In variational Bayesian methods, the evidence lower bound (often abbreviated ELBO, also sometimes called the variational lower bound[1] or negative variational free energy) is a useful lower bound on the log-likelihood of some observed data.

The ELBO is useful because it provides a guarantee on the worst-case for the log-likelihood of some distribution (e.g. ) which models a set of data. The actual log-likelihood may be higher (indicating an even better fit to the distribution) because the ELBO includes a Kullback-Leibler divergence (KL divergence) term which decreases the ELBO due to an internal part of the model being inaccurate despite good fit of the model overall. Thus improving the ELBO score indicates either improving the likelihood of the model or the fit of a component internal to the model, or both, and the ELBO score makes a good loss function, e.g., for training a deep neural network to improve both the model overall and the internal component. (The internal component is , defined in detail later in this article.)

Definition edit

Let   and   be random variables, jointly distributed with distribution  . For example,   is the marginal distribution of  , and   is the conditional distribution of   given  . Then, for a sample  , and any distribution  , the ELBO is defined as

 
The ELBO can equivalently be written as[2]

 

In the first line,   is the entropy of  , which relates the ELBO to the Helmholtz free energy.[3] In the second line,   is called the evidence for  , and   is the Kullback-Leibler divergence between   and  . Since the Kullback-Leibler divergence is non-negative,   forms a lower bound on the evidence (ELBO inequality)

 

Motivation edit

Variational Bayesian inference edit

Suppose we have an observable random variable  , and we want to find its true distribution  . This would allow us to generate data by sampling, and estimate probabilities of future events. In general, it is impossible to find   exactly, forcing us to search for a good approximation.

That is, we define a sufficiently large parametric family   of distributions, then solve for   for some loss function  . One possible way to solve this is by considering small variation from   to  , and solve for  . This is a problem in the calculus of variations, thus it is called the variational method.

Since there are not many explicitly parametrized distribution families (all the classical distribution families, such as the normal distribution, the Gumbel distribution, etc, are far too simplistic to model the true distribution), we consider implicitly parametrized probability distributions:

  • First, define a simple distribution   over a latent random variable  . Usually a normal distribution or a uniform distribution suffices.
  • Next, define a family of complicated functions   (such as a deep neural network) parametrized by  .
  • Finally, define a way to convert any   into a simple distribution over the observable random variable  . For example, let   have two outputs, then we can define the corresponding distribution over   to be the normal distribution  .

This defines a family of joint distributions   over  . It is very easy to sample  : simply sample  , then compute  , and finally sample   using  .

In other words, we have a generative model for both the observable and the latent. Now, we consider a distribution   good, if it is a close approximation of  :

 
since the distribution on the right side is over   only, the distribution on the left side must marginalize the latent variable   away.
In general, it's impossible to perform the integral  , forcing us to perform another approximation.

Since   (Bayes' Rule), it suffices to find a good approximation of  . So define another distribution family   and use it to approximate  . This is a discriminative model for the latent.

The entire situation is summarized in the following table:

 : observable    : latent
  approximable  , easy
 , easy
  approximable  , easy

In Bayesian language,   is the observed evidence, and   is the latent/unobserved. The distribution   over   is the prior distribution over  ,   is the likelihood function, and   is the posterior distribution over  .

Given an observation  , we can infer what   likely gave rise to   by computing  . The usual Bayesian method is to estimate the integral  , then compute by Bayes' rule  . This is expensive to perform in general, but if we can simply find a good approximation   for most  , then we can infer   from   cheaply. Thus, the search for a good   is also called amortized inference.

All in all, we have found a problem of variational Bayesian inference.

Deriving the ELBO edit

A basic result in variational inference is that minimizing the Kullback–Leibler divergence (KL-divergence) is equivalent to maximizing the log-likelihood:

 
where   is the entropy of the true distribution. So if we can maximize  , we can minimize  , and consequently find an accurate approximation  .

To maximize  , we simply sample many  , i.e. use Importance sampling

 
where   is the number of samples drawn from the true distribution. This approximation can be seen as overfitting.[note 1]

In order to maximize  , it's necessary to find  :

 
This usually has no closed form and must be estimated. The usual way to estimate integrals is Monte Carlo integration with importance sampling:
 
where   is a sampling distribution over   that we use to perform the Monte Carlo integration.

So we see that if we sample  , then   is an unbiased estimator of  . Unfortunately, this does not give us an unbiased estimator of  , because   is nonlinear. Indeed, we have by Jensen's inequality,

 
In fact, all the obvious estimators of   are biased downwards, because no matter how many samples of   we take, we have by Jensen's inequality:
 
Subtracting the right side, we see that the problem comes down to a biased estimator of zero:
 
At this point, we could branch off towards the development of an importance-weighted autoencoder[note 2], but we will instead continue with the simplest case with  :
 
The tightness of the inequality has a closed form:
 
We have thus obtained the ELBO function:
 

Maximizing the ELBO edit

For fixed  , the optimization   simultaneously attempts to maximize   and minimize  . If the parametrization for   and   are flexible enough, we would obtain some  , such that we have simultaneously

 
Since
 
we have
 
and so
 
In other words, maximizing the ELBO would simultaneously allow us to obtain an accurate generative model   and an accurate discriminative model  .[5]

Main forms edit

The ELBO has many possible expressions, each with some different emphasis.

 

This form shows that if we sample  , then   is an unbiased estimator of the ELBO.

 

This form shows that the ELBO is a lower bound on the evidence  , and that maximizing the ELBO with respect to   is equivalent to minimizing the KL-divergence from   to  .

 

This form shows that maximizing the ELBO simultaneously attempts to keep   close to   and concentrate   on those   that maximizes  . That is, the approximate posterior   balances between staying close to the prior   and moving towards the maximum likelihood  .

 

This form shows that maximizing the ELBO simultaneously attempts to keep the entropy of   high, and concentrate   on those   that maximizes  . That is, the approximate posterior   balances between being a uniform distribution and moving towards the maximum a posteriori  .

Data-processing inequality edit

Suppose we take   independent samples from  , and collect them in the dataset  , then we have empirical distribution  .


Fitting   to   can be done, as usual, by maximizing the loglikelihood  :

 
Now, by the ELBO inequality, we can bound  , and thus
 
The right-hand-side simplifies to a KL-divergence, and so we get:
 
This result can be interpreted as a special case of the data processing inequality.

In this interpretation, maximizing   is minimizing  , which upper-bounds the real quantity of interest   via the data-processing inequality. That is, we append a latent space to the observable space, paying the price of a weaker inequality for the sake of more computationally efficient minimization of the KL-divergence.[6]

References edit

  1. ^ Kingma, Diederik P.; Welling, Max (2014-05-01). "Auto-Encoding Variational Bayes". arXiv:1312.6114 [stat.ML].
  2. ^ Goodfellow, Ian; Bengio, Yoshua; Courville, Aaron (2016). "Chapter 19". Deep learning. Adaptive computation and machine learning. Cambridge, Mass: The MIT press. ISBN 978-0-262-03561-3.
  3. ^ Hinton, Geoffrey E; Zemel, Richard (1993). "Autoencoders, Minimum Description Length and Helmholtz Free Energy". Advances in Neural Information Processing Systems. 6. Morgan-Kaufmann.
  4. ^ Burda, Yuri; Grosse, Roger; Salakhutdinov, Ruslan (2015-09-01). "Importance Weighted Autoencoders". arXiv:1509.00519 [stat.ML].
  5. ^ Neal, Radford M.; Hinton, Geoffrey E. (1998), "A View of the Em Algorithm that Justifies Incremental, Sparse, and other Variants", Learning in Graphical Models, Dordrecht: Springer Netherlands, pp. 355–368, doi:10.1007/978-94-011-5014-9_12, ISBN 978-94-010-6104-9, S2CID 17947141
  6. ^ Kingma, Diederik P.; Welling, Max (2019-11-27). "An Introduction to Variational Autoencoders". Foundations and Trends in Machine Learning. 12 (4). Section 2.7. arXiv:1906.02691. doi:10.1561/2200000056. ISSN 1935-8237. S2CID 174802445.

Notes edit

  1. ^ In fact, by Jensen's inequality,   The estimator is biased upwards. This can be seen as overfitting: for some finite set of sampled data   , there is usually some   that fits them better than the entire   distribution.
  2. ^ By the delta method, we have
     
    If we continue with this, we would obtain the importance-weighted autoencoder.[4]