Bayesian Deep Learning

Basics

Probability Theory and Bayesian Approach

Why go Bayesian?

  • There's always some uncertainty
    • Complex partially observed world
    • Few observations
  • Probability Theory is a great tool to reason about uncertainty
    • Bayesians quantify subjective uncertainty
    • Frequentists quantify inherent randomness in the long run
  • People seem to interpret probability as beliefs and hence are Bayesians

The Bayesian Workflow

We want to make predictions about some \( x \)

  • We formulate our prior beliefs about how the \( x \) might be generated
  • We collect some data of already generated \( x \): $$ \mathcal{D}_\text{train} = (x_1, ..., x_N) $$
  • We update our beliefs regarding what kind of data exist by incorporating collected data
  • We now can make predictions about unseen data
    • And collect some more data to improve our beliefs

Probability Theory 101

  • We'll assume random variables have and are described by their densities
    • \(X\) – random variable
    • \(p(X=x)\) (\(p(x)\) for short) – its probability density function
    • \(\text{Pr}[X \in A] = \int_{A} p(X=x) dx\) – distribution function
  • In general several random variables \(X_1, ..., X_N\) have joint density $$ p(X_1=x_1, ..., X_N=x_n) $$
    • It describes joint probability $$\text{Pr}(X_1 \in A_1, ..., X_N \in A_N) = \int_{A_1} ... \int_{A_N} p(x_1, ..., x_N) dx_N ... dx_1 $$
    • If (and only if) random variables are independent, the joint density is just a product of individual densities
    • Vector random variables are just a bunch of scalar random variables
  • For 2 and more random variables you should be considering their joint distribution

Probability Theory 101

  • \(\mathbb{E}_{p(x)} X = \int x p(x) dx\) – expected value
    • \( \mathbb{E} [\alpha X + \beta Y] = \alpha \mathbb{E} X + \beta \mathbb{E} Y \)
    • Law of the Unconscious Statistician $$ \mathbb{E} f(x) = \int f(x) p(x) dx $$
  • \( \mathbb{V} X = \mathbb{E} [X^2] - (\mathbb{E} X)^2 = \mathbb{E}(X - \mathbb{E} X)^2 \) variance
    • ​If \(X\) and \(Y\) are independent, \( \mathbb{V}[\alpha X + \beta Y] = \alpha^2 \mathbb{V} X + \beta^2 \mathbb{V} Y \)
  • \( \text{Cov}(X, Y) = \mathbb{E} [X Y] - \mathbb{E} X \mathbb{E} Y \) – covariance
    • \( \mathbb{V} [\alpha X + \beta Y] = \alpha^2 \mathbb{V}[X] + \beta^2 \mathbb{V}[Y] + 2 \alpha \beta \text{Cov}(X, Y) \)
  • \(q_\alpha\) is called an \(\alpha\)-quantile of \(X\) if \(\text{Pr}[X < q_\alpha] = \alpha\) (equivalently \(\text{Pr}[X \ge q_\alpha] = 1 - \alpha\))
    • In particular, \(q_{1/2}\) is called the median

ExampleS

  • \(X\) is said to be Uniformly distributed over \((a, b)\) (denoted \(X \sim U(a, b)\) if its probability density function is $$ p(x) = \begin{cases} \tfrac{1}{b-a}, & a < x < b \\ 0, &\text{otherwise} \end{cases} \quad\quad \mathbb{E} U = \frac{a+b}{2} \quad\quad \mathbb{V} U = \frac{(b-a)^2}{12} $$
  • \(X\) is called a Multivariate Gaussian (Normal) random vector with mean \(\mu \in \mathbb{R}^n\) and positive-definite covariance matrix \(\Sigma \in \mathbb{R}^{n \times n}\) (denoted \(x \sim \mathcal{N}(\mu, \Sigma)\)) if its joint probability density function is $$ \quad\quad\quad\quad\quad\quad\quad\quad\quad \quad \quad p(x) = \frac{1}{\sqrt{\text{det}(2 \pi \Sigma)}} \exp \left( -\tfrac{1}{2} (x-\mu)^T \Sigma^{-1} (x-\mu) \right) $$

$$ \mathbb{E} X = \mu $$

$$ \text{Cov}(X_i, X_j) = \Sigma_{ij} $$

More ExampleS

  • \(X\) is said to be Categorically distributed with probabilities   (non-negative and sum to 1) \(\pi_1, ..., \pi_K\) (denoted \(X \sim \text{Cat}(\pi_1, ..., \pi_K)\)) if its probability density mass function is

 

 

 

 

 

  • \(X\) is called a Bernoulli random variable with probability (of success) \(p \in [0, 1]\) (denoted \(X \sim \text{Bern}(\pi)\)) if its probability mass function is $$ p(X = 1) = \pi \Leftrightarrow p(x) = \pi^{x} (1-\pi)^{1-x} $$ (yes, this is a special case of the categorical distribution)

$$ p(X = k) = \pi_k \Leftrightarrow p(x) = \prod_{k=1}^K \pi_k^{[x = k]} $$

Probability Theory 101

  • Joint density on \(x\) and \(y\) defines the marginal density on each of them: $$ p(x) = \int p(x, y) dy \quad\quad p(y) = \int p(x,y) dx $$
  • Knowing value of \(y\) can reduce uncertainty about \(x\), expressed via the conditional density $$ p(X=x|Y=y) = \frac{p(x, y)}{p(x)} = \frac{p(X=x, Y=y)}{\int p(X=x, Y=z) dz} $$
  • Thus $$ p(x, y) = p(y|x) p(x) = p(x|y) p(y) $$
  • In general the chain rule is $$ p(x_1, \dots, x_N) = p(x_1) p(x_2 | x_1) p(x_3 | x_1, x_2) ... p(x_N | x_1, ..., x_{N-1}) $$

Example

  • Suppose we're having two jointly Gaussian random variables \(X\) and \(Y\): $$(X, Y) \sim \mathcal{N}\left(\left[\begin{array}{c}\mu_x \\ \mu_y \end{array} \right], \left[\begin{array}{cc}\sigma^2_x & \rho_{xy} \\ \rho_{xy} & \sigma^2_y\end{array}\right]\right)$$
  • Then one can show that marginal and conditionals are also Gaussian $$ p(x) = \mathcal{N}(x \mid \mu_x, \sigma^2_x) $$  $$ p(y) = \mathcal{N}(y \mid \mu_y, \sigma^2_y) $$  $$p(x|y) = \mathcal{N}\left(x \mid \mu_x + \tfrac{\rho}{\sigma_x^2} (y - \mu_y), \sigma^2_x - \tfrac{\rho_{xy}^2}{\sigma_y^2}\right)$$

Probability Theory 101

  • This leads to the Bayes' theorem relating conditional distributions $$ p(y|x)= \frac{p(x, y)}{p(x)} = \frac{p(x|y) p(y)}{p(x)} $$
  • If we're interested in \(y\), then these distributions are called
    • \( p(x|y) \) – likelihood function
    • \( p(y) \) – prior distribution
    • \( p(y|x) \) posterior distribution
    • \( p(x) \) – model evidence or marginal likelihood

Bayesian Machine Learning 101

  • We assume some data-generating model $$p(y, \theta \mid x) = p(y \mid x, \theta) p(\theta) $$
    • \(x\) and \(y\) are observed
    • \(\theta\) is not observed, called latent variable
  • We obtain some observations \( \mathcal{D} = \{(x_n, y_n)\}_{n=1}^N \)
  • We seek to make make predictions regarding \(y\) for previously unseen \(x\) having observed the training set \(\mathcal{D}\). Its uncertainty quantified by the posterior predictive distribution $$p(y \mid x, \mathcal{D}) = \int p(y \mid x, \theta) p(\theta \mid \mathcal{D}) d\theta $$
  • This requires us to know the posterior distribution on model parameters \(p(\theta \mid \mathcal{D})\) which we obtain using the Bayes' rule

Example: Bayesian Linear Regression

  • Suppose the model \(y \sim \mathcal{N}(\theta^T x, \sigma^2)\), with \( \theta \sim \mathcal{N}(\mu_0, \sigma_0^2 I) \)
  • Suppose we observed some data from this model \( \mathcal{D} = \{(x_n, y_n)\}_{n=1}^N \) (generated using the same \( \theta^* \))
  • We don't know the optimal \(\theta\), but the more data we observe the more certain we are about possible values of \(\theta\) $$ p(\theta | \mathcal{D}) = \frac{\prod_{n=1}^N p(y_n| x_n, \theta) p(\theta)}{p(\mathcal{D})} = \mathcal{N}(\theta \mid \mu_N, \Sigma_N) $$
  • Posterior predictive would also be Gaussian $$ p(y|x, \mathcal{D}) = \mathcal{N}(y \mid \mu_N^T x, \sigma_N^2) $$
  • This is called Bayesian Linear Regression

Bayesian Linear Regression

Another Example: Coin Tossing

  • Suppose we observe a sequence of coin flips \((x_1, ..., x_N, ...)\), but don't know whether the coin is fair $$ x \sim \text{Bern}(\pi), \quad \pi \sim U(0, 1) $$
  • First, we infer posterior distribution on a hidden parameter \(\pi\) having observed \(x_{<N} = (x_1, ..., x_{N-1}) \): $$ p(\pi | x_{<N}) \propto p(\pi, x_{<N}) = \pi^{\sum_{n=1}^{N-1} x_n} (1-\pi)^{N - 1 - \sum_{n=1}^{N-1} x_n} [0 < \pi < 1] $$ (this is called Beta-distribution)
  • Posterior predictive then is simplified to $$ p(x_N = x \mid x_{<N}) = \left(\tfrac{1 + \sum_{n=1}^{N-1} x_n}{N+1}\right)^{x} \left( \tfrac{1 + \sum_{n=1}^{N-1} (1-x_n)}{N + 1}\right)^{1-x}$$
  • Similarly, one may take \(\text{Beta}(\alpha, \beta)\) as a prior, and obtain the following posterior predictive $$ p(x_N = x \mid x_{<N}) = \left(\tfrac{\alpha + \sum_{n=1}^{N-1} x_n}{N-1 + \alpha + \beta}\right)^{x} \left( \tfrac{\beta + \sum_{n=1}^{N-1} (1-x_n)}{N -1 + \alpha + \beta}\right)^{1-x}$$

ASyMptopia

  • Maximum Likelihood Estimation (MLE) and Maximum a Posteriori  (MAP) are crude approximations to the full Bayesian inference
  • In most cases in the limit of infinite data the true posterior \(p(\theta \mid \mathcal{D})\) collapses into a point, that is we recover the \(\theta^* \) exactly
  • Point estimates assume we already have so much data that the true posterior is very concentrated and hence the approximation by a single point is justifiable
    • MAP allows prevents us from overfitting when we don't have abundance of data, so we essentially rely on prior information
    • MLE builds upon a fact that in the limit of infinite data the particular choice of the prior doesn't matter much
    • There exist many other point estimates, but they are much less frequent

ASyMptopia Example

  • Recall the coin toss problem
  • MLE and MAP are equivalent due to the Uniform prior $$ p(\pi \mid x_{<N}) \propto p(x_{<N} | \pi) [0 < \pi < 1] \to \max_{\pi} $$
  • Equivalent to the following optimization $$ \sum_{n=1}^{N-1} (x_n \log \pi + (1 - x_n) \log (1 - \pi)) \to \max_\pi $$ $$ \pi^* = \tfrac{1}{N-1} \sum_{n=1}^{N-1} x_n $$
  • Predictive distribution is $$ p(x_N|x_{<N}) = \left( \tfrac{1}{N-1} \sum_{n=1}^{N-1} x_n \right)^{x_N} \left( \tfrac{1}{N-1} \sum_{n=1}^{N-1} (1 - x_n) \right)^{1-x_N}$$
  • Overconfident compared to the full Bayesian but the same in the limit of infinite data

Ok, but why

  • Classical point estimates MLE and MAP only work in the limit of infinite data
    • Technically, it means one we have a ton of data, the approximation error becomes negligible
    • However, its not the amount of data itself that matters, but how much data per parameter we have
  • Hence no need for Bayes in classical models in the Big Data regime
  • Not the case for Deep Learning
    • Millions of parameters, relatively little data
    • Neural Networks can easily overfit

Bayesian Hardship

  • Man, why so much math?
    • Idk, I just like it. Seriously though, its just formal language, not much of the actual math is involved
  • We don't need no Bayes, we already learned a lot without it
    • Oh, no, you didn't
  • Why are the slides in english?
    • ​Jeez, how is that related to this slide?
  • How do we know our model assumptions are right?
    • We don't. All our models are just approximations of reality
  • How do we integrate the denominator in the Bayes' rule?
    • Often we can't, we use approximate posteriors
    • Or avoid posterior density altogether, just sample from it
  • How do I chose my priors?
    • You use the prior to express your preferences on a model
    • There are priors that express absence of any preferences

Example Models

Putting Bayesian into Neural Networks

and Neural Networks in Bayesian

Bayesian Neural Networks

  • We have a problem of classifying some objects \(x\) (images, for example) into one of K classes with the correct class given by \(y\)
  • We assume the data is generated using some (partially known) classifier \(\pi_{\theta^*}\): $$ y \mid x, \pi_{\theta^*} \sim \text{Categorical}(\pi_{\theta^*}(x)) $$ where \(\pi_{\theta^*}(\cdot)\) is a neural network of a known structure and unknown weights \(\theta^*\) believed to come from \(p(\theta)\) 
  • After observing the training set \(\mathcal{D}\) the learning boils down to finding \( p(\theta \mid \mathcal{D}) \propto p(\theta) \prod_{n=1}^N p(y_n \mid x_n, \pi_\theta) \)

INTRACTABLE

Bayesian Generative Modeling

  • We want to model uncertainties in, say, images \(x\) (and maybe sample them), but these are very complicated objects
  • We assume that each image \(x\) has some high-level features \(z\) that can help explain its uncertainty in a non-linear way \( p(x \mid f(z)) \ne p(x) \) where \(f\) is a neural network
  • The features are believed to follow some simple distribution \(p(z)\)
  • Having this model would allow us to
    • Sample unseen images via \(z \sim p(z)\), \(x \sim p(x \mid z) \)
    • Detect out-of-domain data using marginal density \(p(x)\)

INTRACTABLE

Bayesian Control Flow

  • Suppose we have a residual neural network $$ H_l(x) = F_l(x) + x $$
  • Can we drop unnecessary computations for easy inputs?
  • Lets equip the network with a mechanism to decide when to stop processing and prefer networks that stop early

Bayesian Control Flow

  • Let \(z\) indicate the number of layers to use. Then $$H_l(x) = [z \le l] F_l(x) + x$$
  • Thus we have $$p(y|x,z) = \text{Categorical}(y \mid \pi(x, z))$$ where \(\pi(x, z)\) is a residual network with \(z\) that controls when to stop processing the \(x\)
  • We chose the prior on \(z\) s.t. lower values are more preferable
  • How to decide upon number of layers at the test time?
    • Obtain a sample from (or the mode statistic of) the true posterior $$ p(y, z \mid x) \propto p(y|x, z) p(z) $$

INTRACTABLE

Approximation Methods

Overcoming the intractability

Variational Inference

  • We define some joint model \(p(y, \theta | x) = p(y | x, \theta) p(\theta) \)
  • We obtain observations \( \mathcal{D} = \{ (x_1, y_1), ..., (x_N, y_N) \} \)
  • We would like to infer possible values of \(\theta\) given  observed data \(\mathcal{D}\) $$ p(\theta \mid \mathcal{D}) = \frac{p(\mathcal{D} | \theta) p(\theta)}{\int p(\mathcal{D}|\theta) p(\theta) d\theta} $$
    • The denominator is generally intractable 😭
    • We will be approximating true posterior distribution with an approximate one
  • Need a distance between distributions to measure how good the approximation is $$ \text{KL}(q(x) || p(x)) = \mathbb{E}_{q(x)} \log \frac{q(x)}{p(x)} \quad\quad \textbf{Kullback-Leibler divergence}$$
    • Not an actual distance, but \( \text{KL}(q(x) || p(x)) = 0 \) iff \(q(x) = p(x)\) for all \(x\) and is strictly positive otherwise
    • Will be minimizing \( \text{KL}(q(\theta) || p(\theta | \mathcal{D})) \) over \(q\)

Variational Inference

  • We'll take \(q(\theta)\) from some tractable parametric family, for example Gaussian $$ q(\theta | \Lambda) = \mathcal{N}(\theta \mid \mu(\Lambda), \Sigma(\Lambda)) $$
  • Then we reformulate the objective s.t. we don't need the exact true posterior $$ \text{KL}(q(\theta | \Lambda) || p(\theta | \mathcal{D})) = \log p(\mathcal{D}) - \mathbb{E}_{q(\theta | \Lambda)} \log \frac{p(\mathcal{D}, \theta)}{q(\theta | \Lambda)} $$
  • Hence we seek parameters \(\Lambda_*\) maximizing the following objective (the ELBO) $$ \Lambda_* = \text{argmax}_\Lambda \left[ \mathbb{E}_{q(\theta | \Lambda)} \log \frac{p(\mathcal{D}, \theta)}{q(\theta|\Lambda)} = \mathbb{E}_{q(\theta|\Lambda)} \log p(\mathcal{D}|\theta) - \text{KL}(q(\theta|\Lambda)||p(\theta)) \right]$$
  • Once we find these, we get an approximate posterior predictive $$ q(y \mid x, \mathcal{D}) = \mathbb{E}_{q(\theta \mid \Lambda_*)} p(y \mid x, \theta) $$
  • We can't compute this quantity analytically either, but can sample from \(q\) to get Monte Carlo estimates of the approximate posterior predictive distribution: $$ q(y \mid x, \mathcal{D}) \approx \hat{q}(y|x, \mathcal{D}) = \frac{1}{M} \sum_{m=1}^M p(y \mid x, \theta^m), \quad\quad \theta^m \sim q(\theta \mid \Lambda_*) $$

Gradient-Based Variational Inference

  • Recall the objective for variational inference $$ \mathcal{L}(\Lambda_*) = \mathbb{E}_{q(\theta | \Lambda)} \log \frac{p(\mathcal{D}, \theta)}{q(\theta|\Lambda)} \to \max_{\Lambda} $$
  • We'll be using well-known optimization method Stochastic Gradient Descent
  • We need (stochastic) gradient \(\hat{g}\) of \(\mathcal{L}(\Lambda)\) s.t. \(\mathbb{E} \hat{g} = \nabla_\Lambda \mathcal{L}(\Lambda) \)
    • Problem: We can't just take \(\hat{g} = \nabla_\Lambda \log \frac{p(\mathcal{D}, \theta)}{q(\theta | \Lambda)} \) as the samples themselves depend on \(\Lambda\) through \(q(\theta|\Lambda)\)
  • Remember the expectation is just an integral, and apply the log-derivative trick $$ \nabla_\Lambda q(\theta | \Lambda) = q(\theta | \Lambda) \nabla_\Lambda \log q(\theta|\Lambda) $$ $$ \nabla_\Lambda \mathcal{L}(\Lambda) = \int q(\theta|\Lambda) \log \frac{p(\mathcal{D}, \theta)}{q(\theta | \Lambda)} \nabla_\Lambda \log q(\theta | \Lambda) d\theta = \mathbb{E}_{q(\theta|\Lambda)} \log \frac{p(\mathcal{D}, \theta)}{q(\theta|\Lambda)} \nabla \log q(\theta | \Lambda) $$
  • Though general, this gradient estimator has too much variance in practice
    • Estimator is called REINFORCE, random search in disguise

Bayesian Neural Networks

Preferred Neural Networks

Bayesian Neural Networks

  • We assume the data is generated using some (partially known) classifier \(\pi_{\theta}\): $$ p(y \mid x, \theta) = \text{Cat}(y | \pi_\theta(x)) \quad\quad \theta \sim p(\theta) $$
  • True posterior is intractable $$ p(\theta \mid \mathcal{D}) \propto p(\theta) \prod_{n=1}^N p(y_n \mid x_n, \pi_\theta) $$
  • Approximate it using \(q(\theta | \Lambda)\): $$ \Lambda_* = \text{argmax} \; \mathbb{E}_{q(\theta | \Lambda)} \left[\sum_{n=1}^N \log p(y_n | x_n, \theta)  - \text{KL}(q(\theta | \Lambda) || p(\theta))\right] $$ 
  • Essentially, instead of learning a single neural network that would solve the problem, we learn a distribution over networks that solve the problem – a (Bayesian) ensemble
  • \(p(\theta)\) encodes our preferences on which networks we'd like to see

Dropout as a Bayesian Procedure

  • Let \(q(\theta_i | \Lambda)\) be s.t. with some fixed probability \(p\) it's 0 and with probability \(1-p\) it's some learnable value \(\Lambda_i\)
  • Then for some prior \(p(\theta)\) our optimization objective is $$ \mathbb{E}_{q(\theta|\Lambda)} \sum_{n=1}^N \log p(y_n | x_n, \theta) \to \max_{\Lambda} $$ where the KL term is missing due to the model choice
  • No need to take special care about differentiating through samples
  • We recover DropConnect
    • The same can be done with dropout
    • Turns out, these are bayesian approximate inference procedures
  • What if we want to tune dropout rates \(p\)?
  • Gradient-based optimization in discrete models is hard
    • Invoke the Central Limit Theorem and turn the model into a continuous one

Sparse Bayesian Neural Networks

  • Consider a model with continuous noise on weights $$ q(\theta_i | \Lambda) = \mathcal{N}(\theta_i | \mu_i(\Lambda), \alpha_i(\Lambda) \mu^2_i(\Lambda)) $$
  • Neural Networks have lots of parameters, surely there's some redundancy in them
  • Let's take a prior \(p(\theta)\) that would encourage large \(\alpha\)
  • Large \(\alpha_i\) would imply that weight \(\theta_i\) is unbounded noise that corrupts predictions
    • Such weights won't be doing anything useful, hence it should be zeroed out by putting \(\mu_i(\Lambda) = 0\)
    • Thus the weight \(\theta_i\) would effectively turn into a deterministic 0
    • Large \(\alpha\) encourage sparsification
  • How do we backpropagate through samples \(\theta_i\)?

Reparametrization Trick

  • We have a continuous density \(q(\theta_i | \mu_i(\Lambda), \sigma_i^2(\Lambda))\) and would like to compute the gradient of $$ \mathbb{E}_{q(\theta|\Lambda)} \log \frac{p(\mathcal{D}|\theta) p(\theta)}{q(\theta|\Lambda)} $$
  • The gradient consists of two parts
    • The inner part – expected gradients of \(\log \frac{p(\mathcal{D}|\theta) p(\theta)}{q(\theta|\Lambda)} \)
      • This is easy
    • Sampling part – gradients through samples \( \theta \sim q(\theta|\Lambda) \)
  • Reparametrization trick: replace a sample from \(q(\theta_i | \mu_i(\Lambda), \sigma_i^2(\Lambda))\) with a transformation of a sample from a parameter-less distribution: $$ \theta \sim \mathcal{N}(\mu, \sigma^2) \Leftrightarrow \theta = \mu + \sigma \varepsilon, \quad \varepsilon \sim \mathcal{N}(0, 1) $$
  • The objective then becomes $$ \mathbb{E}_{\varepsilon \sim \mathcal{N}(0, 1)} \log \tfrac{p(\mathcal{D}, \mu + \varepsilon \sigma)}{q(\mu + \varepsilon \sigma | \Lambda)} $$

Sparsification in action

Variational Dropout Sparsifies Deep Neural Networks

D. Molchanov, A. Ashukha, D. Vetrov, ICML 2017

Bayesian Neural Networks: Conclusion

  • The objective then becomes $$ \mathbb{E}_{\varepsilon \sim \mathcal{N}(0, 1)} \left[\sum_{n=1}^N \log p(y_n | \theta=\mu(\Lambda) + \varepsilon \sigma(\Lambda)) \right] - \text{KL}(q(\theta|\Lambda) || p(\theta)) $$
  • Training a neural network with special kind of noise upon weights
    • The magnitude of the noise is encouraged to increase
    • Zeroes out unnecessary weights completely
  • Essentially, training a whole ensemble of neural networks
    • Actually using the ensemble is costly: \(k\) times slow for an ensemble of \(k\) models
    • Single network (single-sample ensemble) also work

Bayesian Generative Models

Generating everything out of nothing

Bayesian Generative Models

  • We assume the two-phase data-generating process:
    • First, we decide upon high-level abstract features of the datum \(z \sim p(z)\)
    • Then, we unpack these features using Neural Networks into an actual observable \(x\) using the (learnable) generator \(f_\theta\)
  • This leads to the following model \(p(x, z) = p(x|z) p(z)\) where $$ p(x|z) = p(z) \prod_{d=1}^D p(x_d | f_\theta(z)) $$ $$ p(z) = \mathcal{N}(z | 0, I) $$ and \(f_\theta\) is some neural network
  • We can sample new \(x\) by passing samples \(z\) through the generator once we learn it
  • Would like to maximize log-marginal density of observed variables \(\log p(x)\)
    • Intractable integral \( \log p(x) = \log \int p(x|z) p(z) dz \)
    • Monte Carlo doesn't help much

Variational Inference to the rescue

  • Introduce approximate posterior \(q(z|x)\): $$ q(z|x) = \mathcal{N}(z|\mu_\Lambda(x), \Sigma_\Lambda(x))$$
    • Where \(\mu, \Sigma\) are generated using auxiliary inference network from the observation \(x\)
  • Invoking the ELBO we obtain the following objective $$ \tfrac{1}{N} \sum_{n=1}^N \left[ \mathbb{E}_{q(z_n|x_n)} \log p(x_n | z_n) - \text{KL}(q(z_n|x_n)||p(z_n)) \right] \to \max_\Lambda $$
  • Turns out, the ELBO is also a lower bound on marginal log-likelihood (hence the name), we can maximize it w.r.t. to parameters \(\theta\) of the generator also!
  • Generator network and inference network essentially give us autoencoder
    • Inference network encodes observations into latent code
    • Generator network decodes latent code into observations

VAE

  • Latent-variable Generative Model
    • Generates observation from noise
    • Can infer high-level abstract features of existing objects
      • Useful for representation learning
      • And semi-supervised learning
  • Uses neural network to amortize inference

Auto-Encoding Variational Bayes

D. P Kingma, M. Welling, ICLR 2013

Conclusion

What this all was for

ConCLUSION

  • Bayesian methods are useful when we have low data-to-parameters ratio
    • The Deep Learning case!
  • Bayesian methods can
    • Impose useful priors on Neural Networks helping discover solutions of special form
    • Provide better predictions
    • Provide Neural Networks with uncertainty estimates (uncovered)
  • Neural Networks help us make more efficient Bayesian inference
  • Uses a lot of math
  • Active area of research

Shameless Plug(s)

I sometimes blog about different cutting-edge-like topics:

http://artem.sobolev.name/

Made with Slides.com