b

DiscoverSearch
About
My stuff
Boltzmann machines and energy-based models
2017·arXiv
Abstract
Abstract

We review Boltzmann machines and energy-based models. A Boltzmann machine defines a probability distribution over binary-valued patterns. One can learn parameters of a Boltzmann machine via gradient based approaches in a way that log likelihood of data is increased. The gradient and Hessian of a Boltzmann machine admit beautiful mathematical representations, although computing them is in general intractable. This intractability motivates approximate methods, including Gibbs sampler and contrastive divergence, and tractable alternatives, namely energy-based models.

The Boltzmann machine has received considerable attention particularly after the publication of the seminal paper by Hinton and Salakhutdinov on autoencoder with stacked restricted Boltzmann machines [21], which leads to today’s success of and expectation to deep learning [54, 13] as well as a wide range of applications of Boltzmann machines such as collaborative filtering [1], classification of images and documents [33], and human choice [45, 47]. The Boltzmann machine is a stochastic (generative) model that can represent a probability distribution over binary patterns and others (see Section 2). The stochastic or generative capability of the Boltzmann machine has not been fully exploited in today’s deep learning. For further advancement of the field, it is important to understand basics of the Boltzmann machine particularly from probabilistic perspectives. In this paper, we review fundamental properties of the Boltzmann machines with particular emphasis on probabilistic representations that allow intuitive interpretations in terms of probabilities.

A core of this paper is in the learning rules based on gradients or stochastic gradients for Boltzmann machines (Section 3-Section 4). These learning rules maximize the log likelihood of given dataset or minimize the Kullback-Leibler (KL) divergence to a target distribution. In particular, Boltzmann machines admit concise mathematical representations for its gradients and Hessians. For example, Hessians can be represented with covariance matrices.

The exact learning rules, however, turn out to be computationally intractable for general Boltzmann machines. We then review approximate learning methods such as Gibbs sampler and contrastive divergence in Section 5.

We also review other models that are related to the Boltzmann machine in Section 6. For example, the Markov random field is a generalization of the Boltzmann machine. We also discuss how to deal with real valued distributions by modifying the Boltzmann machine.

The intractability of exact learning of the Boltzmann machine motivates tractable energy-based learning. Some of the approximate learning methods for the Boltzmann machine may be considered as a form of energy-based learning. As a practical example, we review an energy-based model for face detection in Section 7.

This survey paper is based on a personal note prepared for the first of the four parts of a tutorial given at the 26th International Joint Conference on Artificial Intelligence (IJCAI-17) held in

image

Figure 1: A Boltzmann machine.

Melbourne, Australia on August 21, 2017. See a tutorial webpage1for information about the tutorial. A survey corresponding to the third part of the tutorial (Boltzmann machines for time-series) can be found in [43].

A Boltzmann machine is a network of units that are connected to each other (see Figure 1). Let N be the number of units. Each unit takes a binary value (0 or 1). Let  Xibe the random variable representing the value of the i-th unit for  i ∈[1, N]. We use a column vector X to denote the random values of the N units. The Boltzmann machine has two types of parameters: bias and weight. Let  bibe the bias for the i-th unit for  i ∈[1, N], and let  wi,jbe the weight between unit i and unit j for (i, j) ∈[1, N −1]  × [i+ 1, N]. We use a column vector b to denote the bias for all units and a matrix W to denote the weight for all pairs of units. Namely, the (i, j)-the element of W is  wi,j. We let wi,j= 0 for  i ≥ jand for the pair of units (i, j) that are disconnected each other. The parameters are collectively denoted by

image

From the energy, the Boltzmann machine defines the probability distribution over binary patterns as follows:

image

where the summation with respect to  ˜xis over all of the possible N bit binary values. Namely, the higher the energy of a pattern x, the less likely that the x is generated. For a moment, we do not

image

Figure 2: Boltzmann machines with hidden units, input, and output

address the computational aspect of the denominator, which involves a summation of 2Nterms. This denominator is also known as a partition function:

image

A Boltzmann machine can be used to model the probability distribution,  Ptarget(·), of target patterns. Namely, by optimally setting the values of  θ, we approximate  Ptarget(·) with  Pθ(·). Here, some of the units of the Boltzmann machine are allowed to be hidden, which means that those units do not directly correspond to the target patterns (see Figure 2b). The units that directly correspond to the target patterns are called visible. The primary purpose of the hidden units is to allow particular dependency between visible units, which cannot be represented solely with visible units. The visible units may be divided into input and output (see Figure 2c). Then the Boltzmann machine can be used to model the conditional distribution of the output patterns given an input pattern.

Now we consider the problem of optimally setting the values of  θin a way that  Pθ(·) best approximates a given  Ptarget(·). Specifically, we seek to minimize the Kullback-Leibler (KL) divergence from  Pθto Ptarget[2]:

image

The first term of (7) is independent of  θ. It thus suffices to maximize the negation of the second term:

image

A special case of  Ptargetis the empirical distribution of the patterns in a given training dataset:

image

where D is the number of the patterns in D, Then the objective function (8) becomes

image

which is the log-likelihood of D with respect to  Pθwhen multiplied by D. By defining

image

we can represent  f(θ) as follows:

image

To find the optimal values of  θ, we take the gradient of  f(θ) with respect to  θ:

image

3.1 All of the units are visible

We start with the simplest case where all of the units are visible (see Figure 2a). Then the energy of the Boltzmann machine is simply given by (3), and the probability distribution is given by (4).

3.1.1 Gradient

We will derive a specific representation of  ∇log  Pθ(x) to examine the form of  ∇f(θ) in this case:

image

where the summation with respect to  ˆxis over all of the possible binary patterns, similar to the summation with respect to  ˜x. Here, (18) follows from (4) and (17). Plugging the last expression into (14), we obtain

image

The last expression allows intuitive interpretation of a gradient-based method for increasing the value of  f(θ):

image

where  ηis the learning rate (or the step size). Namely, for each pattern  ˜x, we compare  Pθ(˜x) against Ptarget(˜x). If  Pθ(˜x) is greater than  Ptarget(˜x), we update  θin a way that it increases the energy  Eθ(˜x)

so that the  ˜xbecomes less likely to be generated with  Pθ. If  Pθ(˜x) is smaller than  Ptarget(˜x), we update  θin a way that  Eθ(˜x) decreases. We will also write (20) as follows:

image

where  Etarget[·] is the expectation with respect to  Ptarget, Eθ[·] is the expectation with respect to  Pθ, and X is the vector of random variables denoting the values of the N units. Note that the expression of the gradient in (23) holds for any form of energy, as long as the energy is used to define the probability as in (4).

Now we take into account the specific form of the energy given by (3). Taking the derivative with respect to each parameter, we obtain

image

for  i ∈[1, N] and (i, j) ∈[1, N −1]  × [i+ 1, N]. From (23), we then find

image

where  Xiis the random variable denoting the value of the i-th unit for each  i ∈[1, N]. Notice that the expected value of  Xiis the same as the probability of  Xi= 1, because  Xiis binary. In general, exact evaluation of  Eθ[Xi] or  Eθ[Xi Xj] is computationally intractable, but we will not be concerned with this computational aspect until Section 5.

A gradient ascent method is thus to iteratively update the parameters as follows:

image

for  i ∈[1, N] and (i, j) ∈[1, N −1]  × [i+ 1, N]. Intuitively,  bicontrols how likely that the i-th unit takes the value 1, and  wi,jcontrols how likely that the i-th unit and the j-th unit simultaneously take the value 1. For example, when  Eθ[Xi Xj] is smaller than  Etarget[Xi Xj], we increase  wi,jto increase Eθ[Xi Xj]. This form of learning rule appears frequently in the context of Boltzmann machines. Namely, we compare our prediction  Eθ[·] against the target  Etarget[·] and update  θin a way that  Eθ[·] gets closer to  Etarget[·].

3.1.2 Stochastic gradient

We now rewrite (20) as follows:

image

Namely,  ∇f(θ) is given by the expected value of  −∇Eθ(X) +  Eθ [∇Eθ(X)], where the first X is distributed with respect to  Ptarget. Recall that the second term is an expectation with respect to  Pθ. This suggests stochastic gradient methods [3, 28, 9, 61, 50]. At each step, we sample a pattern  X(ω) according to  Ptargetand update  θaccording to the stochastic gradient:

image

image

Figure 3: How the energy is updated

where

image

When the target distribution is the empirical distribution given by the training data D, we only need to take a sample  X(ω) from D uniformly at random.

The stochastic gradient method based on (31)-(32) allows intuitive interpretation. At each step, we sample a pattern according to the target distribution (or from the training data) and update  θin a way that the energy of the sampled pattern is reduced. At the same time, the energy of every pattern is increased, where the amount of the increase is proportional to the probability for the Boltzmann machine with the latest parameter  θto generate that pattern (see Figure 3).

Taking into account the specific form of the energy given by (3), we can derive the specific form of the stochastic gradient:

image

which suggests a stochastic gradient method of iteratively sampling a pattern  X(ω) according to the target probability distribution and updating the parameters as follows:

image

for  i ∈[1, N] and (i, j) ∈[1, N −1]  × [i+ 1, N].

3.1.3 Giving theoretical foundation for Hebb’s rule

The learning rule of (36) has a paramount importance of providing a theoretical foundation for Hebb’s rule of learning in biological neural networks [16]:

When an axon of cell A is near enough to excite a cell B and repeatedly or persistently takes part in firing it, some growth process or metabolic change takes place in one or both cells such that A’s efficiency, as one of the cells firing B, is increased.

In short, “neurons wire together if they fire together” [37]. A unit of a Boltzmann machine corresponds to a neuron, and  Xi= 1 means that the i-th neuron fires. When two neurons, i and j, fire (Xi(ω) Xj(ω) = 1), the wiring weight  wi,jbetween the two

neurons gets stronger according to (36). Here, notice that we have 0  < Eθ[Xi Xj] <1 as long as the values of  θare finite.

The learning rule of the Boltzmann machine also involves a mechanism beyond what is suggested by Hebb’s rule. Namely, the amount of the change in  wi,jwhen the two neurons (i and j) fire depends on how likely those two neurons fire according to  Pθat that time. More specifically, it the two neurons are already expected to fire together (i.e., Eθ[Xi Xj] ≈1), we increase  wi,jonly by a small amount (i.e., η(1  − Eθ[Xi Xj])) even if the two neurons fire together (i.e., Xi(ω) Xj(ω) = 1).

Without this additional term (−Eθ[Xi] or  −Eθ[Xi Xj]) in (36), all of the parameters monotonically increases. If  Xi= 1 with nonzero probability in  Ptarget, then  bidiverges to  ∞almost surely. Otherwise, bistays unchanged from the initial value. Likewise, if  Xi Xj= 1 with nonzero probability in  Ptarget, then  wi,jdiverges to  ∞almost surely. Otherwise,  wi,jstays unchanged.

What is important is that this additional term is formally derived instead of being introduced in an ad hoc manner. Specifically, the learning rule is derived from a stochastic model (i.e., a Boltzmann machine) and an objective function (i.e., minimizing the KL divergence to the target distribution or maximizing the log-likelihood of training data) by taking the gradient with respect to the parameters.

3.1.4 Hessian

We now derive the Hessian of  f(θ) to examine its landscape. Starting from the expression in (27), we obtain

image

where the last expression is obtained from (18) and (25). The last expression consists of expectations with respect to  Pθand can be represented conveniently as follows:

image

where  COVθ[A, B] denotes the covariance between random variables A and B with respect to  Pθ. Likewise, we have

image

Therefore, the Hessian of  f(θ) is a covariance matrix:

image

where we use  COVθto denote a covariance matrix with respect to  Pθ. When  θis finite, this covariance matrix is positive semidefinite, and  f(θ) is concave. This justifies (stochastic) gradient based approaches to optimizing  θ. This concavity has been known [19], but I am not aware of the literature that explicitly represent the Hessian with a covariance matrix.

3.1.5 Summary

Consider a Boltzmann machine with parameters  θ= (b, W). When all of the N units of the Boltzmann machine are visible, the Boltzmann machine defines a probability distribution  Pθof N-bit binary patterns by

image

where the energy is

image

The KL divergence from  Pθto  Ptargetcan be minimized (or the log-likelihood of the target data having the empirical distribution  Ptargetcan be maximized) by maximizing

image

The gradient and the Hessian of  f(θ) is given by

image

where S denotes the vector of the random variables representing the value of a unit or the product of the values of a pair of units:

image

3.2 Some of the units are hidden

In this section, we consider Boltzmann machines that have both visible units and hidden units. Let N be the number of visible units and M be the number of hidden units.

3.2.1 Necessity of hidden units

We first study the necessity of hidden units [2]. The Boltzmann machine with N units have

image

parameters. This Boltzmann machine is used to model N-bit binary patterns. There are 2Npossible N-bit binary patterns, and the general distribution of N-bit patterns assigns probabilities to those 2N

patterns. We need

image

parameters to characterize this general distribution.

The number of parameters of the Boltzmann machine is smaller than the number of parameters needed to characterize the general distribution as long as N > 2. This suggests that the probability distribution that can be represented by the Boltzmann machine is limited. One way to extend the flexibility of the Boltzmann machine is the use of hidden units.

3.2.2 Free energy

Let x denote the visible values (i.e., the values of visible units), h denote the hidden values, and (x, h) denote the values of all units. We write the marginal probability distribution of the visible values as follows:

image

where the summation is over all of the possible binary patterns of the hidden values, and

image

Here, we write energy as follows:

image

Now, we define free energy as follows:

image

We can then represent  Pθ(x) in a way similar to the case where all of the units are visible, replacing energy with free energy:

image

3.2.3 Gradient

In (20), we simply replace energy with free energy to obtain the gradient of our objective function when some of the units are hidden:

image

What we then need is the gradient of free energy:

image

where  Pθ(h | x) is the conditional probability that the hidden values are h given that the visible values are x:

image

Observe in (66) that the gradient of free energy is expected gradient of energy, where the expectation is with respect to the conditional distribution of hidden values given the visible values. We thus obtain

image

The first term in the last expression (except the minus sign) is the expected value of the gradient of energy, where the expectation is with respect to the distribution defined with  Ptargetand  Pθ. Specifically, the visible values follow  Ptarget, and given the visible values, x, the hidden values follow Pθ(· | x). We will write this expectation with  Etarget [Eθ[· | X]]. The second term is expectation with respect to  Pθ, which we denote with  Eθ[·]. Because the energy (56) has the form equivalent to (47), ∇f(θ) can then be represented analogously to (49):

image

where X is the vector of random values of the visible units, H is the vector of  Hifor  i ∈[1, N], and S is defined analogously to (51) for all of the (visible or hidden) units:

image

where  Ui ≡ Xifor  i ∈[1, N], and  UN+i ≡ Hifor  i ∈[1, M], where  Hiis the random variable denoting the i-th hidden value.

3.2.4 Stochastic gradient

The expression with (72) suggests stochastic gradient analogous to (31)-(32). Observe that  ∇f(θ) can be represented as

image

where  Eθ [∇Eθ(X, H)] is the expected value of the gradient of the energy when both visible values and hidden values follow  Pθ, and  Eθ [∇Eθ(˜x, H) | ˜x] is the corresponding conditional expectation when the hidden values follow  Pθ(· | ˜x) given the visible values  ˜x.

A stochastic gradient method is then to sample visible values,  X(ω), according to  Ptargetand update  θaccording to the stochastic gradient:

image

By taking into account the specific form of the energy, we find the following specific update rule:

image

where each unit (i or j) may be either visible or hidden. Specifically, let M be the number of hidden units and N be the number of visible units. Then (i, j) ∈[1, N + M −1]  × [i+ 1, N + M]. Here,  Uidenotes the value of the i-th unit, which may be visible or hidden. When the i-th unit is visible, its expected value is simply  Eθ[Ui | X(ω)] =  Xi(ω), and  Eθ[Ui Uj | X(ω)] =  Xi(ω) Eθ[Uj | X(ω)]. When both i and j are visible, we have  Eθ[Ui Uj | X(ω)] =  Xi(ω)Xj(ω).

Namely, we have

image

for a visible unit  i ∈[1, N],

image

for a hidden unit  i ∈ [N+ 1, N + M],

image

for a pair of visible units (i, j) ∈[1, N −1]  × [i+ 1, N],

image

for a pair of a visible unit and a hidden unit (i, j) ∈[1, N] × [N+ 1, N + M], and

image

for a pair of hidden units (i, j) ∈ [N+ 1, N + M −1]  × [i+ 1, N + M].

3.2.5 Hessian

We now derive the Hessian of  f(θ) when some of the units are hidden. From the gradient of  f(θ) in (72), we can write the partial derivatives as follows:

image

where recall that  ui ≡ xifor  i ∈[1, N] and  uN+i ≡ hifor  i ∈[1, M]. When all of the units are visible, the first term in (27) is the expectation with respect to the target distribution and is independent of  θ. Now that some of the units are hidden, the corresponding first term in (85) depends on  θ. Here, the first term is expectation, where the visible units follow the target distribution, and the hidden units follow the conditional distribution with respect to the Boltzmann machine with parameters  θgiven the visible values. Notice that the second term in the right-hand side of (86) has the form equivalent to (38). Hence, we have

image

The first term in the right-hand side of (86) has a from similar to (38) with two differences. The first difference is that it is an expectation with respect to  Ptarget. The second difference is that the probability  Pθis conditional probability given the visible values. We now show that there exists a Boltzmann distribution that has  Pθ(· | x) as its probability distribution (see also Figure 4).

Lemma 1. Consider a Boltzmann machine having visible units and hidden units whose energy is given by (57). The conditional probability distribution of the hidden units given the visible values x is given by the probability distribution of a Boltzmann machine with bias b(x) and weight W(x) that are defined as follows:

image

Proof. Recall the expression of the conditional probability distribution in (67):

image

where we rewrite the energy in (57) as follows:

image

The first term in the right-hand side of the last expression is independent of h and canceled out between the numerator and the denominator in (90). We thus consider the Boltzmann machine with parameters  θ(x) ≡ (b(x), W(x)) as defined in (88)-(89). Then we have

image

which completes the proof of the lemma.

By (86) and Lemma 1, we can represent the second partial derivative as follows:

image

where  COVθ[·, · | X] denotes the conditional covariance with respect to  Pθ(X)(·) ≡ Pθ(· | X) given the visible values X.

image

Figure 4: The conditional distribution of h given x in (a) is equal to the distribution of h in (b), illustrating Lemma 1.

Therefore, the Hessian of  f(θ) is

image

where we define S as in (75), and  COVθ[·, · | X] denotes the conditional covariance matrix with respect to  Pθ(X)given the visible values X. In general, the Hessian is not positive semidefinite, and  f(θ) is not concave. Hence, (stochastic) gradient based approaches do not necessarily find globally optimal parameters.

3.2.6 Summary

Consider a Boltzmann machine with parameters  θ ≡ (b, W), where at least one of the units are hidden. The Boltzmann machine defines a probability distribution  Pθof the visible values x and the hidden values h by

image

where energy is given by

image

The marginal probability distribution of the visible values is

image

where free energy is defined as follows:

image

The KL divergence from  Pθto  Ptargetcan be minimized (or the log-likelihood of the target data having the empirical distribution  Ptargetcan be maximized) by maximizing

image

The gradient and the Hessian of  f(θ) are given by

image

where S denotes the vector of the random variables representing the value of a unit (Ui = Xifor i ∈[1, N] and  UN+i = Hifor  i ∈[1, M]) or the product of the values of a pair of units:

image

In this section, we study the Boltzmann machine whose visible units are divided into input units and output units (see Figure 2c). Such a Boltzmann machine is sometimes called a conditional Boltzmann machine [20]. Let  Ninbe the number of input units,  Noutbe the number of output units,  N = Nin+Noutbe the number of visible units, and M be the number of hidden units.

Such a Boltzmann machine can be used to model a conditional probability distribution of the output values y given the input values x. Let  Pθ(y | x) denote the conditional probability of y given x. When the Boltzmann machine has hidden units, we also write

image

4.1 Objective function

A training dataset for a discriminative model is a set of the pairs of input and output:

image

When this training dataset is given, a natural objective function corresponding to the log-likelihood (10) for the generative model is

image

This objective function may also be related to a KL divergence. Let  Ptarget(· | ·) be the target conditional distribution, which we seek to model with  Pθ(· | ·). Consider the following expected KL divergence from  Pθ(· | X) to  Ptarget(· | X), where X is the random variable denoting the input values that are distributed according to  Ptarget:

image

The first term of the last expression is independent of  θ. To minimize the expected KL divergence, it thus suffices to maximize the negation of the second term:

image

When  Ptargetis the empirical distribution of the pairs of input and output in the training dataset, (112) is reduced to (108).

4.2 Gradient, stochastic gradient, Hessian

Observe that the first term of (115) is equivalent to the objective function of the generative model where both of input and output are visible (i.e., input and output in (115) should be regarded as visible in (102)). The second term is analogous to the first term, but now only the input should be regarded as visible (i.e., output and hidden in (115) should be regarded as hidden in (102)).

The gradient thus follows from (102):

image

where S denotes the vector of random variables representing the value of a unit or the product of the values of a pair of units:

image

where  Ui ≡ Xifor  i ∈[1, Nin], UNin+i ≡ Yifor  i ∈[1, Nout], and  UN+i ≡ Hifor  i ∈[1, M]. Because (117) is expressed as an expectation with respect to  Ptarget, it directly gives the following stochastic gradient:

image

where (X(ω), Y(ω)) is sampled according to  Ptarget. Specifically, we have the following learning rule of a stochastic gradient method:

image

for an output unit  i ∈ [Nin+ 1, N],

image

for a hidden unit  i ∈ [N+ 1, N + M],

image

for a pair of an input unit and an output unit (i, j) ∈[1, Nin] × [Nin+ 1, N],

image

for a pair of an input unit and a hidden unit (i, j) ∈[1, Nin] × [N+ 1, N + M],

image

for a pair of output units (i, j) ∈ [Nin+ 1, N −1]  × [i+ 1, N],

image

for a pair of an output unit and a hidden unit (i, j) ∈ [Nin+ 1, N] × [N+ 1, N + M], and

image

for an input unit  i ∈[1, Nin] and

image

for a pair of input units (i, j) ∈[1, Nin −1]  × [i+ 1, Nin]. One can easily see that these parameters are redundant and do not play any role in  Pθ(y | x). The Hessian of the objective of discriminative learning follows from (103) and (115):

image

where  COVθ[· | X, Y] denotes the conditional covariance matrix with respect to the conditional distribution of the hidden values  Pθ(· | X, Y) given (X, Y), and  COVθ[· | X] denotes the conditional covariance matrix with respect to the conditional distribution of the output and hidden values  Pθ(· | X) given X.

4.2.1 Summary

Consider a Boltzmann machine with parameters  θ= (b, W), where the units can be classified into input, output, and hidden. The Boltzmann machine defines a conditional probability distribution of the output values y given the input values x:

image

where  Pθ(x, y, h) is the probability distribution of the Boltzmann machine where (x, y) is visible, and h is hidden.

The expected KL divergence from  Pθ(· | X) to  Ptarget(· | X), where the expectation is with respect to the target distribution of the input values X, can be minimized (or the sum of the conditional log-likelihood of the output values given the input values, where the input and output follow the empirical distribution  Ptarget, can be maximized) by maximizing

image

image

Figure 5: A discriminative model with a Boltzmann machine with no hidden units and no weight between output units.

The gradient and the Hessian of  f(θ) are given by

image

where S denotes the vector of the random variables representing the value of a unit (Ui = Xifor i ∈[1, Nin], UNin+i = Yifor  i ∈[1, Nout], and  UN+i = Hifor  i ∈[1, M]) or the product of the values of a pair of units:

image

4.3 Simplest case with relation to logistic regression

Here, we study the simplest but non-trivial case of a discriminative model of a Boltzmann machine. Specifically, we assume that the Boltzmann machine has no hidden units, and there are no weight between output units. As we have discussed in Section 4, the bias associated with the input unit and the weight between input units are redundant and do not play any role. Therefore, it suffices to consider the bias for output units and the weight between input units and output units (see Figure 5). Let  bObe the bias associated with output units and  WIObe the weight matrix where the (i, j) element denotes the weight between the i-th input unit and the j-th output unit.

4.3.1 Conditional probability

We first apply Lemma 1, where the input units correspond to the visible units in the lemma, and the output units correspond to the hidden units in the lemma. Then we can see that the probability distribution of output values given input values x is given by a Boltzmann machine with the following bias and no weight:

image

The conditional probability of output values y given input values x is thus given by

image

In this case, the partition function, which consists of 2Noutterms, can be computed in  O(Nout) time:

Lemma 2. The partition function (the denominator) in the right-hand side of (137) can be written as follows:

image

where the summation with respect to  ˜yis over all of the possible binary patterns of length  Nout.

Proof.

image

Now, it can be easily shown that the output values are conditionally independent of each other given the input values:

Corollary 1. The conditional probability of output values (137) can be written as the product of the conditional probabilities of each output value:

image

where

image

Proof. By Lemma 2, we have

image

image

4.3.2 Relation to logistic regression

By (143), the probability that the i-th output unit takes the value 1 is

image

where  wiis the i-th column of  WIO. The last expression has the form of logistic regression, where the explanatory variables x are binary.

Due to the conditional independence shown in Corollary 1, we can say that the Boltzmann machine shown in Figure 5 consists of N independent models of logistic regression that have common explanatory variables.

When we train a Boltzmann machine with a stochastic gradient method, we need to evaluate expected values with respect to the distribution defined by the Boltzmann machine. Such expected values appear for example as  Eθ[S] in (49) and (102) or as  Eθ[S | X] in (133). In general, exact expressions for these expectations are unavailable in closed forms.

5.1 Gibbs sampler

A practical approach that can be used to estimate those expectations is Markov Chain Monte Carlo in general and Glauber dynamics [36] or Gibbs samplers [5] in particular. For example, we can sample K patterns according to the distribution given by a Boltzmann machine via a Gibbs sampler shown in Algorithm 1. The expected values can then be estimated using the K samples. In practice, we often ignore initial samples and consider only every nth sample for a sufficiently large n.

In Step 5 of Algorithm 1, the conditional distribution of  x(k)igiven  x(k−1)jfor  j ̸= iis defined by

image

for  x(k)i ∈ {0, 1}, where

image

5.2 Contrastive divergence

Another approach to deal with the computational intractability of evaluation the expectations is to avoid it. Namely, we modify our objective function.

Recall, from (7) and (10), that our objective function has been the KL divergence from  Pθto  Ptarget(or equivalently the log likelihood of data with respect to  Pθ). The gradient of the KL divergence with respect to  θinvolves the computationally intractable term of expectation with respect to  Pθ.

Here, we study an alternative objective function of Contrastive divergence [18, 4]. Consider a Gibbs sampler (Algorithm 1) that initializes the values by sampling from  Ptarget(or uniformly at random from the dataset D). Because this is the distribution at the beginning of the Gibbs sampler, we write P0 ≡ Ptarget. The distribution of the patterns sampled by the Gibbs sampler at step k is referred to as  Pθk. Because  Pθk → Pθas  k → ∞, we also write  Pθ∞ ≡ Pθ.The KL divergence in (7) can now be written as

image

where

image

and its gradient (recall (18)) as

image

The first term of the right-hand side of (152) is the expectation with respect to the target distribution P0, which is readily computable. The second term is the expectation with respect to the model (with parameter  θ), which is in general computationally intractable for a Boltzmann machine.

To cancel out this computationally intractable second term, consider the following contrastive divergence:

image

Here, we quote an intuitive motivation of the contrastive divergence from [18], using our notations (shown within [·]):

The intuitive motivation for using this “contrastive divergence” is that we would like the Markov chain that is implemented by Gibbs sampling to leave the initial distribution [  P0] over the visible variables unaltered. Instead of running the chain to equilibrium and comparing the initial and final derivatives we can simply run the chain for one full step and then update the parameters to reduce the tendency of the chain to wander away from the initial distribution on the first step. Because [  Pθ1] is one step closer to the equilibrium dis- tribution than [  P0], we are guaranteed that [ KL(P0 || Pθ∞) ] exceeds [ KL(Pθ1 || Pθ∞) ] unless [  P0] equals [  Pθ1], so the contrastive divergence can never be negative. Also, for Markov chains in which all transitions have non-zero probability, [  P0 = Pθ1] implies [  P0=  Pθ1] so the contrastive divergence can only be zero if the model is perfect.

The gradient of the second term of (153) can be derived as follows:

image

where

image

Hinton has empirically shown that  ε(θ) is small and recommended the approximation of  ε(θ) ≈0 in [18]. The first term of the right-hand side of (158) is an expectation with respect to  P0and can be readily evaluated. The second term is an expectation with respect to  Pθ1and can be estimated with the samples from the Gibbs sampler in one step.

In [18],  ε(θ) is represented in the following equivalent form:

image

5.2.1 Score matching (Fisher divergence)

A particularly interesting objective function for energy-based models is score matching [22], which has been subsequently used for example in [66, 29, 24, 30, 60, 62]. Similar to contrastive divergence, score matching is an objective function, which we can avoid computationally intractable evaluation of expectation with.

Specifically, Hyv¨arinen [22] defines score matching as

image

which is also referred to as Fisher divergence [38]. Note that the gradient is with respect to x and not θ. Hyv¨arinen shows that

image

is a consistent estimator [22]. A key property of score matching is that there is no need for calculating the partition function. In particular,

image

Also, although  ∇˜xlog  Ptarget(x) might appear to be intractable, it has been shown in [22] that

image

is asymptotically equivalent to FD(Ptarget || Pθ) as  D ≡ |D| → ∞, where D is the set of data (samples from  Ptarget), Nis the dimension of  x ∈ D, and const is the term independent of  θ.

A limitation of the estimator in (165) is that it assumes among others that the variables x are continuous and and  Pθ(·) is differentiable. There has been prior work for relaxing these assumptions. For example, Hyv¨arinen studies an extension for binary or non-negative data [23], and Kingma and LaCun study an approach of adding Gaussian noise to data samples for smoothness condition [29].

5.3 A separate generator

The learning rule that follows from maximization of loglikelihood (minimization of KL divergence) via a gradient ascent method may be considered as decreasing the energy of “positive” or “real” samples that are generated according to a target distribution  Ptargetand increasing the energy of “negative” or “fake” samples that are generated according to the current model (recall (18)):

image

With this learning rule, one can let the model to be able to better “discriminate between the positive examples from the original data and the negative examples generated by sampling from the current density estimate” [65].

Then an energy-based model may be considered as taking both the role of a generator and the role of a discriminator in a generative adversarial network (GAN) [53, 14]. There is a line of work [15, 8, 67, 26] that prepares a separate generator with an energy-based model. If the separate generator allows more efficient sampling than the energy-based model, the expectation with respect to the distribution of the current generator (i.e., �˜x Pθ(˜x) ∇θEθ(ˆx) in (166)) can be more efficiently evaluated [26, 15].

5.4 Mean-field Boltzmann machines

There are also approaches that avoid computationally expensive evaluation of expectation. An example is a mean-field Boltzmann machine [49], which can be used to approximate a Boltzmann machine. A mean-field Boltzmann machine ignores connections between units and chooses the real value  mifor each unit i in a way that the distribution defined through

image

well approximates the distribution  Pθ(x) of a Boltzmann machine in the sense of the KL divergence [63].

Here we review stochastic models that are related to the Boltzmann machine.

6.1 Markov random fields

A Boltzmann machine is a Markov random field [27] having a particular structure. A Markov random field consists of a finite set of units similar to a Boltzmann machine. Each unit of a Markov random field takes a value in a finite set. The probability distribution of the values (configurations) of the Markov random field can be represented as

image

where the summation with respect to  ˜xis over all of the possible configurations of the values in the finite set, for which the Markov random field is defined. The energy of a configuration is defined as follows:

image

where  φ(x) is a feature vector of x. A Markov random field is also called an undirected graphical model.

6.1.1 Boltzmann machine and Ising model

A Markov random field is reduced to a Boltzmann machine when it has the following two properties. First,  φ(x) is a vector of monomials of degrees up to 2. Second, each unit takes a binary value.

An Ising model [25] is essentially equivalent to a Boltzmann machine but the binary variable takes values in  {−1,+1}.

6.1.2 Higher-order Boltzmann machine

A higher-order Boltzmann machine extends a Boltzmann machine by allowing  φ(x) to include monomials of degree greater than 2 [55]. Each unit of a higher-order Boltzmann machine takes a binary value.

6.2 Determinantal point process

A determinantal point process (DPP) defines a probability distribution over the subsets of a given ground set [39, 32, 56, 57]. In our context, the ground set Y can be considered as the set of all units:

image

A subset X can be considered as a set of units that take the value 1:

image

A DPP can be characterized by a kernel L, which is an  N × Npositive semi-definite matrix. The probability that the subset X is selected (i.e., the units in X take the value 1) is then given by

image

where  LXdenotes the principal submatrix of L indexed by X, and I is the  N × Nidentity matrix. A DPP can be seen as an energy-based model, whose energy is given by E(X) =  −log det(LX). In general, the kernel L can be represented as

image

by the use of a  K × Nmatrix B, where K is the rank of L. For  i ∈[1, N], let  bibe the i-th column of B, which may be understood as a feature vector of the i-th item (unit). One can further decompose biinto  bi = qi φi, where  ||φi||= 1 and  qi ≥0. Then we can write the (i, j)-th element of L as follows:

image

In particular,

image

so that  qican be understood as the ”quality” of the i-th item. Specifically, given that exactly one item is selected, the conditional probability that the i-th item is selected is proportional to  q2i. Likewise,

image

where

image

may be understood as the similarity between item i and item j. Equation (176) implies that the similar items are unlikely to be selected together. In general, a DPP tends to give high probability to diverse subsets of items having high quality.

Finally, we discuss a computational aspect of a DPP. Let’s compare the denominator of the right-hand side of (172) against the corresponding denominator (partition function) of the Boltzmann machine in (4). The partition function of the Boltzmann machine is a summation over 2Nterms, which suggest that we need  O(2N) time to evaluate it. On the other hand, the determinant of an  N × Nmatrix can be computed in  O(N 3) time.

When L has a low rank (K < N), one can further reduce the computational complexity. Let

image

be the  K × Kpositive semi-definite matrix, defined from (173). Because the eigenvalues  {λ1, . . . , λK}of L are eigenvalues of C and vice versa, we have

image

where the first identity matrix I is  N ×N, and the second is  K×K. Therefore, (172) can be represented as follows:

image

The denominator of this dual representation of a DPP can be evaluated in  O(K3) time. The DPP has been receiving increasing attention in machine learning, and various learning algorithms have been proposed in the literature [12, 11, 40, 46].

6.3 Gaussian Boltzmann machines

Here we review energy based models that deal with real values with a particular focus on those models that are extended from Boltzmann machines. A standard approach to extend the Boltzmann machine to deal with real values is the use of a Gaussian unit [41, 64, 31].

6.3.1 Gaussian Bernoulli restricted Boltzmann machines

For real values  x ∈ RNand binary values  h ∈ {0, 1}M, Krizhevsky studies the following energy [31]:

image

where  θ ≡ (bV, bH, W, σ) is the set of parameters.

Krizhevsky shows that the conditional probability of x given h has a normal distribution, and the conditional probability of h given x has a Bernoulli distribution [31]. Here, we re-derive them with our notations and in a simpler manner.

Theorem 1. Consider the energy given by (181). Then the elements of x are conditionally independent of each other given h, and the elements of h are conditionally independent of each other given x. Also, the conditional probability density of  xigiven h is given by

image

for  xi ∈ R, and the conditional mass probability of  hjfor  hj ∈ {0, 1}given x is given by

image

Proof. First observe that

image

where

image

This means that the elements of x are conditionally independent of each other given h. The conditional independence of the elements of h given x can be shown analogously. Then we have

image

where

image

for  xi ∈ Rand  hj ∈ {0, 1}. By taking into account the normalization for the total probability to become 1, we obtain (182) and (183).

One might wonder why the product  σi wi,jappears in (182), and the quotient  wi,j/σiappears in (183). We can shed light on these expressions by studying natural parameters as in [64]. The natural parameters of the normal distribution with mean  µand standard deviation  σare  µ/σ2and  −1/(2 σ2). Hence, the natural parameters in (182) are

image

Likewise, the natural parameter in (183) is

image

Therefore, only the quotient  wi,j/σiappears in natural parameters.

6.3.2 Spike and slab restricted Boltzmann machines

Courville et al. study a class of particularly structured higher-order Boltzmann machines with Gaussian units, which they refer to as spike and slab RBMs [6, 7]. For example, the energy may be represented as follows:

image

where x denotes real-valued visible values, h denotes binary hidden values (“spikes”), S denotes real-valued “slabs,” and  θ ≡ (λ, α, W, b) denotes the parameters. The term  wi,j,k xi hj sj,krepresents a three way interactions.

Similar to Theorem 1, one can show that the elements of x are conditionally independent of each other and have normal distributions given h and S:

image

for  xi ∈ R. Likewise, the elements of S are conditionally independent of each other and have normal distributions given x and h:

image

for  sj,k ∈ R. Given x and S, the elements of h are conditionally independent of each other and have Bernoulli distributions:

image

for  hj ∈ {0, 1}. It has been argued and experimentally confirmed that spike and slab RBMs can generate images with sharper boundaries than those generated by models with binary hidden units [6, 7].

6.4 Using expected values to represent real values

Here, we discuss the approach of using expected values given by the probability distribution defined by a Boltzmann machine. Because a unit of a Boltzmann machine takes a binary value, 0 or 1, the expected value is in [0, 1]. With appropriate scaling, any closed interval can be mapped to [0, 1].

6.4.1 Expected values in visible units

Recall, from Section 3, that a Boltzmann machine with parameter  θ ≡ (b, W) defines a probability distribution over binary values:

image

where

image

for  x ∈ {0, 1}N. The expected values can then be given as

image

which take values in [0, 1]N.

A question is whether the values given by the expectation are suitable for a particular purpose. In other words, how should we set  θso that the expectation gives suitable values for the purpose under consideration?

Here, we discuss a particular learning rule of simply using the values in [0, 1]Nin the learning rules for binary values. In particular, a gradient ascent method for generative learning without hidden units is given by (28) and (29). A stochastic gradient method with mini-batches is then given by

image

where we take K samples,  X(ω1), . . . , X(ωK), at each step of the mini-batch stochastic gradient method. As  K → ∞, these learning rules converge to the gradient ascent method given by (28) and (29). If the real values under consideration are actually expectation of some binary random variables, we can justify a stochastic gradient method of updating  bibased on a sampled real value  Ri(ω) according to

image

However, the corresponding update of  wi,jcannot be justified, because  Eθ[Xi Xj] cannot be represented solely by  Eθ[Xi] and  Eθ[Xj] unless  Xiand  Xjare independent. If  Xiand  Xjare independent, we should have  wi,j= 0.

In general, the use of expected values can be justified only for the special case where the corresponding random variables are conditionally independent of each other given input values. The simple discriminative model in Figure 5 is an example of such a special case. In this case, one can apply a stochastic gradient method of updating the parameters as follows:

image

for a sampled pair (Xi(ω), Rj(ω)), where  Xi(ω) is an input binary value for  i ∈[1, Nin], and  Rj(ω) is an output real value in [0, 1] for  j ∈[1, Nout].

image

Figure 6: A restricted Boltzmann machine

6.4.2 Expected values in hidden units

Expected values are more often used for hidden units [58, 59, 10, 44] than for visible units.

Consider a Boltzmann machine with visible units and hidden units, which have no connections between visible units or between hidden units (namely, a restricted Boltzmann machine or RBM; see Figure 6). Let  bHbe the bias associated with hidden units,  bVbe the bias associated with visible units, and W be the weight between visible units and hidden units. Then, given the visible values x, the hidden values h are conditionally independent of each other, and we can represent the conditional expected value of the j-th hidden unit as follows:

image

where

image

Because the energy is a linear function of h, we can represent the expected energy, where the expectation is with respect to the conditional distribution of the hidden values H given the visible values x, using the conditional expected values of hidden units:

image

Notice that the distribution of visible values is then given by

image

This expected energy may be compared against the corresponding free energy:

image

where

image

The free energy can be represented as follows:

image

Theorem 2.

image

where  −Eθ[log  Pθ(H | x)] is the entropy of the conditional distribution of hidden values H given visible values x.

Proof. By Corollary 1, hidden values are conditionally independent of each other given visible values:

image

Then, using the notation in (206), we obtain

image

The theorem now follows by adding (209) and (220), comparing it against (215).

Figure 7 compares expected energy and free energy. Specifically, the blue curve shows the value of

image

which appears in the expression of expected energy (209), and the green curve shows the value of

image

which appears in the expression of free energy (215). The difference between the two curves is largest (log 2  ≈ 0.30) when  bHj+  x⊤W:,j= 0. The two curves are essentially indistinguishable when  |bHj+ x⊤W:,j|is sufficiently large. This suggests that the Boltzmann machine with expected energy is different from the corresponding Boltzmann machine (with free energy), but the two models have some similarity.

image

Figure 7: Comparison between expected energy (209) and free energy (215) associated with the j-th hidden unit of a restricted Boltzmann machine.

So far, we have studied probabilistic models. A difficulty that we often face with probabilistic models is in high computational complexity for evaluating the partition function, or the normalization for the total probability to become 1. The probabilistic models that we have studied can be turned into (non-probabilistic) energy-based models. Such energy-based models do not require normalization, which is attractive from computational point of view. A difficulty in energy-based models is in designing appropriate objective functions. In this section, we will study (non-probabilistic) energy-based models through an example [42]. For more details, see a tutorial by LeCun [34].

7.1 Objective functions for energy-based models

In learning an energy-based model, one needs to carefully design an objective (loss) function [34, 35] in a way that minimizing the objective function leads to desired results. We can then optimize the parameters  θof the energy-based model by minimizing the objective function. The energy-based model with the optimized parameters  θshould give low energy to desirable values of the variables of the energy-based model and high energy to other values.

Here, we consider energy-based models with input and output. Let  Eθ(x, y) be the energy for a pair of an input x and an output y. An energy-based model with parameter  θgives

image

as the output for input x. A desirable pair of input and output should give lower energy than undesirable ones.

When we optimize an energy-based model, minimizing the energy of a given data is usually inappropriate. In particular, such an objective function may be unbounded. It may not distinguish two patterns, one is good and the other is very good, as both of the two patterns have the minimum energy.

An objective function of an energy-based model should have a contrastive term, which naturally appear in the objective function of a probabilistic model (i.e., the KL divergence or the log likelihood). For example, recall from (4) that the average negative log likelihood of a set of patterns D with respect

image

Figure 4: Architecture of the Minimum Energy Machine.

1 Figure 8: An architecture of the energy (225): Figure 4 from [42], where W is used to denote parameters θ.

to a Boltzmann machine is given by

image

The second term of the right-hand side of (224) is a contrastive term. In particular, to minimize this objective function, we should not only reduce the energy of the patterns in D but also increase the energy of the patterns not in D. Recall also Figure 3. However, the contrastive term involves the summation over exponentially many (2N) patterns, and is the source of computational intractability. In designing an objective function for an energy-based model, we want to design a contrastive term that can be more efficiently evaluated. We will see an example in the following.

7.2 An example of face detection with pose estimation

Osadchy et al. study an energy-based approach for classifying images into “face” or “non-face” and estimating the facial pose at the same time [42]. Let x be a vector representing an image, y be a variable indicating “face” or “non-face,” and z be a vector representing a facial pose. They consider an energy function of the following form:

image

where  Gθ(·) is a convolutional neural network (CNN), having parameter  θ, that maps an x into a lower dimensional vector,  F(·) is an arbitrarily defined function that maps a z on to a manifold embedded in the low dimensional space of the output of the CNN, and T is a constant (see Figure 8).

An image x is classified as “face” (y = 1) if

image

which is equivalent to

image

We want to learn the parameters  θfrom a given training dataset. A training dataset D consists of two subsets. The first subset  D1includes facial images with their facial poses (z, x). The second subset  D2includes non-facial images x.

It would be computationally intractable to maximize the log-likelihood of the training dataset with respect to the probability distribution that could be defined with the energy (225) through

image

Osadchy et al. instead minimizes the following objective function [42]:

image

where  κis a positive constant (hyper-parameter). The first term of the right-hand side of (229) is the average squared energy of the “face” samples. Hence, minimizing the energy of “face” samples leads to minimizing the objective function. The second (contrastive) term involves the energy of “non-face” samples when they are classified as “face” (y = 1), where the face pose is set optimally so that the energy is minimized. By minimizing the second term, one can expect to make the energy of those “non-face” samples with the face label (y = 1) greater than the corresponding energy with the non-face label (y = 0).

We have reviewed basic properties of Boltzmann machines with a particular focus on those that are relevant for gradient-based approaches to learning their parameters. As it turns out that exact learning is in general intractable, we have discussed general approaches to approximation as well as tractable alternative, namely energy-based models.

This paper has covered limited perspectives of Boltzmann machines and energy-based models. For example, we have only briefly discussed restricted Boltzmann machines (RBMs), which deserves intensive review in its own. In fact, RBMs and deep Boltzmann machines [51] are the topics that are covered in the second part of an IJCAI-17 tutorial2. This paper does not cover Boltzmann machines for time-series, which are reviewed in a companion paper [43].

The use of energy is also restrictive in this paper and in line with the tutorial by LeCun [34]. Our perspective on energy-based learning in the tutorial is, however, broader than what is suggested by LeCun. We may use energy for other purposes. An example is the use of free energy [52, 17, 48] or expected energy [10] to approximate the Q-function in reinforcement learning. This energy-based reinforcement learning is the topic covered in the fourth part of the IJCAI-17 tutorial. A key aspect of energy-based reinforcement learning is that the energy used to approximate the Q-function naturally defines a probability distribution, which is used for exploration. Thus, energy-based models that allow efficient sampling have the advantage in sampling the actions to explore in reinforcement learning. There is recent work on the use of deep neural networks to approximate the Q-function in the framework of energy-based reinforcement learning [15], where they use a separate sampling network for efficiency.

This work was supported by JST CREST Grant Number JPMJCR1304, Japan. The author thanks Chen-Feng Liu and Samuel Laferriere for pointing out typographical errors in the original version.

[1] Restricted Boltzmann machines for collaborative filtering. In Proceedings of the 24th international conference on Machine learning (ICML 2017), pages 791–798, June 2007.

[2] D. H. Ackley, G. E. Hinton, and T. J. Sejnowski. A learning algorithm for Boltzmann machines. Cognitive Science, 9:147–169, 1985.

[3] L. Bottou. Online learning and stochastic approximations. In D. Saad, editor, On-Line Learning in Neural Networks, chapter 2, pages 9–42. Cambridge University Press, 2009.

[4] M. ´A. Carreira-Perpi˜n´an and G. E. Hinton. On contrastive divergence learning. In Proceedings of the tenth International Workshop on Artificial Intelligence and Statistics, pages 33–40, January 2005.

[5] G. Casella and E. I. George. Explaining the Gibbs sampler. The Americal Statistician, 46(3):167– 174, 1992.

[6] A. Courville, J. Bergstra, and Y. Bengio. A spike and slab restricted Boltzmann machine. In Proceedings of the 14th International Conference on Artificial Intelligence and Statistics, pages 233–241, 2011.

[7] A. Courville, J. Bergstra, and Y. Bengio. Unsupervised models of images by spike-and-slab RBMs. In Proceedings of the 28th International Conference on Machine Learning, pages 1145–1152, 2011.

[8] Z. Dai, A. Almahairi, P. Bachman, E. Hovy, and A. Courville. Calibrating energy-based generative adversarial networks. CoRR, abs/1702.01691, 2017.

[9] J. Duchi, E. Hazan, and Y. Singer. Adaptive subgradient methods for online learning and stochas- tic optimization. Journal of Machine Learning Research, 12:2121–2159, 2011.

[10] S. Elfwing, , E. Uchibe, and K. Doya. From free energy to expected energy: Improving energy- based value function approximation in reinforcement learning. Neural Networks, 84:17–27, 2016.

[11] M. Gartrell, U. Paquet, and N. Koenigstein. Low-rank factorization of determinantal point pro- cesses. In Proceedings of the Thirty-First AAAI Conference on Artificial Intelligence (AAAI-17), pages 1912–1918, February 2017.

[12] J. A. Gillenwater, A. Kulesza, E. Fox, and B. Taskar. Expectation-maximization for learning determinantal point processes. In Z. Ghahramani, M. Welling, C. Cortes, N. D. Lawrence, and K. Q. Weinberger, editors, Advances in Neural Information Processing Systems 27, pages 3149– 3157. Curran Associates, Inc., 2014.

[13] I. Goodfellow, Y. Bengio, and A. Courville. Deep Learning. MIT Press, 2016.

[14] I. Goodfellow, J. Pouget-Abadie, M. Mirza, B. Xu, D. Warde-Farley, S. Ozair, A. Courville, and Y. Bengio. Generative adversarial nets. In Z. Ghahramani, M. Welling, C. Cortes, N. D. Lawrence, and K. Q. Weinberger, editors, Advances in Neural Information Processing Systems 27, pages 2672–2680. Curran Associates, Inc., 2014.

[15] T. Haarnoja, H. Tang, P. Abbeel, and S. Levine. Reinforcement learning with deep energy-based policies. In Proceedings of the 34th International Conference on Machine Learning, pages 1352– 1361, August 2017.

[16] D. O. Hebb. The organization of behavior: A neuropsychological approach. Wiley, 1949.

[17] N. Heess, D. Silver, and Y. W. Teh. Actor-critic reinforcement learning with energy-based policies. In M. P. Deisenroth, C. Szepesv´ari, and J. Peters, editors, Proceedings of the Tenth European Workshop on Reinforcement Learning, volume 24 of Proceedings of Machine Learning Research, pages 45–58, Edinburgh, Scotland, 2013. PMLR.

[18] G. E. Hinton. Training products of experts by minimizing contrastive divergence. Neural Computation, 14(8):1771–1800, August 2002.

[19] G. E. Hinton. Boltzmann machine. Scholarpedia, 2(5):1668, 2007.

[20] G. E. Hinton. Boltzmann machines. In C. Sammut and G. I. Webb, editors, Encyclopedia of Machine Learning. Springer, 2010.

[21] G. E. Hinton and R. Salakhutdinov. Reducing the dimensionality of data with neural networks. Science, 313:504–507, 2006.

[22] A. Hyv¨arinen. Estimation of non-normalized statistical models by score matching. Journal of Machine Learning Research, 6:695–709, 2005.

[23] A. Hyv¨arinen. Some extensions of score matching. Computational Statistics & Data Analysis, 51(5):2499–2512, 2007.

[24] A. Hyv¨arinen. Optimal approximation of signal priors. Neural Computation, 20:3087–3110, 2008.

[25] E. Ising. Beitrag zur theorie des ferromagnetismus.  Zeitschrift f¨ur Physik, 31(1):253?258, 1925.

[26] T. Kim and Y. Bengio. Deep directed generative models with energy-based probability estimation. CoRR, abs/1606.03439, 2016.

[27] R. Kindermann and J. L. Snell. Markov random field and their applications. Americal Mathematical Society, 1980.

[28] D. P. Kingma and J. Ba. Adam: A method for stochastic optimization. In Proceedings of the International Conference on Learning Representations (ICLR), arXiv:1412.6980, 2015.

[29] D. P. Kingma and Y. LeCun. Regularized estimation of image statistics by score matching. In J. D. Lafferty, C. K. I. Williams, J. Shawe-Taylor, R. S. Zemel, and A. Culotta, editors, Advances in Neural Information Processing Systems 23, pages 1126–1134. Curran Associates, Inc., 2010.

[30] U. K¨oster, J. T. Lindgren, and A. Hyv¨arinen. Estimating Markov random field potentials for natural images. In T. Adali T, C. Jutten C, J. M. T. Romano, and A. K. Barros, editors, Independent Component Analysis and Signal Separation. ICA 2009. Lecture Notes in Computer Science, vol. 5441. Springer, 2009.

[31] A. Krizhevsky. Learning multiple layers of features from tiny images. Master’s thesis, Computer Science Department, University of Toronto, Toronto, Canada, 2009.

[32] A. Kulesza and B. Taskar. Determinantal Point Processes for Machine Learning. Now Publishers Inc., Hanover, MA, USA, 2012.

[33] H. Larochelle and Y. Bengio. Classification using discriminative restricted Boltzmann machines. In Proceedings of the 25th international conference on Machine learning (ICML 2008), pages 536–543, 2008.

[34] Y. LeCun, S. Chopra, R. Hadsell, M. A. Ranzato, and F. J. Huang. A Tutorial on Energy-Based Learning. MIT Press, 2006.

[35] Y. LeCun and F. J. Huang. Loss functions for discriminative training of energy-based models. In Proceedings of the 10-th International Workshop on Artificial Intelligence and Statistics, pages 206–213, 2005.

[36] D. A. Levin, Y. Peres, and E. L. Wilmer. Markov Chains and Mixing Times. American Mathematical Society, first edition, 2008.

[37] S. L¨owel and W. Singer. Selection of intrinsic horizontal connections in the visual cortex by correlated neuronal activity. Science, 255:209–212, 1992.

[38] S. Lyu. Interpretation and generalization of score matching. In Proceedings of the 25th Conference in Uncertainty in Artificial Intelligence (UAI09), pages 359–366, June 2009.

[39] O. Macchi. The coincidence approach to stochastic point processes. Advances in Applied Probability, 7(1):83–122, 1975.

[40] Z. Mariet and S. Sra. Kronecker determinantal point processes. In D. D. Lee, M. Sugiyama, U. V. Luxburg, I. Guyon, and R. Garnett, editors, Advances in Neural Information Processing Systems 29, pages 2694–2702. Curran Associates, Inc., 2016.

[41] T. Marks and J. Movellan. Diffusion networks, products of experts, and factor analysis. In Proceedings of the Third International Conference on Independent Component Analysis and Blind Source Separation, 2001.

[42] M. Osadchy, Y. LeCun, and M. Miller. Synergistic face detection and pose estimation with energy-based models. Journal of Machine Learning Research, 8(1197–1215), 2007.

[43] T. Osogami. Boltzmann machines for time-series. Technical Report RT0980, IBM Research - Tokyo, 2017.

[44] T. Osogami, H. Kajino, and T. Sekiyama. Bidirectional learning for time-series models with hidden units. In Proceedings of the 34th International Conference on Machine Learning (ICML 2017), pages 2711–2720, August 2017.

[45] T. Osogami and M. Otsuka. Restricted Boltzmann machines modeling human choice. In Advances in Neural Information Processing Systems 27, pages 73–81. 2014.

[46] T. Osogami, R. Raymond, A. Goel, T. Shirai, and T. Maehara. Dynamic determinantal point processes. 2017.

[47] M. Otsuka and T. Osogami. A deep choice model. In Proceedings of the 30th AAAI Conference on Artificial Intelligence (AAAI-16), pages 850–856, January 2016.

[48] M. Otsuka, J. Yoshimoto, and K. Doya. Free-energy-based reinforcement learning in a partially observable environment. In Proceedings of the European Symposium on Artificial Neural Networks — Computational Intelligence and Machine Learning, pages 28–30, April 2010.

[49] C. Peterson and J. Anderson. A mean field theory learning algorithm for neural networks. Complex Systems, 1(5):995–1019, 1987.

[50] N. Qian. On the momentum term in gradient descent learning algorithms. Neural Networks: The Official Journal of the International Neural Network Society, 12(1):145–151, 1999.

[51] R. R. Salakhutdinov and G. E. Hinton. Deep Boltzmann machines. In Proceedings of the International Conference on Artificial Intelligence and Statistics (AISTATS-09), pages 448–455, April 2009.

[52] B. Sallans and G. E. Hinton. Using free energies to represent Q-values in a multiagent reinforce- ment learning task. In T. K. Leen, T. G. Dietterich, and V. Tresp, editors, Advances in Neural Information Processing Systems 13, pages 1075–1081. MIT Press, 2001.

[53] J. Schmidhuber. Learning factorial codes by predictability minimization. Neural Computation, 4(6):863–879, 1992.

[54] J. Schmidhuber. Deep learning in neural networks: An overview. Neural Networks, 61:85–117, 2015.

[55] T. J. Sejnowski. Higher-order Boltzmann machines. American Institute of Physics, Conference Proceedings, 151(1):398–403, 1986.

[56] Tomoyuki Shirai and Yoichiro Takahashi. Random point fields associated with certain Fredholm determinants. I. Fermion, Poisson and boson point processes. Journal of Functional Analysis, 205(2):414–463, 2003.

[57] Tomoyuki Shirai and Yoichiro Takahashi. Random point fields associated with certain Fredholm determinants. II. Fermion shifts and their ergodic and Gibbs properties. The Annals of Probability, 31(3):1533–1564, 2003.

[58] I. Sutskever and G. E. Hinton. Learning multilevel distributed representations for high-dimensional sequences. In Proceedings of the Eleventh International Conference on Artificial Intelligence and Statistics (AISTATS-07), volume 2, pages 548–555. Journal of Machine Learning Research - Proceedings Track, 2007.

[59] I. Sutskever, G. E. Hinton, and G. W. Taylor. The recurrent temporal restricted Boltzmann machine. In Advances in Neural Information Processing Systems 21, pages 1601–1608. December 2008.

[60] K. Swersky, M. Ranzato, D. Buchman, B. M. Marlin, and N. Freitas. On autoencoders and score matching for energy based models. In Proceedings of the 28th International Conference on Machine Learning (ICML 2011), pages 1201–1208, June 2011.

[61] T. Tieleman and G. E. Hinton. Lecture 6.5—Rmsprop: Divide the gradient by a running average of its recent magnitude. COURSERA: Neural Networks for Machine Learning, 2012.

[62] P. Vincent. A connection between score matching and denoising autoencoders. Neural Computation, 23(7):1661–1674, 2011.

[63] M. Welling and G. E. Hinton. A new learning algorithm for mean field Boltzmann machines. In Proceedings of the 12th International Conference on Artificial Neural Networks, pages 351–357, August 2002.

[64] M. Welling, M. Rosen-Zvi, and G. E. Hinton. Exponential family harmoniums with an application to information retrieval. In Advances in Neural Information Processing Systems 17, pages 1481– 1488. 2004.

[65] M. Welling, R. S. Zemel, and G. E. Hinton. Self supervised boosting. In S. Becker, S. Thrun, and K. Obermayer, editors, Advances in Neural Information Processing Systems 15, pages 681–688. MIT Press, 2003.

[66] S. Zhai, Y. Cheng, W. Lu, and Z. Zhang. Deep structured energy based models for anomaly detection. In Proceedings of the 33rd International Conference on Machine Learning (ICML 2016), pages 1100–1109, June 2016.

[67] J. J. Zhao, M. Mathieu, and Y. LeCun. Energy-based generative adversarial network. CoRR, abs/1609.03126, 2016.


Designed for Accessibility and to further Open Science