently proposed stick-breaking likelihood of [14] which uses piecewise bounds as well. [sent-114, score-0.199]

57 Finally, note that the original log-likelihood may not be concave itself, but if it is such that LJ has a unique solution, then designing a concave variational lower bound will allow us to use concave optimization to efficiently maximize the lower bound. [sent-115, score-1.139]

58 5 Existing Algorithms for Variational Inference In this section, we assume that for each output yn there is a corresponding scalar latent function zn . [sent-116, score-0.333]

59 In variational inference, we find the approximate Gaussian posterior distribution with mean m and covariance V that maximizes Eq. [sent-118, score-0.36]

60 The simplest approach is to use gradient-based methods for optimization, but this can be problematic since the number of variational parameters is quadratic in N due to the covariance matrix V. [sent-120, score-0.295]

61 The authors of [19] speculate that this may perhaps be the reason behind limited use of Gaussian variational approximations. [sent-121, score-0.213]

62 v m 7 and 8 and equate to zero, using gn := ∂fb (yn , mn , vn )/∂mn and gn := ∂fb (yn , mn , vn )/∂vn . [sent-124, score-0.398]

63 Also, gm and gv are the vectors of these gradients, and diag(gv ) is the matrix with gv as its diagonal. [sent-125, score-0.273]

64 This property can be exploited to reduce the number of variational parameters. [sent-127, score-0.213]

65 Opper and Archambeau [19] (and [18]) propose a reparameterization to reduce the number of parameters to O(N ). [sent-128, score-0.145]

66 At the maximum (but not everywhere), α and λ will be equal to gm and gv respectively. [sent-130, score-0.162]

67 Therefore, instead of solving the fixed-point equations to obtain m and V, we can reparameterize the lower bound with respect to α and λ. [sent-131, score-0.185]

68 6 and after simplification using the matrix inversion and determinant lemmas, we get the following new objective function (for a detailed derivation, see [18]), N 1 2 − log(|Bλ ||diag(λ)|) + Tr(B−1 Σ) − αT Σα + λ fb (yn , mn , Vnn ), n=1 4 (11) with Bλ = diag(λ)−1 + Σ. [sent-134, score-0.463]

69 The new lower bound involves vectors of size N , reducing the number of variational parameters to O(N ). [sent-137, score-0.36]

70 The problem with this reparameterization is that the new lower bound is no longer concave, even though it has a unique maximum. [sent-138, score-0.292]

71 We substitute the reparameterization V = (Σ−1 + λ)−1 to get a new function f (λ) = 1 [− log(1 + Σλ) − (1 + Σλ)−1 ]/2. [sent-142, score-0.233]

72 Clearly, this derivative is negative for λ < 1/Σ and non-negative otherwise, making the function neither concave nor convex. [sent-144, score-0.241]

73 With the reparameterization, we loose concavity and therefore the algorithm may have slow convergence. [sent-146, score-0.152]

74 6 Fast Convergent Variational Inference using Coordinate Ascent We now derive an algorithm that reduces the number of variational parameters to 2N while maintaining concavity. [sent-148, score-0.213]

75 14 at the solution, where g22 is the gradient of fb with respect to v22 . [sent-167, score-0.305]

76 v 0 = k22 − Ω22 + 2g22 (14) v 0 = k22 + 1/v22 − Ω22 + 2g22 (15) f (v) = log(v) − (Ω22 − k22 )v + 2fb (y2 , m22 , v) (16) The function f (v) is a strictly concave function and can be optimized by iterating the following v update: v22 ← 1/(Ω22 − k22 − 2g22 ). [sent-174, score-0.271]

77 and denote the old old old values by k22 and v22 . [sent-184, score-0.687]

78 old old K−1 k12 = −(k22 − k22 )vold = −vold /v22 12 12 11 K−1 = Vold − 11 11 K−1 k12 kT K−1 12 11 11 old k22 − k22 old = Vold − vold (vold )T /v22 11 12 12 (17) (18) Note that both K−1 and k12 do not change after the fixed point iteration. [sent-194, score-1.135]

79 This is indeed the case for us since each fixed point iteration solves a concave problem of the form given by Eq. [sent-210, score-0.241]

80 We evaluate the performance of our fast variational inference algorithm against existing inference methods for 1 http://www. [sent-225, score-0.425]

81 For binary classification, we use the UCI ionosphere data (with 351 data examples containing 34 features). [sent-254, score-0.134]

82 We consider GP classification using the Bernoulli logit likelihood, for which we use the piecewise bound of [16] with 20 pieces. [sent-257, score-0.317]

83 In all cases, due to non-concavity, the optimization of the Opper and Archambeau reparameterization (black curve with squares) convergence slowly, passing through flat regions of the objective and requiring a large number of computations to reach convergence. [sent-268, score-0.19]

84 We show results comparing the lower bound vs the number of flops taken in Figure 1(b), for four hyperparameter settings {log(s), log(σ)}. [sent-278, score-0.18]

85 5) 20K 40K 60K 80K 100K 0 10K Mega−flops Mega−Flops (a) Ionosphere data 20K 30K 40K 50K Mega−flops (b) Forensic glass data Figure 1: Convergence results for (a) the binary classification on the ionosphere data set and (b) the multi-class classification on the glass dataset. [sent-296, score-0.244]

86 We plot the negative of the lower bound vs the number of flops. [sent-297, score-0.147]

87 where both methods reach the same lower bound value, but the existing approach converging much slower, with our algorithm always converged within 20 iterations. [sent-300, score-0.181]

88 8 Discussion In this paper we have presented a new variational inference algorithm for non-conjugate GP regression. [sent-301, score-0.287]

89 We derived a concave variational lower bound to the log marginal likelihood, and used concavity to develop an efficient optimization algorithm. [sent-302, score-0.829]

90 Our algorithm uses an efficient approach where we update the marginals of the posterior and then do a rank one update of the covariance matrix. [sent-306, score-0.273]

91 Both EP and our algorithm update posterior marginals, followed by a rank-one update of the covariance. [sent-312, score-0.191]

92 The derivation of our algorithm is based on the observation that the posterior covariance has a special structure, and does not directly use the concavity of the lower bound. [sent-315, score-0.367]

93 An alternate derivation based on the Fenchel duality exists and shows that the fixed-point iterations compute dual variables which are related to the gradients of fb . [sent-316, score-0.391]

94 Efficient bounds for the softmax and applications to approximate inference in hybrid models. [sent-341, score-0.123]

95 Concave Gaussian variational approximations for inference in largescale Bayesian linear models. [sent-351, score-0.287]

96 Data augmentation and MCMC for binary and multiu u nomial logit models. [sent-366, score-0.241]

97 Variational Bayesian multinomial probit regression with Gaussian process priors. [sent-371, score-0.146]

98 A variational approach to Bayesian logistic regression problems and their extensions. [sent-381, score-0.27]

99 A stick-breaking likelihood for categorical data analysis with latent Gaussian models. [sent-394, score-0.187]

100 Data augmentation, frequentist estimation, and the Bayesian analysis of multinomial logit models. [sent-446, score-0.197]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('fb', 0.267), ('concave', 0.241), ('old', 0.229), ('flops', 0.219), ('loglik', 0.219), ('mega', 0.219), ('vold', 0.219), ('variational', 0.213), ('neg', 0.193), ('gp', 0.192), ('vnn', 0.164), ('logit', 0.147), ('reparameterization', 0.145), ('opper', 0.142), ('vnew', 0.137), ('yn', 0.13), ('zn', 0.122), ('mn', 0.121), ('archambeau', 0.111), ('gv', 0.111), ('concavity', 0.108), ('kt', 0.1), ('bound', 0.091), ('lj', 0.086), ('covariance', 0.082), ('fbr', 0.082), ('lvbs', 0.082), ('piecewise', 0.079), ('th', 0.079), ('categorical', 0.077), ('gaussian', 0.077), ('inference', 0.074), ('ep', 0.074), ('ionosphere', 0.072), ('ops', 0.072), ('likelihood', 0.071), ('ordinal', 0.071), ('log', 0.07), ('likelihoods', 0.067), ('lse', 0.067), ('dz', 0.065), ('posterior', 0.065), ('update', 0.063), ('binary', 0.062), ('regression', 0.057), ('lower', 0.056), ('derivation', 0.056), ('glass', 0.055), ('forensic', 0.055), ('lvb', 0.055), ('bayesian', 0.053), ('khan', 0.051), ('gm', 0.051), ('multinomial', 0.05), ('count', 0.05), ('marginal', 0.05), ('bounds', 0.049), ('arch', 0.048), ('sigmoids', 0.048), ('get', 0.046), ('diag', 0.046), ('convergence', 0.045), ('braun', 0.045), ('challis', 0.045), ('slow', 0.044), ('vn', 0.043), ('diagonal', 0.043), ('classi', 0.043), ('corner', 0.043), ('substitute', 0.042), ('scalar', 0.042), ('convergent', 0.041), ('expectation', 0.041), ('marlin', 0.04), ('latent', 0.039), ('process', 0.039), ('respect', 0.038), ('markers', 0.035), ('gn', 0.035), ('iterations', 0.035), ('cholesky', 0.034), ('mohamed', 0.034), ('existing', 0.034), ('gradients', 0.033), ('hyperparameter', 0.033), ('fr', 0.032), ('augmentation', 0.032), ('unimodal', 0.032), ('prior', 0.031), ('assessing', 0.031), ('pieces', 0.031), ('ym', 0.03), ('hyperparameters', 0.03), ('fast', 0.03), ('iterating', 0.03), ('jointly', 0.03), ('sigmoid', 0.029), ('inversion', 0.029), ('bernoulli', 0.029), ('element', 0.029)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 1.0 127 nips-2012-Fast Bayesian Inference for Non-Conjugate Gaussian Process Regression

Author: Emtiyaz Khan, Shakir Mohamed, Kevin P. Murphy

Abstract: We present a new variational inference algorithm for Gaussian process regression with non-conjugate likelihood functions, with application to a wide array of problems including binary and multi-class classification, and ordinal regression. Our method constructs a concave lower bound that is optimized using an efficient fixed-point updating algorithm. We show that the new algorithm has highly competitive computational complexity, matching that of alternative approximate inference methods. We also prove that the use of concave variational bounds provides stable and guaranteed convergence – a property not available to other approaches. We show empirically for both binary and multi-class classification that our new algorithm converges much faster than existing variational methods, and without any degradation in performance. 1

2 0.19008015 272 nips-2012-Practical Bayesian Optimization of Machine Learning Algorithms

Author: Jasper Snoek, Hugo Larochelle, Ryan P. Adams

Abstract: The use of machine learning algorithms frequently involves careful tuning of learning parameters and model hyperparameters. Unfortunately, this tuning is often a “black art” requiring expert experience, rules of thumb, or sometimes bruteforce search. There is therefore great appeal for automatic approaches that can optimize the performance of any given learning algorithm to the problem at hand. In this work, we consider this problem through the framework of Bayesian optimization, in which a learning algorithm’s generalization performance is modeled as a sample from a Gaussian process (GP). We show that certain choices for the nature of the GP, such as the type of kernel and the treatment of its hyperparameters, can play a crucial role in obtaining a good optimizer that can achieve expertlevel performance. We describe new algorithms that take into account the variable cost (duration) of learning algorithm experiments and that can leverage the presence of multiple cores for parallel experimentation. We show that these proposed algorithms improve on previous automatic procedures and can reach or surpass human expert-level optimization for many algorithms including latent Dirichlet allocation, structured SVMs and convolutional neural networks. 1

3 0.1590573 33 nips-2012-Active Learning of Model Evidence Using Bayesian Quadrature

Author: Michael Osborne, Roman Garnett, Zoubin Ghahramani, David K. Duvenaud, Stephen J. Roberts, Carl E. Rasmussen

Abstract: Numerical integration is a key component of many problems in scientific computing, statistical modelling, and machine learning. Bayesian Quadrature is a modelbased method for numerical integration which, relative to standard Monte Carlo methods, offers increased sample efficiency and a more robust estimate of the uncertainty in the estimated integral. We propose a novel Bayesian Quadrature approach for numerical integration when the integrand is non-negative, such as the case of computing the marginal likelihood, predictive distribution, or normalising constant of a probabilistic model. Our approach approximately marginalises the quadrature model’s hyperparameters in closed form, and introduces an active learning scheme to optimally select function evaluations, as opposed to using Monte Carlo samples. We demonstrate our method on both a number of synthetic benchmarks and a real scientific problem from astronomy. 1

4 0.13645227 187 nips-2012-Learning curves for multi-task Gaussian process regression

Author: Peter Sollich, Simon Ashton

Abstract: We study the average case performance of multi-task Gaussian process (GP) regression as captured in the learning curve, i.e. the average Bayes error for a chosen task versus the total number of examples n for all tasks. For GP covariances that are the product of an input-dependent covariance function and a free-form intertask covariance matrix, we show that accurate approximations for the learning curve can be obtained for an arbitrary number of tasks T . We use these to study the asymptotic learning behaviour for large n. Surprisingly, multi-task learning can be asymptotically essentially useless, in the sense that examples from other tasks help only when the degree of inter-task correlation, ρ, is near its maximal value ρ = 1. This effect is most extreme for learning of smooth target functions as described by e.g. squared exponential kernels. We also demonstrate that when learning many tasks, the learning curves separate into an initial phase, where the Bayes error on each task is reduced down to a plateau value by “collective learning” even though most tasks have not seen examples, and a final decay that occurs once the number of examples is proportional to the number of tasks. 1 Introduction and motivation Gaussian processes (GPs) [1] have been popular in the NIPS community for a number of years now, as one of the key non-parametric Bayesian inference approaches. In the simplest case one can use a GP prior when learning a function from data. In line with growing interest in multi-task or transfer learning, where relatedness between tasks is used to aid learning of the individual tasks (see e.g. [2, 3]), GPs have increasingly also been used in a multi-task setting. A number of different choices of covariance functions have been proposed [4, 5, 6, 7, 8]. These differ e.g. in assumptions on whether the functions to be learned are related to a smaller number of latent functions or have free-form inter-task correlations; for a recent review see [9]. Given this interest in multi-task GPs, one would like to quantify the benefits that they bring compared to single-task learning. PAC-style bounds for classification [2, 3, 10] in more general multi-task scenarios exist, but there has been little work on average case analysis. The basic question in this setting is: how does the Bayes error on a given task depend on the number of training examples for all tasks, when averaged over all data sets of the given size. For a single regression task, this learning curve has become relatively well understood since the late 1990s, with a number of bounds and approximations available [11, 12, 13, 14, 15, 16, 17, 18, 19] as well as some exact predictions [20]. Already two-task GP regression is much more difficult to analyse, and progress was made only very recently at NIPS 2009 [21], where upper and lower bounds for learning curves were derived. The tightest of these bounds, however, either required evaluation by Monte Carlo sampling, or assumed knowledge of the corresponding single-task learning curves. Here our aim is to obtain accurate learning curve approximations that apply to an arbitrary number T of tasks, and that can be evaluated explicitly without recourse to sampling. 1 We begin (Sec. 2) by expressing the Bayes error for any single task in a multi-task GP regression problem in a convenient feature space form, where individual training examples enter additively. This requires the introduction of a non-trivial tensor structure combining feature space components and tasks. Considering the change in error when adding an example for some task leads to partial differential equations linking the Bayes errors for all tasks. Solving these using the method of characteristics then gives, as our primary result, the desired learning curve approximation (Sec. 3). In Sec. 4 we discuss some of its predictions. The approximation correctly delineates the limits of pure transfer learning, when all examples are from tasks other than the one of interest. Next we compare with numerical simulations for some two-task scenarios, finding good qualitative agreement. These results also highlight a surprising feature, namely that asymptotically the relatedness between tasks can become much less useful. We analyse this effect in some detail, showing that it is most extreme for learning of smooth functions. Finally we discuss the case of many tasks, where there is an unexpected separation of the learning curves into a fast initial error decay arising from “collective learning”, and a much slower final part where tasks are learned almost independently. 2 GP regression and Bayes error We consider GP regression for T functions fτ (x), τ = 1, 2, . . . , T . These functions have to be learned from n training examples (x , τ , y ), = 1, . . . , n. Here x is the training input, τ ∈ {1, . . . , T } denotes which task the example relates to, and y is the corresponding training output. We assume that the latter is given by the target function value fτ (x ) corrupted by i.i.d. additive 2 2 Gaussian noise with zero mean and variance στ . This setup allows the noise level στ to depend on the task. In GP regression the prior over the functions fτ (x) is a Gaussian process. This means that for any set of inputs x and task labels τ , the function values {fτ (x )} have a joint Gaussian distribution. As is common we assume this to have zero mean, so the multi-task GP is fully specified by the covariances fτ (x)fτ (x ) = C(τ, x, τ , x ). For this covariance we take the flexible form from [5], fτ (x)fτ (x ) = Dτ τ C(x, x ). Here C(x, x ) determines the covariance between function values at different input points, encoding “spatial” behaviour such as smoothness and the lengthscale(s) over which the functions vary, while the matrix D is a free-form inter-task covariance matrix. One of the attractions of GPs for regression is that, even though they are non-parametric models with (in general) an infinite number of degrees of freedom, predictions can be made in closed form, see e.g. [1]. For a test point x for task τ , one would predict as output the mean of fτ (x) over the (Gaussian) posterior, which is y T K −1 kτ (x). Here K is the n × n Gram matrix with entries 2 K m = Dτ τm C(x , xm ) + στ δ m , while kτ (x) is a vector with the n entries kτ, = Dτ τ C(x , x). The error bar would be taken as the square root of the posterior variance of fτ (x), which is T Vτ (x) = Dτ τ C(x, x) − kτ (x)K −1 kτ (x) (1) The learning curve for task τ is defined as the mean-squared prediction error, averaged over the location of test input x and over all data sets with a specified number of examples for each task, say n1 for task 1 and so on. As is standard in learning curve analysis we consider a matched scenario where the training outputs y are generated from the same prior and noise model that we use for inference. In this case the mean-squared prediction error ˆτ is the Bayes error, and is given by the average posterior variance [1], i.e. ˆτ = Vτ (x) x . To obtain the learning curve this is averaged over the location of the training inputs x : τ = ˆτ . This average presents the main challenge for learning curve prediction because the training inputs feature in a highly nonlinear way in Vτ (x). Note that the training outputs, on the other hand, do not appear in the posterior variance Vτ (x) and so do not need to be averaged over. We now want to write the Bayes error ˆτ in a form convenient for performing, at least approximately, the averages required for the learning curve. Assume that all training inputs x , and also the test input x, are drawn from the same distribution P (x). One can decompose the input-dependent part of the covariance function into eigenfunctions relative to P (x), according to C(x, x ) = i λi φi (x)φi (x ). The eigenfunctions are defined by the condition C(x, x )φi (x ) x = λi φi (x) and can be chosen to be orthonormal with respect to P (x), φi (x)φj (x) x = δij . The sum over i here is in general infinite (unless the covariance function is degenerate, as e.g. for the dot product kernel C(x, x ) = x · x ). To make the algebra below as simple as possible, we let the eigenvalues λi be arranged in decreasing order and truncate the sum to the finite range i = 1, . . . , M ; M is then some large effective feature space dimension and can be taken to infinity at the end. 2 In terms of the above eigenfunction decomposition, the Gram matrix has elements K m = Dτ 2 λi φi (x )φi (xm )+στ δ τm m δτ = i ,τ φi (x )λi δij Dτ τ φj (xm )δτ 2 ,τm +στ δ m i,τ,j,τ or in matrix form K = ΨLΨT + Σ where Σ is the diagonal matrix from the noise variances and Ψ = δτ ,iτ ,τ φi (x ), Liτ,jτ = λi δij Dτ τ (2) Here Ψ has its second index ranging over M (number of kernel eigenvalues) times T (number of tasks) values; L is a square matrix of this size. In Kronecker (tensor) product notation, L = D ⊗ Λ if we define Λ as the diagonal matrix with entries λi δij . The Kronecker product is convenient for the simplifications below; we will use that for generic square matrices, (A ⊗ B)(A ⊗ B ) = (AA ) ⊗ (BB ), (A ⊗ B)−1 = A−1 ⊗ B −1 , and tr (A ⊗ B) = (tr A)(tr B). In thinking about the mathematical expressions, it is often easier to picture Kronecker products over feature spaces and tasks as block matrices. For example, L can then be viewed as consisting of T × T blocks, each of which is proportional to Λ. To calculate the Bayes error, we need to average the posterior variance Vτ (x) over the test input x. The first term in (1) then becomes Dτ τ C(x, x) = Dτ τ tr Λ. In the second one, we need to average kτ, (x)kτ,m = Dτ τ C(x , x)C(x, xm ) x Dτm τ = x Dτ τ λi λj φi (x ) φi (x)φj (x) x φj (xm )Dτm τ ij = Dτ τ Ψl,iτ λi λj δij Ψm,jτ Dτ τ i,τ ,j,τ T In matrix form this is kτ (x)kτ (x) x = Ψ[(Deτ eT D) ⊗ Λ2 ]ΨT = ΨMτ ΨT Here the last τ equality defines Mτ , and we have denoted by eτ the T -dimensional vector with τ -th component equal to one and all others zero. Multiplying by the inverse Gram matrix K −1 and taking the trace gives the average of the second term in (1); combining with the first gives the Bayes error on task τ ˆτ = Vτ (x) x = Dτ τ tr Λ − tr ΨMτ ΨT (ΨLΨT + Σ)−1 Applying the Woodbury identity and re-arranging yields = Dτ τ tr Λ − tr Mτ ΨT Σ−1 Ψ(I + LΨT Σ−1 Ψ)−1 = ˆτ Dτ τ tr Λ − tr Mτ L−1 [I − (I + LΨT Σ−1 Ψ)−1 ] But tr Mτ L−1 = tr {[(Deτ eT D) ⊗ Λ2 ][D ⊗ Λ]−1 } τ = tr {[Deτ eT ] ⊗ Λ} = eT Deτ tr Λ = Dτ τ tr Λ τ τ so the first and second terms in the expression for ˆτ cancel and one has = tr Mτ L−1 (I + LΨT Σ−1 Ψ)−1 = tr L−1 Mτ L−1 (L−1 + ΨT Σ−1 Ψ)−1 = tr [D ⊗ Λ]−1 [(Deτ eT D) ⊗ Λ2 ][D ⊗ Λ]−1 (L−1 + ΨT Σ−1 Ψ)−1 τ = ˆτ tr [eτ eT ⊗ I](L−1 + ΨT Σ−1 Ψ)−1 τ The matrix in square brackets in the last line is just a projector Pτ onto task τ ; thought of as a matrix of T × T blocks (each of size M × M ), this has an identity matrix in the (τ, τ ) block while all other blocks are zero. We can therefore write, finally, for the Bayes error on task τ , ˆτ = tr Pτ (L−1 + ΨT Σ−1 Ψ)−1 (3) Because Σ is diagonal and given the definition (2) of Ψ, the matrix ΨT Σ−1 Ψ is a sum of contributions from the individual training examples = 1, . . . , n. This will be important for deriving the learning curve approximation below. We note in passing that, because τ Pτ = I, the sum of the Bayes errors on all tasks is τ ˆτ = tr (L−1 +ΨT Σ−1 Ψ)−1 , in close analogy to the corresponding expression for the single-task case [13]. 3 3 Learning curve prediction To obtain the learning curve τ = ˆτ , we now need to carry out the average . . . over the training inputs. To help with this, we can extend an approach for the single-task scenario [13] and define a response or resolvent matrix G = (L−1 + ΨT Σ−1 Ψ + τ vτ Pτ )−1 with auxiliary parameters vτ that will be set back to zero at the end. One can then ask how G = G and hence τ = tr Pτ G changes with the number nτ of training points for task τ . Adding an example at position x for task −2 τ increases ΨT Σ−1 Ψ by στ φτ φT , where φτ has elements (φτ )iτ = φi (x)δτ τ . Evaluating the τ −1 −2 difference (G + στ φτ φT )−1 − G with the help of the Woodbury identity and approximating it τ with a derivative gives Gφτ φT G ∂G τ =− 2 ∂nτ στ + φT Gφτ τ This needs to be averaged over the new example and all previous ones. If we approximate by averaging numerator and denominator separately we get 1 ∂G ∂G = 2 ∂nτ στ + tr Pτ G ∂vτ (4) Here we have exploited for the average over x that the matrix φτ φT x has (i, τ ), (j, τ )-entry τ φi (x)φj (x) x δτ τ δτ τ = δij δτ τ δτ τ , hence simply φτ φT x = Pτ . We have also used the τ auxiliary parameters to rewrite − GPτ G = ∂ G /∂vτ = ∂G/∂vτ . Finally, multiplying (4) by Pτ and taking the trace gives the set of quasi-linear partial differential equations ∂ τ 1 = 2 ∂nτ στ + τ ∂ τ ∂vτ (5) The remaining task is now to find the functions τ (n1 , . . . , nT , v1 , . . . , vT ) by solving these differential equations. We initially attempted to do this by tracking the τ as examples are added one task at a time, but the derivation is laborious already for T = 2 and becomes prohibitive beyond. Far more elegant is to adapt the method of characteristics to the present case. We need to find a 2T -dimensional surface in the 3T -dimensional space (n1 , . . . , nT , v1 , . . . , vT , 1 , . . . , T ), which is specified by the T functions τ (. . .). A small change (δn1 , . . . , δnT , δv1 , . . . , δvT , δ 1 , . . . , δ T ) in all 3T coordinates is tangential to this surface if it obeys the T constraints (one for each τ ) δ τ ∂ τ ∂ τ δnτ + δvτ ∂nτ ∂vτ = τ 2 From (5), one sees that this condition is satisfied whenever δ τ = 0 and δnτ = −δvτ (στ + τ ) It follows that all the characteristic curves given by τ (t) = τ,0 = const., vτ (t) = vτ,0 (1 − t), 2 nτ (t) = vτ,0 (στ + τ,0 ) t for t ∈ [0, 1] are tangential to the solution surface for all t, so lie within this surface if the initial point at t = 0 does. Because at t = 0 there are no training examples (nτ (0) = 0), this initial condition is satisfied by setting −1 τ,0 = tr Pτ −1 L + vτ ,0 Pτ τ Because t=1 τ (t) is constant along the characteristic curve, we get by equating the values at t = 0 and −1 τ,0 = tr Pτ L −1 + vτ ,0 Pτ = τ ({nτ = vτ 2 ,0 (στ + τ ,0 )}, {vτ = 0}) τ Expressing vτ ,0 in terms of nτ gives then τ = tr Pτ L−1 + τ nτ 2 στ + −1 Pτ (6) τ This is our main result: a closed set of T self-consistency equations for the average Bayes errors 2 τ . Given L as defined by the eigenvalues λi of the covariance function, the noise levels στ and the 4 number of examples nτ for each task, it is straightforward to solve these equations numerically to find the average Bayes error τ for each task. The r.h.s. of (6) is easiest to evaluate if we view the matrix inside the brackets as consisting of M × M blocks of size T × T (which is the reverse of the picture we have used so far). The matrix is then block diagonal, with the blocks corresponding to different eigenvalues λi . Explicitly, because L−1 = D −1 ⊗ Λ−1 , one has τ λ−1 D −1 + diag({ i = i 4 2 στ nτ + −1 }) τ (7) ττ Results and discussion We now consider the consequences of the approximate prediction (7) for multi-task learning curves in GP regression. A trivial special case is the one of uncorrelated tasks, where D is diagonal. Here one recovers T separate equations for the individual tasks as expected, which have the same form as for single-task learning [13]. 4.1 Pure transfer learning Consider now the case of pure transfer learning, where one is learning a task of interest (say τ = 1) purely from examples for other tasks. What is the lowest average Bayes error that can be obtained? Somewhat more generally, suppose we have no examples for the first T0 tasks, n1 = . . . = nT0 = 0, but a large number of examples for the remaining T1 = T − T0 tasks. Denote E = D −1 and write this in block form as E00 E01 E= T E01 E11 2 Now multiply by λ−1 and add in the lower right block a diagonal matrix N = diag({nτ /(στ + i −1 −1 τ )}τ =T0 +1,...,T ). The matrix inverse in (7) then has top left block λi [E00 + E00 E01 (λi N + −1 −1 T T E11 − E01 E00 E01 )−1 E01 E00 ]. As the number of examples for the last T1 tasks grows, so do all −1 (diagonal) elements of N . In the limit only the term λi E00 survives, and summing over i gives −1 −1 1 = tr Λ(E00 )11 = C(x, x) (E00 )11 . The Bayes error on task 1 cannot become lower than this, placing a limit on the benefits of pure transfer learning. That this prediction of the approximation (7) for such a lower limit is correct can also be checked directly: once the last T1 tasks fτ (x) (τ = T0 + 1, . . . T ) have been learn perfectly, the posterior over the first T0 functions is, by standard Gaussian conditioning, a GP with covariance C(x, x )(E00 )−1 . Averaging the posterior variance of −1 f1 (x) then gives the Bayes error on task 1 as 1 = C(x, x) (E00 )11 , as found earlier. This analysis can be extended to the case where there are some examples available also for the first T0 tasks. One finds for the generalization errors on these tasks the prediction (7) with D −1 replaced by E00 . This is again in line with the above form of the GP posterior after perfect learning of the remaining T1 tasks. 4.2 Two tasks We next analyse how well the approxiation (7) does in predicting multi-task learning curves for T = 2 tasks. Here we have the work of Chai [21] as a baseline, and as there we choose D= 1 ρ ρ 1 The diagonal elements are fixed to unity, as in a practical application where one would scale both task functions f1 (x) and f2 (x) to unit variance; the degree of correlation of the tasks is controlled by ρ. We fix π2 = n2 /n and plot learning curves against n. In numerical simulations we ensure integer values of n1 and n2 by setting n2 = nπ2 , n1 = n − n2 ; for evaluation of (7) we use 2 2 directly n2 = nπ2 , n1 = n(1 − π2 ). For simplicity we consider equal noise levels σ1 = σ2 = σ 2 . As regards the covariance function and input distribution, we analyse first the scenario studied in [21]: a squared exponential (SE) kernel C(x, x ) = exp[−(x − x )2 /(2l2 )] with lengthscale l, and one-dimensional inputs x with a Gaussian distribution N (0, 1/12). The kernel eigenvalues λi 5 1 1 1 1 ε1 ε1 0.8 1 1 ε1 ε1 0.8 0.001 1 ε1 0.8 0.001 n 10000 ε1 1 0.01 1 n 10000 0.6 0.6 0.4 0.4 0.4 0.2 0.2 n 1000 0.6 0.2 0 0 100 200 n 300 400 0 500 0 100 200 n 300 400 500 0 0 100 200 n 300 400 500 Figure 1: Average Bayes error for task 1 for two-task GP regression with kernel lengthscale l = 0.01, noise level σ 2 = 0.05 and a fraction π2 = 0.75 of examples for task 2. Solid lines: numerical simulations; dashed lines: approximation (7). Task correlation ρ2 = 0, 0.25, 0.5, 0.75, 1 from top to bottom. Left: SE covariance function, Gaussian input distribution. Middle: SE covariance, uniform inputs. Right: OU covariance, uniform inputs. Log-log plots (insets) show tendency of asymptotic uselessness, i.e. bunching of the ρ < 1 curves towards the one for ρ = 0; this effect is strongest for learning of smooth functions (left and middle). are known explicitly from [22] and decay exponentially with i. Figure 1(left) compares numerically simulated learning curves with the predictions for 1 , the average Bayes error on task 1, from (7). Five pairs of curves are shown, for ρ2 = 0, 0.25, 0.5, 0.75, 1. Note that the two extreme values represent single-task limits, where examples from task 2 are either ignored (ρ = 0) or effectively treated as being from task 1 (ρ = 1). Our predictions lie generally below the true learning curves, but qualitatively represent the trends well, in particular the variation with ρ2 . The curves for the different ρ2 values are fairly evenly spaced vertically for small number of examples, n, corresponding to a linear dependence on ρ2 . As n increases, however, the learning curves for ρ < 1 start to bunch together and separate from the one for the fully correlated case (ρ = 1). The approximation (7) correctly captures this behaviour, which is discussed in more detail below. Figure 1(middle) has analogous results for the case of inputs x uniformly distributed on the interval [0, 1]; the λi here decay exponentially with i2 [17]. Quantitative agreement between simulations and predictions is better for this case. The discussion in [17] suggests that this is because the approximation method we have used implicitly neglects spatial variation of the dataset-averaged posterior variance Vτ (x) ; but for a uniform input distribution this variation will be weak except near the ends of the input range [0, 1]. Figure 1(right) displays similar results for an OU kernel C(x, x ) = exp(−|x − x |/l), showing that our predictions also work well when learning rough (nowhere differentiable) functions. 4.3 Asymptotic uselessness The two-task results above suggest that multi-task learning is less useful asymptotically: when the number of training examples n is large, the learning curves seem to bunch towards the curve for ρ = 0, where task 2 examples are ignored, except when the two tasks are fully correlated (ρ = 1). We now study this effect. When the number of examples for all tasks becomes large, the Bayes errors τ will become small 2 and eventually be negligible compared to the noise variances στ in (7). One then has an explicit prediction for each τ , without solving T self-consistency equations. If we write, for T tasks, 2 nτ = nπτ with πτ the fraction of examples for task τ , and set γτ = πτ /στ , then for large n τ = i λ−1 D −1 + nΓ i −1 ττ = −1/2 −1 [λi (Γ1/2 DΓ1/2 )−1 i (Γ + nI]−1 Γ−1/2 )τ τ 1/2 where Γ = diag(γ1 , . . . , γT ). Using an eigendecomposition of the symmetric matrix Γ T T a=1 δa va va , one then shows in a few lines that (8) can be written as τ −1 ≈ γτ 2 a (va,τ ) δa g(nδa ) 6 (8) 1/2 DΓ = (9) 1 1 1 50000 ε 5000 r 0.1 ε 0.5 n=500 10 100 1000 n 0.1 0 0 0.2 0.4 ρ 2 0.6 0.8 1 1 10 100 1000 n Figure 2: Left: Bayes error (parameters as in Fig. 1(left), with n = 500) vs ρ2 . To focus on the error reduction with ρ, r = [ 1 (ρ) − 1 (1)]/[ 1 (0) − 1 (1)] is shown. Circles: simulations; solid line: predictions from (7). Other lines: predictions for larger n, showing the approach to asymptotic uselessness in multi-task learning of smooth functions. Inset: Analogous results for rough functions (parameters as in Fig. 1(right)). Right: Learning curve for many-task learning (T = 200, parameters otherwise as in Fig. 1(left) except ρ2 = 0.8). Notice the bend around 1 = 1 − ρ = 0.106. Solid line: simulations (steps arise because we chose to allocate examples to tasks in order τ = 1, . . . , T rather than randomly); dashed line: predictions from (7). Inset: Predictions for T = 1000, with asymptotic forms = 1 − ρ + ρ˜ and = (1 − ρ)¯ for the two learning stages shown as solid lines. −1 where g(h) = tr (Λ−1 + h)−1 = + h)−1 and va,τ is the τ -th component of the a-th i (λi eigenvector va . This is the general asymptotic form of our prediction for the average Bayes error for task τ . To get a more explicit result, consider the case where sample functions from the GP prior have (mean-square) derivatives up to order r. The kernel eigenvalues λi then decay as1 i−(2r+2) for large i, and using arguments from [17] one deduces that g(h) ∼ h−α for large h, with α = (2r +1)/(2r + 2). In (9) we can then write, for large n, g(nδa ) ≈ (δa /γτ )−α g(nγτ ) and hence τ ≈ g(nγτ ){ 2 1−α } a (va,τ ) (δa /γτ ) (10) 2 When there is only a single task, δ1 = γ1 and this expression reduces to 1 = g(nγ1 ) = g(n1 /σ1 ). 2 Thus g(nγτ ) = g(nτ /στ ) is the error we would get by ignoring all examples from tasks other than τ , and the term in {. . .} in (10) gives the “multi-task gain”, i.e. the factor by which the error is reduced because of examples from other tasks. (The absolute error reduction always vanishes trivially for n → ∞, along with the errors themselves.) One observation can be made directly. Learning of very smooth functions, as defined e.g. by the SE kernel, corresponds to r → ∞ and hence α → 1, so the multi-task gain tends to unity: multi-task learning is asymptotically useless. The only exception occurs when some of the tasks are fully correlated, because one or more of the eigenvalues δa of Γ1/2 DΓ1/2 will then be zero. Fig. 2(left) shows this effect in action, plotting Bayes error against ρ2 for the two-task setting of Fig. 1(left) with n = 500. Our predictions capture the nonlinear dependence on ρ2 quite well, though the effect is somewhat weaker in the simulations. For larger n the predictions approach a curve that is constant for ρ < 1, signifying negligible improvement from multi-task learning except at ρ = 1. It is worth contrasting this with the lower bound from [21], which is linear in ρ2 . While this provides a very good approximation to the learning curves for moderate n [21], our results here show that asymptotically this bound can become very loose. When predicting rough functions, there is some asymptotic improvement to be had from multi-task learning, though again the multi-task gain is nonlinear in ρ2 : see Fig. 2(left, inset) for the OU case, which has r = 1). A simple expression for the gain can be obtained in the limit of many tasks, to which we turn next. 1 See the discussion of Sacks-Ylvisaker conditions in e.g. [1]; we consider one-dimensional inputs here though the discussion can be generalized. 7 4.4 Many tasks We assume as for the two-task case that all inter-task correlations, Dτ,τ with τ = τ , are equal to ρ, while Dτ,τ = 1. This setup was used e.g. in [23], and can be interpreted as each task having a √ component proportional to ρ of a shared latent function, with an independent task-specific signal in addition. We assume for simplicity that we have the same number nτ = n/T of examples for 2 each task, and that all noise levels are the same, στ = σ 2 . Then also all Bayes errors τ = will be the same. Carrying out the matrix inverses in (7) explicitly, one can then write this equation as = gT (n/(σ 2 + ), ρ) (11) where gT (h, ρ) is related to the single-task function g(h) from above by gT (h, ρ) = 1−ρ T −1 (1 − ρ)g(h(1 − ρ)/T ) + ρ + T T g(h[ρ + (1 − ρ)/T ]) (12) Now consider the limit T → ∞ of many tasks. If n and hence h = n/(σ 2 + ) is kept fixed, gT (h, ρ) → (1 − ρ) + ρg(hρ); here we have taken g(0) = 1 which corresponds to tr Λ = C(x, x) x = 1 as in the examples above. One can then deduce from (11) that the Bayes error for any task will have the form = (1 − ρ) + ρ˜, where ˜ decays from one to zero with increasing n as for a single task, but with an effective noise level σ 2 = (1 − ρ + σ 2 )/ρ. Remarkably, then, ˜ even though here n/T → 0 so that for most tasks no examples have been seen, the Bayes error for each task decreases by “collective learning” to a plateau of height 1 − ρ. The remaining decay of to zero happens only once n becomes of order T . Here one can show, by taking T → ∞ at fixed h/T in (12) and inserting into (11), that = (1 − ρ)¯ where ¯ again decays as for a single task but with an effective number of examples n = n/T and effective noise level σ 2 /(1 − ρ). This final stage of ¯ ¯ learning therefore happens only when each task has seen a considerable number of exampes n/T . Fig. 2(right) validates these predictions against simulations, for a number of tasks (T = 200) that is in the same ballpark as in the many-tasks application example of [24]. The inset for T = 1000 shows clearly how the two learning curve stages separate as T becomes larger. Finally we can come back to the multi-task gain in the asymptotic stage of learning. For GP priors with sample functions with derivatives up to order r as before, the function ¯ from above will decay as (¯ /¯ 2 )−α ; since = (1 − ρ)¯ and σ 2 = σ 2 /(1 − ρ), the Bayes error is then proportional n σ ¯ to (1 − ρ)1−α . This multi-task gain again approaches unity for ρ < 1 for smooth functions (α = (2r + 1)/(2r + 2) → 1). Interestingly, for rough functions (α < 1), the multi-task gain decreases for small ρ2 as 1 − (1 − α) ρ2 and so always lies below a linear dependence on ρ2 initially. This shows that a linear-in-ρ2 lower error bound cannot generally apply to T > 2 tasks, and indeed one can verify that the derivation in [21] does not extend to this case. 5 Conclusion We have derived an approximate prediction (7) for learning curves in multi-task GP regression, valid for arbitrary inter-task correlation matrices D. This can be evaluated explicitly knowing only the kernel eigenvalues, without sampling or recourse to single-task learning curves. The approximation shows that pure transfer learning has a simple lower error bound, and provides a good qualitative account of numerically simulated learning curves. Because it can be used to study the asymptotic behaviour for large training sets, it allowed us to show that multi-task learning can become asymptotically useless: when learning smooth functions it reduces the asymptotic Bayes error only if tasks are fully correlated. For the limit of many tasks we found that, remarkably, some initial “collective learning” is possible even when most tasks have not seen examples. A much slower second learning stage then requires many examples per task. The asymptotic regime of this also showed explicitly that a lower error bound that is linear in ρ2 , the square of the inter-task correlation, is applicable only to the two-task setting T = 2. In future work it would be interesting to use our general result to investigate in more detail the consequences of specific choices for the inter-task correlations D, e.g. to represent a lower-dimensional latent factor structure. One could also try to deploy similar approximation methods to study the case of model mismatch, where the inter-task correlations D would have to be learned from data. More challenging, but worthwhile, would be an extension to multi-task covariance functions where task and input-space correlations to not factorize. 8 References [1] C K I Williams and C Rasmussen. Gaussian Processes for Machine Learning. MIT Press, Cambridge, MA, 2006. [2] J Baxter. A model of inductive bias learning. J. Artif. Intell. Res., 12:149–198, 2000. [3] S Ben-David and R S Borbely. A notion of task relatedness yielding provable multiple-task learning guarantees. Mach. Learn., 73(3):273–287, December 2008. [4] Y W Teh, M Seeger, and M I Jordan. Semiparametric latent factor models. In Workshop on Artificial Intelligence and Statistics 10, pages 333–340. Society for Artificial Intelligence and Statistics, 2005. [5] E V Bonilla, F V Agakov, and C K I Williams. Kernel multi-task learning using task-specific features. In Proceedings of the 11th International Conference on Artificial Intelligence and Statistics (AISTATS). Omni Press, 2007. [6] E V Bonilla, K M A Chai, and C K I Williams. Multi-task Gaussian process prediction. In J C Platt, D Koller, Y Singer, and S Roweis, editors, NIPS 20, pages 153–160, Cambridge, MA, 2008. MIT Press. [7] M Alvarez and N D Lawrence. Sparse convolved Gaussian processes for multi-output regression. In D Koller, D Schuurmans, Y Bengio, and L Bottou, editors, NIPS 21, pages 57–64, Cambridge, MA, 2009. MIT Press. [8] G Leen, J Peltonen, and S Kaski. Focused multi-task learning using Gaussian processes. In Dimitrios Gunopulos, Thomas Hofmann, Donato Malerba, and Michalis Vazirgiannis, editors, Machine Learning and Knowledge Discovery in Databases, volume 6912 of Lecture Notes in Computer Science, pages 310– 325. Springer Berlin, Heidelberg, 2011. ´ [9] M A Alvarez, L Rosasco, and N D Lawrence. Kernels for vector-valued functions: a review. Foundations and Trends in Machine Learning, 4:195–266, 2012. [10] A Maurer. Bounds for linear multi-task learning. J. Mach. Learn. Res., 7:117–139, 2006. [11] M Opper and F Vivarelli. General bounds on Bayes errors for regression with Gaussian processes. In M Kearns, S A Solla, and D Cohn, editors, NIPS 11, pages 302–308, Cambridge, MA, 1999. MIT Press. [12] G F Trecate, C K I Williams, and M Opper. Finite-dimensional approximation of Gaussian processes. In M Kearns, S A Solla, and D Cohn, editors, NIPS 11, pages 218–224, Cambridge, MA, 1999. MIT Press. [13] P Sollich. Learning curves for Gaussian processes. In M S Kearns, S A Solla, and D A Cohn, editors, NIPS 11, pages 344–350, Cambridge, MA, 1999. MIT Press. [14] D Malzahn and M Opper. Learning curves for Gaussian processes regression: A framework for good approximations. In T K Leen, T G Dietterich, and V Tresp, editors, NIPS 13, pages 273–279, Cambridge, MA, 2001. MIT Press. [15] D Malzahn and M Opper. A variational approach to learning curves. In T G Dietterich, S Becker, and Z Ghahramani, editors, NIPS 14, pages 463–469, Cambridge, MA, 2002. MIT Press. [16] D Malzahn and M Opper. Statistical mechanics of learning: a variational approach for real data. Phys. Rev. Lett., 89:108302, 2002. [17] P Sollich and A Halees. Learning curves for Gaussian process regression: approximations and bounds. Neural Comput., 14(6):1393–1428, 2002. [18] P Sollich. Gaussian process regression with mismatched models. In T G Dietterich, S Becker, and Z Ghahramani, editors, NIPS 14, pages 519–526, Cambridge, MA, 2002. MIT Press. [19] P Sollich. Can Gaussian process regression be made robust against model mismatch? In Deterministic and Statistical Methods in Machine Learning, volume 3635 of Lecture Notes in Artificial Intelligence, pages 199–210. Springer Berlin, Heidelberg, 2005. [20] M Urry and P Sollich. Exact larning curves for Gaussian process regression on large random graphs. In J Lafferty, C K I Williams, J Shawe-Taylor, R S Zemel, and A Culotta, editors, NIPS 23, pages 2316–2324, Cambridge, MA, 2010. MIT Press. [21] K M A Chai. Generalization errors and learning curves for regression with multi-task Gaussian processes. In Y Bengio, D Schuurmans, J Lafferty, C K I Williams, and A Culotta, editors, NIPS 22, pages 279–287, 2009. [22] H Zhu, C K I Williams, R J Rohwer, and M Morciniec. Gaussian regression and optimal finite dimensional linear models. In C M Bishop, editor, Neural Networks and Machine Learning. Springer, 1998. [23] E Rodner and J Denzler. One-shot learning of object categories using dependent Gaussian processes. In Michael Goesele, Stefan Roth, Arjan Kuijper, Bernt Schiele, and Konrad Schindler, editors, Pattern Recognition, volume 6376 of Lecture Notes in Computer Science, pages 232–241. Springer Berlin, Heidelberg, 2010. [24] T Heskes. Solving a huge number of similar tasks: a combination of multi-task learning and a hierarchical Bayesian approach. In Proceedings of the Fifteenth International Conference on Machine Learning (ICML’98), pages 233–241. Morgan Kaufmann, 1998. 9

5 0.12031204 74 nips-2012-Collaborative Gaussian Processes for Preference Learning

Author: Neil Houlsby, Ferenc Huszar, Zoubin Ghahramani, Jose M. Hernández-lobato

Abstract: We present a new model based on Gaussian processes (GPs) for learning pairwise preferences expressed by multiple users. Inference is simplified by using a preference kernel for GPs which allows us to combine supervised GP learning of user preferences with unsupervised dimensionality reduction for multi-user systems. The model not only exploits collaborative information from the shared structure in user behavior, but may also incorporate user features if they are available. Approximate inference is implemented using a combination of expectation propagation and variational Bayes. Finally, we present an efficient active learning strategy for querying preferences. The proposed technique performs favorably on real-world data against state-of-the-art multi-user preference learning algorithms. 1

6 0.12008685 354 nips-2012-Truly Nonparametric Online Variational Inference for Hierarchical Dirichlet Processes

7 0.11877798 355 nips-2012-Truncation-free Online Variational Inference for Bayesian Nonparametric Models

8 0.11722178 121 nips-2012-Expectation Propagation in Gaussian Process Dynamical Systems

9 0.11484013 37 nips-2012-Affine Independent Variational Inference

10 0.11433063 233 nips-2012-Multiresolution Gaussian Processes

11 0.10983168 104 nips-2012-Dual-Space Analysis of the Sparse Linear Model

12 0.10760481 129 nips-2012-Fast Variational Inference in the Conjugate Exponential Family

13 0.10474052 55 nips-2012-Bayesian Warped Gaussian Processes

14 0.10230425 58 nips-2012-Bayesian models for Large-scale Hierarchical Classification

15 0.092642158 5 nips-2012-A Conditional Multinomial Mixture Model for Superset Label Learning

16 0.091383375 312 nips-2012-Simultaneously Leveraging Output and Task Structures for Multiple-Output Regression

17 0.087229535 278 nips-2012-Probabilistic n-Choose-k Models for Classification and Ranking

18 0.086832799 206 nips-2012-Majorization for CRFs and Latent Likelihoods

19 0.082492821 126 nips-2012-FastEx: Hash Clustering with Exponential Families

20 0.077577718 138 nips-2012-Fully Bayesian inference for neural models with negative-binomial spiking


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, 0.222), (1, 0.056), (2, 0.061), (3, 0.005), (4, -0.161), (5, -0.01), (6, 0.025), (7, 0.051), (8, 0.019), (9, -0.244), (10, -0.109), (11, 0.049), (12, -0.024), (13, 0.088), (14, -0.041), (15, 0.115), (16, -0.048), (17, 0.005), (18, -0.054), (19, -0.01), (20, -0.057), (21, -0.021), (22, 0.034), (23, 0.088), (24, -0.005), (25, -0.001), (26, 0.035), (27, 0.014), (28, -0.118), (29, 0.039), (30, 0.027), (31, 0.02), (32, -0.025), (33, -0.053), (34, 0.017), (35, -0.005), (36, 0.057), (37, -0.089), (38, -0.091), (39, 0.076), (40, 0.036), (41, 0.049), (42, -0.019), (43, -0.062), (44, 0.024), (45, -0.016), (46, -0.018), (47, -0.01), (48, -0.031), (49, 0.066)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.93530673 127 nips-2012-Fast Bayesian Inference for Non-Conjugate Gaussian Process Regression

Author: Emtiyaz Khan, Shakir Mohamed, Kevin P. Murphy

Abstract: We present a new variational inference algorithm for Gaussian process regression with non-conjugate likelihood functions, with application to a wide array of problems including binary and multi-class classification, and ordinal regression. Our method constructs a concave lower bound that is optimized using an efficient fixed-point updating algorithm. We show that the new algorithm has highly competitive computational complexity, matching that of alternative approximate inference methods. We also prove that the use of concave variational bounds provides stable and guaranteed convergence – a property not available to other approaches. We show empirically for both binary and multi-class classification that our new algorithm converges much faster than existing variational methods, and without any degradation in performance. 1

2 0.82468081 55 nips-2012-Bayesian Warped Gaussian Processes

Author: Miguel Lázaro-gredilla

Abstract: Warped Gaussian processes (WGP) [1] model output observations in regression tasks as a parametric nonlinear transformation of a Gaussian process (GP). The use of this nonlinear transformation, which is included as part of the probabilistic model, was shown to enhance performance by providing a better prior model on several data sets. In order to learn its parameters, maximum likelihood was used. In this work we show that it is possible to use a non-parametric nonlinear transformation in WGP and variationally integrate it out. The resulting Bayesian WGP is then able to work in scenarios in which the maximum likelihood WGP failed: Low data regime, data with censored values, classification, etc. We demonstrate the superior performance of Bayesian warped GPs on several real data sets.

3 0.79554129 33 nips-2012-Active Learning of Model Evidence Using Bayesian Quadrature

Author: Michael Osborne, Roman Garnett, Zoubin Ghahramani, David K. Duvenaud, Stephen J. Roberts, Carl E. Rasmussen

Abstract: Numerical integration is a key component of many problems in scientific computing, statistical modelling, and machine learning. Bayesian Quadrature is a modelbased method for numerical integration which, relative to standard Monte Carlo methods, offers increased sample efficiency and a more robust estimate of the uncertainty in the estimated integral. We propose a novel Bayesian Quadrature approach for numerical integration when the integrand is non-negative, such as the case of computing the marginal likelihood, predictive distribution, or normalising constant of a probabilistic model. Our approach approximately marginalises the quadrature model’s hyperparameters in closed form, and introduces an active learning scheme to optimally select function evaluations, as opposed to using Monte Carlo samples. We demonstrate our method on both a number of synthetic benchmarks and a real scientific problem from astronomy. 1

4 0.75265032 272 nips-2012-Practical Bayesian Optimization of Machine Learning Algorithms

Author: Jasper Snoek, Hugo Larochelle, Ryan P. Adams

Abstract: The use of machine learning algorithms frequently involves careful tuning of learning parameters and model hyperparameters. Unfortunately, this tuning is often a “black art” requiring expert experience, rules of thumb, or sometimes bruteforce search. There is therefore great appeal for automatic approaches that can optimize the performance of any given learning algorithm to the problem at hand. In this work, we consider this problem through the framework of Bayesian optimization, in which a learning algorithm’s generalization performance is modeled as a sample from a Gaussian process (GP). We show that certain choices for the nature of the GP, such as the type of kernel and the treatment of its hyperparameters, can play a crucial role in obtaining a good optimizer that can achieve expertlevel performance. We describe new algorithms that take into account the variable cost (duration) of learning algorithm experiments and that can leverage the presence of multiple cores for parallel experimentation. We show that these proposed algorithms improve on previous automatic procedures and can reach or surpass human expert-level optimization for many algorithms including latent Dirichlet allocation, structured SVMs and convolutional neural networks. 1

5 0.72728842 37 nips-2012-Affine Independent Variational Inference

Author: Edward Challis, David Barber

Abstract: We consider inference in a broad class of non-conjugate probabilistic models based on minimising the Kullback-Leibler divergence between the given target density and an approximating ‘variational’ density. In particular, for generalised linear models we describe approximating densities formed from an affine transformation of independently distributed latent variables, this class including many well known densities as special cases. We show how all relevant quantities can be efficiently computed using the fast Fourier transform. This extends the known class of tractable variational approximations and enables the fitting for example of skew variational densities to the target density. 1

6 0.69349438 129 nips-2012-Fast Variational Inference in the Conjugate Exponential Family

7 0.67405379 58 nips-2012-Bayesian models for Large-scale Hierarchical Classification

8 0.64852923 233 nips-2012-Multiresolution Gaussian Processes

9 0.64022028 270 nips-2012-Phoneme Classification using Constrained Variational Gaussian Process Dynamical System

10 0.62493008 187 nips-2012-Learning curves for multi-task Gaussian process regression

11 0.6242246 355 nips-2012-Truncation-free Online Variational Inference for Bayesian Nonparametric Models

12 0.60254973 354 nips-2012-Truly Nonparametric Online Variational Inference for Hierarchical Dirichlet Processes

13 0.57788533 74 nips-2012-Collaborative Gaussian Processes for Preference Learning

14 0.57180709 359 nips-2012-Variational Inference for Crowdsourcing

15 0.56099832 206 nips-2012-Majorization for CRFs and Latent Likelihoods

16 0.5273456 54 nips-2012-Bayesian Probabilistic Co-Subspace Addition

17 0.49619207 268 nips-2012-Perfect Dimensionality Recovery by Variational Bayesian PCA

18 0.49232978 246 nips-2012-Nonparametric Max-Margin Matrix Factorization for Collaborative Prediction

19 0.49005333 244 nips-2012-Nonconvex Penalization Using Laplace Exponents and Concave Conjugates

20 0.48676559 342 nips-2012-The variational hierarchical EM algorithm for clustering hidden Markov models


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(0, 0.02), (21, 0.014), (38, 0.189), (42, 0.037), (54, 0.016), (55, 0.037), (65, 0.223), (74, 0.036), (76, 0.178), (80, 0.1), (92, 0.069)]

similar papers list:

simIndex simValue paperId paperTitle

1 0.87247401 85 nips-2012-Convergence and Energy Landscape for Cheeger Cut Clustering

Author: Xavier Bresson, Thomas Laurent, David Uminsky, James V. Brecht

Abstract: This paper provides both theoretical and algorithmic results for the 1 -relaxation of the Cheeger cut problem. The 2 -relaxation, known as spectral clustering, only loosely relates to the Cheeger cut; however, it is convex and leads to a simple optimization problem. The 1 -relaxation, in contrast, is non-convex but is provably equivalent to the original problem. The 1 -relaxation therefore trades convexity for exactness, yielding improved clustering results at the cost of a more challenging optimization. The first challenge is understanding convergence of algorithms. This paper provides the first complete proof of convergence for algorithms that minimize the 1 -relaxation. The second challenge entails comprehending the 1 energy landscape, i.e. the set of possible points to which an algorithm might converge. We show that 1 -algorithms can get trapped in local minima that are not globally optimal and we provide a classification theorem to interpret these local minima. This classification gives meaning to these suboptimal solutions and helps to explain, in terms of graph structure, when the 1 -relaxation provides the solution of the original Cheeger cut problem. 1

2 0.86845672 336 nips-2012-The Coloured Noise Expansion and Parameter Estimation of Diffusion Processes

Author: Simon Lyons, Amos J. Storkey, Simo Särkkä

Abstract: Stochastic differential equations (SDE) are a natural tool for modelling systems that are inherently noisy or contain uncertainties that can be modelled as stochastic processes. Crucial to the process of using SDE to build mathematical models is the ability to estimate parameters of those models from observed data. Over the past few decades, significant progress has been made on this problem, but we are still far from having a definitive solution. We describe a novel method of approximating a diffusion process that we show to be useful in Markov chain Monte-Carlo (MCMC) inference algorithms. We take the ‘white’ noise that drives a diffusion process and decompose it into two terms. The first is a ‘coloured noise’ term that can be deterministically controlled by a set of auxilliary variables. The second term is small and enables us to form a linear Gaussian ‘small noise’ approximation. The decomposition allows us to take a diffusion process of interest and cast it in a form that is amenable to sampling by MCMC methods. We explain why many state-of-the-art inference methods fail on highly nonlinear inference problems, and we demonstrate experimentally that our method performs well in such situations. Our results show that this method is a promising new tool for use in inference and parameter estimation problems. 1

same-paper 3 0.85264719 127 nips-2012-Fast Bayesian Inference for Non-Conjugate Gaussian Process Regression

Author: Emtiyaz Khan, Shakir Mohamed, Kevin P. Murphy

Abstract: We present a new variational inference algorithm for Gaussian process regression with non-conjugate likelihood functions, with application to a wide array of problems including binary and multi-class classification, and ordinal regression. Our method constructs a concave lower bound that is optimized using an efficient fixed-point updating algorithm. We show that the new algorithm has highly competitive computational complexity, matching that of alternative approximate inference methods. We also prove that the use of concave variational bounds provides stable and guaranteed convergence – a property not available to other approaches. We show empirically for both binary and multi-class classification that our new algorithm converges much faster than existing variational methods, and without any degradation in performance. 1

4 0.80409288 104 nips-2012-Dual-Space Analysis of the Sparse Linear Model

Author: Yi Wu, David P. Wipf

Abstract: Sparse linear (or generalized linear) models combine a standard likelihood function with a sparse prior on the unknown coefficients. These priors can conveniently be expressed as a maximization over zero-mean Gaussians with different variance hyperparameters. Standard MAP estimation (Type I) involves maximizing over both the hyperparameters and coefficients, while an empirical Bayesian alternative (Type II) first marginalizes the coefficients and then maximizes over the hyperparameters, leading to a tractable posterior approximation. The underlying cost functions can be related via a dual-space framework from [22], which allows both the Type I or Type II objectives to be expressed in either coefficient or hyperparmeter space. This perspective is useful because some analyses or extensions are more conducive to development in one space or the other. Herein we consider the estimation of a trade-off parameter balancing sparsity and data fit. As this parameter is effectively a variance, natural estimators exist by assessing the problem in hyperparameter (variance) space, transitioning natural ideas from Type II to solve what is much less intuitive for Type I. In contrast, for analyses of update rules and sparsity properties of local and global solutions, as well as extensions to more general likelihood models, we can leverage coefficient-space techniques developed for Type I and apply them to Type II. For example, this allows us to prove that Type II-inspired techniques can be successful recovering sparse coefficients when unfavorable restricted isometry properties (RIP) lead to failure of popular ℓ1 reconstructions. It also facilitates the analysis of Type II when non-Gaussian likelihood models lead to intractable integrations. 1

5 0.78654116 361 nips-2012-Volume Regularization for Binary Classification

Author: Koby Crammer, Tal Wagner

Abstract: We introduce a large-volume box classification for binary prediction, which maintains a subset of weight vectors, and specifically axis-aligned boxes. Our learning algorithm seeks for a box of large volume that contains “simple” weight vectors which most of are accurate on the training set. Two versions of the learning process are cast as convex optimization problems, and it is shown how to solve them efficiently. The formulation yields a natural PAC-Bayesian performance bound and it is shown to minimize a quantity directly aligned with it. The algorithm outperforms SVM and the recently proposed AROW algorithm on a majority of 30 NLP datasets and binarized USPS optical character recognition datasets. 1

6 0.78562671 179 nips-2012-Learning Manifolds with K-Means and K-Flats

7 0.78403473 226 nips-2012-Multiclass Learning Approaches: A Theoretical Comparison with Implications

8 0.78386194 319 nips-2012-Sparse Prediction with the $k$-Support Norm

9 0.78305268 34 nips-2012-Active Learning of Multi-Index Function Models

10 0.7825163 335 nips-2012-The Bethe Partition Function of Log-supermodular Graphical Models

11 0.78239381 199 nips-2012-Link Prediction in Graphs with Autoregressive Features

12 0.78226066 139 nips-2012-Fused sparsity and robust estimation for linear models with unknown variance

13 0.78098869 227 nips-2012-Multiclass Learning with Simplex Coding

14 0.78061146 148 nips-2012-Hamming Distance Metric Learning

15 0.78050715 292 nips-2012-Regularized Off-Policy TD-Learning

16 0.78047192 252 nips-2012-On Multilabel Classification and Ranking with Partial Feedback

17 0.7797094 187 nips-2012-Learning curves for multi-task Gaussian process regression

18 0.77966636 333 nips-2012-Synchronization can Control Regularization in Neural Systems via Correlated Noise Processes

19 0.77928263 186 nips-2012-Learning as MAP Inference in Discrete Graphical Models

20 0.77915817 16 nips-2012-A Polynomial-time Form of Robust Regression