Gaussian Mixture Model

A Gaussian Mixture Model is a probability model that describes the probability of data using a mixture of gaussian distributions. Given a fixed parameter $K$ denoting the number of gaussian distrbutions in the mixture the probability of a point $x$ is given by: \begin{equation} P(x; \pi, \mu, \Sigma) = \sum_{i=1}^{K} \pi_i \mathcal{N}(x|\mu_i, \Sigma_i) \end{equation}

It can be seen that a Gaussian Mixture Model (GMM) is parameterised by the following parameters:

  1. $\pi$: $\pi_i$ denotes the prior probability of a given point $x$ belonging to $i^{th}$ gaussian component
  2. $\mu$: $\mu_i$ denotes the mean for the $i^{th}$ gaussian component
  3. $\Sigma$: $\Sigma_i$ denotes the covariance matrix for the $i^{th}$ gaussian component

It is easy to see that $P(x)$ is indeed a valid probability density function. To generate a data point from this distribution we first sample a variable $z \sim \pi$. If the sampled variable $z = k$ then the point $x$ is generated by sampling from $\mathcal{N}(\mu_k, \Sigma_k)$

GMMs can also be interpretted as latent variable models. The observed data is $x$ and the associated latent variable is $z$ which denotes the component of the mixture which has generated $x$. More precisely, the following equations govern the data:

\begin{align} P(z = k) &= \pi_k \\ P(x|z = k) &= \mathcal{N}(x|\mu_k, \Sigma_k) \\ \end{align}

Generating Samples from GMMs

As described earlier, generating a sample from a GMM can be viewed as a two step process.

  1. Sample $z \sim \pi$, to decide the mixture component to which $x$ must belong
  2. Let the sampled value of $z$ in first step be $k$. Sample $x \sim \mathcal{N}(x|\mu_k, \Sigma_k)$.

The file gm_generate.py contains the code to generate samples from a GMM. Here we will use gm_generate.py to generate $100$ data samples from a GMM with $K=3$. All the generated $x_i \in \mathbb{R}^2$. We then plot these points using different colors for different values of $z$.


In [1]:
import numpy as np
import matplotlib.pyplot as plt
import random
from gm_generate import *

# Initialize P(z=k) for all k (pi)
pz = [1.0/3]*3 # Using uniform distribution

# Initialize the means mu_k
mu_1 = np.asarray([0.0, 0.0])
mu_2 = np.asarray([5.0, 3.0])
mu_3 = np.asarray([-4.5, -3.0])
mu = [mu_1, mu_2, mu_3]

# Initialize the covariance sigma_k
sigma_1 = np.asarray([[1.0, 0.0], [0.0, 1.0]])
sigma_2 = np.asarray([[3.0, 0.0], [0.0, 1.0]])
sigma_3 = np.asarray([[0.5, 2.0], [0.5, 3.0]])
sigma = [sigma_1, sigma_2, sigma_3]

# Get the GMM object
gmm = GMM(pz=pz, muks=mu, sigmaks=sigma, k=3, d=2)

# Generate 100 data points
(x, z) = gmm.generate_points(n=100)

# Plot the data
idx_1 = np.where(z == 0) # Indices for first cluster
plt.plot(x[idx_1, 0], x[idx_1, 1], 'ro')
idx_2 = np.where(z == 1) # Indices for second cluster
plt.plot(x[idx_2, 0], x[idx_2, 1], 'go')
idx_3 = np.where(z == 2) # Indices for third cluster
plt.plot(x[idx_3, 0], x[idx_3, 1], 'bo')
plt.show()


Note that gm_generate.py can also be used to sample from a single gaussian distribution. Just set the value of $K$ to $1$. The following code uses gm_generate.py to generate $1000$ data points drawn from a 1-D gaussian distribution.


In [2]:
import numpy as np
import matplotlib.pyplot as plt
import random
from gm_generate import *

# Initlialize P(z=k) for all k, here k can take only 1 value
pz = [1.0]

# Initialize mu and sigma for the distribution
mu = [np.asarray([0.0])]
sigma = [np.asarray([1]).reshape((1, 1))]

# Get the GMM object
gmm = GMM(pz=pz, muks=mu, sigmaks=sigma, k=1, d=1)

# Generate 1000 data points
x, _ = gmm.generate_points(n=1000)

# Show the histogram
plt.hist(x)
plt.show()


Probabilistic Inference

Often we have more than one random variables which interact with each other in complicated ways. At any given time we can observe the value of some of these random variables. We call these observed random variables. There are other random variables whose value we can not observe. These are called unobserved random variables. Inference deals with finding a probability distribution over one or more unobserved variables given the values of all observed variables.

In the context of GMMs we observe the data points $x_i$'s and usually the aim is to infer the value of corresponding latent variable $z_i$'s. Using Bayes' Rule we can calculate:

\begin{equation} P(z|x) = \frac{P(x|z)P(z)}{P(x)} \end{equation}

The denominator term can be calculated using the total probability theorem as $P(x) = \sum_{z} P(x|z)P(z)$. The summation is over all possible values of $z$. Note that in the particular case of GMMs, the data points are drawn iid and $z$ is a discrete random variable that takes $K$ distinct values, hence this summation is easy to calculate. In general, however, calculating $P(x)$ is an intractible problem. Thus exact inference is intractible in general and that is why we need to resort to approximation methods. Variational Inference is one such method.

Variational Inference

The objective is to approximate the posterior $p(z|x)$. Variational Inference turns the problem of inference into an optimization problem. Let $\mathcal{D}$ be a parameteric family of distribution, then we find a distribution $q^*(z) \in \mathcal{D}$ as:

\begin{equation} q^*(z) = arg\min_{q(z) \in \mathcal{D}} KL(q(z) || p(z|x)) \end{equation}

$q^*(z) \in \mathcal{D}$ serves as an approximation to the true posterior $p(z|x)$. In the equation above KL-divergence has been used as a measure for calculating the degree of "closeness" between the distributions $q$ and $p$. This is the most common choice although it is not the only choice.

\begin{align} KL(q(z) || p(z|x)) &= \mathbb{E}_{z \sim q(z)} \big[\log q(z) - \log p(z|x)\big] \\ &= \mathbb{E}_{z \sim q(z)} \big[\log q(z) - \log p(x,z) + \log p(x)\big] \\ &= \mathbb{E}_{z \sim q(z)} \big[\log q(z) - \log p(x,z)\big] + \log p(x) \end{align}

Note that the calculation of $KL(q||p)$ involves the calculation of $p(x)$ which as we noted earlier is intractible. Hence calculating the exact value for $KL(q||p)$ is intractible. However since we wish to minimize $KL(q||p)$ one can resort to minimizing an upper bound on $KL(q||p)$. Note that since $\log p(x)$ will always be negative the following inequality follows from the previous equation:

\begin{equation} KL(q(z) || p(z|x)) \leq \mathbb{E}_{z \sim q(z)} \big[\log q(z) - \log p(x,z)\big] \end{equation}

For most models that are used in practice the term $p(x, z)$ is easy to calculate.

Evidence Lower Bound Objective (ELBO)

The negative of the quantity on the right hand side of the previous equation is known as ELBO.

\begin{equation} ELBO(q) = \mathbb{E}_{z \sim q(z)} \big[\log p(x,z) - \log q(z)\big] \end{equation}

The following statements are easy to prove:

  1. $ELBO(q) \leq \log p(x)$, that is why it is called the evidence lower bound
  2. Maximizing ELBO is equivalent to minimizing an upper bound on $KL(q||p)$
  3. Maximizing ELBO is also equivalent to maximizing $\log p(x)$

Choosing $\mathcal{D}$

In order to completely specify the optimization problem one final thing that we need is a specification for $\mathcal{D}$. The complexity of $\mathcal{D}$ determines the overall complexity of the optimization problem. In general $\mathcal{D}$ is chosen such that:

  1. $\mathcal{D}$ is expressive enough to approximate the posterior $p(z|x)$ but simple enough to make the optimization problem tractable
  2. It must be easy to calculate the expectation term in the objective function

For example one can set $\mathcal{D} = \{\mathcal{N}(\mu, \Sigma)\,\,:\,\, \mu \in \mathbb{R}^d, \Sigma \in \mathbb{R}^{d \times d}\}$. $\mathcal{D}$ is a parametric family of normal distribution and the parameters are $\mu$ and $\Sigma$.

The optimization problem is: \begin{equation} q^*(z) = arg\max_{q(z) \sim \mathcal{D}} ELBO(q) \end{equation}

Toy Example 1 - Gaussian Distribution

We begin our exploration of Variational Inference with a very simple problem. We are given a set of $N$ real numbers $x_1, x_2, ..., x_N$ and the task is to fit a gaussian distribution $\mathcal{N}(\mu, \sigma^2)$ to this data. This can be very easily done by using Maximum Likelihood Estimation to compute point estimates for parameters $\mu$ and $\sigma^2$. Another approach is to pose this problem into the Variational Inference framework to get an approximation for the unknown distribution. We implement both these approaches.

Using Maximum Likelihood Estimation

Assuming that the data points are drawn iid from a gaussian distrubution $\mathcal{N}(\mu, \sigma^2)$, the log probability of data is given by:

\begin{align} \mathcal{L}(x, \mu, \sigma^2) &= \log P(x; \mu, \sigma^2) \\ &= \sum_{i=1}^{N} \log \mathcal{N}(x_i| \mu, \sigma^2) \\ &= \sum_{i=1}^{N} \big(-\frac{1}{2}\log 2\pi - \frac{1}{2}\log \sigma^2 - \frac{(x_i - \mu)^2}{2\sigma^2}\big) \end{align}

Under Maximum Likelihood Estimation we maximize $\mathcal{L}(x, \mu, \sigma^2)$ with respect to the paramters $\mu$ and $\sigma^2$. Taking the gradient of $\mathcal{L}$ with respect to the parameters and equating it to zero we get the following equations:

\begin{align} \frac{\partial \mathcal{L}}{\partial \mu} = 0 &\Rightarrow \mu = \frac{1}{N} \sum_{i=1}^{N} x_i \\ \frac{\partial \mathcal{L}}{\partial \sigma^2} = 0 &\Rightarrow \sigma^2 = \frac{1}{N} \sum_{i=1}^{N} (x_i - \mu)^2 \end{align}

The following code first generates $100$ data points from a gaussian distrubution $\mathcal{N}(0, 1)$ using gm_generate.py and then fits a gaussian distribution to the data using Maximum Likelihood Estimation as described above.


In [3]:
import numpy as np
import matplotlib.pyplot as plt
import random
from gm_generate import *

# Generate the data
pz = [1.0]
mu = [np.asarray([0.0])]
sigma = [np.asarray([1]).reshape((1, 1))]
gmm = GMM(pz=pz, muks=mu, sigmaks=sigma, k=1, d=1)
x, _ = gmm.generate_points(n=100)

# Compute maximum likelihood estimation for the parameters
mu_mle = np.mean(x) # mu_mle = sample mean
sigma_mle = np.var(x) # sigma_mle = sample variance

# Show the histogram, true distribution and estimated distribution
plt.hist(x, normed=True, color='#cccccc')
def true_dist(x):
    return (1.0/np.sqrt(2*np.pi*sigma[0])*np.exp(-0.5*((x-mu[0])/sigma[0])**2))
def estimated_dist(x):
    return (1.0/np.sqrt(2*np.pi*sigma_mle)*np.exp(-0.5*((x-mu_mle)/sigma_mle)**2))
x_axis = np.arange(np.min(x)-0.5, np.max(x)+0.5, 0.01)
plt.plot(x_axis, true_dist(x_axis).reshape((-1,)), 'ro', label='True Distribution')
plt.plot(x_axis, estimated_dist(x_axis).reshape((-1,)), 'co', label='Estimated Distribution')
plt.legend(bbox_to_anchor=(1.05, 1), loc=2, borderaxespad=0.)
plt.show()


Using Variational Inference

We now pose this problem in the Variational Inference framework. $\mu$ and $\sigma^2$ will be treated as random variables and we are interested in finding the posterior distribution $P(\mu, \sigma^2 | X)$, where $X = \{x_1, ..., x_N\}$. As described earlier, we approximate $P(\mu, \sigma^2 | X)$ by a parametric distribution $q(\mu, \sigma^2)$. Further we assume that $q(\mu, \sigma^2)$ can be factorised into $q_1(\mu)q_2(\sigma^2)$. This assumption is known as the mean-field assumption and it will be discussed in detail later on. The first step is to write a parametric form for $q_1(\mu)$ and $q_2(\sigma^2)$. We assign the following parametric forms:

\begin{align} q_1(\mu) &= \mathcal{N}(\mu|m, s^2) \\ q_2(\sigma^2) &= Gamma(\alpha, \beta) \end{align}

The parameters $m$, $s^2$, $\alpha$ and $\beta$ are called Variational Parameters and they will be determined by using the available data $X$. We also need a prior distribution over $\mu$ and $\sigma^2$ in order to calculate the joint distribution $P(x, \mu, \sigma^2)$. The prior reflects our knowledge about the domain of the problem. This is the step where subject experts can step in and help in proper modelling of the problem. We use the following priors:

\begin{align} p(\mu) &= \mathcal{N}(\mu|m_p, s_p^2) \\ p(\sigma^2) &= Gamma(\alpha_p, \beta_p) \end{align}

The parameters $m_p$, $s_p^2$, $\alpha_p$ and $\beta_p$ are the hyperparameters that will be provided by the user. Now we are ready to calculate the ELBO.

\begin{align} ELBO(q) &= \mathbb{E}_{\mu, \sigma^2 \sim q}\big[\log p(X, \mu, \sigma^2) - \log q(\mu, \sigma^2)\big] \\ &= \mathbb{E}_{\mu, \sigma^2 \sim q}\big[\log p(X | \mu, \sigma^2)p(\mu)p(\sigma^2) - \log q_1(\mu)q_2(\sigma^2)\big] \\ &= \mathbb{E}_{\mu, \sigma^2 \sim q}\big[\log p(X | \mu, \sigma^2) + \log p(\mu) + \log p(\sigma^2) - \log q_1(\mu) - \log q_2(\sigma^2)\big] \\ &= \mathbb{E}_{\mu, \sigma^2 \sim q}\big[\sum_{i=1}^{N}\log p(x_i | \mu, \sigma^2) + \log p(\mu) + \log p(\sigma^2) - \log q_1(\mu) - \log q_2(\sigma^2)\big] \\ &= K + \mathbb{E}_{\mu, \sigma^2 \sim q}\big[\sum_{i=1}^{N}\big(-\frac{\log \sigma^2}{2} - \frac{(x_i-\mu)^2}{2\sigma^2}\big) -\frac{(\mu-m_p)^2}{2s_p^2} + (\alpha_p - 1)\log \sigma^2 - \frac{\sigma^2}{\beta_p}\big] - \mathbb{E}_{\mu \sim q_1}\big[\log q_1(\mu)\big] - \mathbb{E}_{\sigma^2 \sim q_2}\big[\log q_2(\sigma^2)\big] \\ &= K + \mathbb{E}_{\mu, \sigma^2 \sim q}\big[\sum_{i=1}^{N}\big(-\frac{\log \sigma^2}{2} - \frac{(x_i-\mu)^2}{2\sigma^2}\big) -\frac{(\mu-m_p)^2}{2s_p^2} + (\alpha_p - 1)\log \sigma^2 - \frac{\sigma^2}{\beta_p}\big] + \mathcal{H}(q_1) + \mathcal{H}(q_2) \\ &= K + \mathbb{E}_{\mu, \sigma^2 \sim q}\big[-\frac{N \log \sigma^2}{2} - \sum_{i=1}^{N}\frac{(x_i-\mu)^2}{2\sigma^2} -\frac{N(\mu-m_p)^2}{2s_p^2} + (\alpha_p - 1)\log \sigma^2 - \frac{\sigma^2}{\beta_p}\big] + \mathcal{H}(q_1) + \mathcal{H}(q_2) \\ &= K + \frac{\Gamma(\alpha-1)}{\beta\Gamma(\alpha)}\big(m\sum_{i=1}^{N} x_i - \frac{N}{2}(m^2 + s^2) - \frac{1}{2}\sum_{i=1}^{N} x_i^2\big) + (\alpha_p - \frac{N}{2})(\psi(\alpha) + \log \beta) - \frac{\alpha\beta}{\beta_p} + \alpha + \log \Gamma(\alpha) -\alpha\psi(\alpha) + \frac{\log 2s^2\pi e}{2} \end{align}

Here we have used the following results:

  1. $\mathcal{H}(q_1) = \frac{1}{2} \log 2\sigma^2\pi e$
  2. $\mathcal{H}(q_2) = \alpha + \log \beta + \log \Gamma(\alpha) + (1-\alpha)\psi(\alpha)$
  3. $q(\mu, \sigma^2) = q_1(\mu)q_2(\sigma^2)$

The function $\psi(\alpha)$ is called Digamma function. $\psi(\alpha) = \frac{\Gamma'(\alpha)}{\Gamma(\alpha)}$. $K$ contains all the terms which are independent of variational parameters (and hence can be ignored while optimizing ELBO). The given expression for ELBO is maximized with respect to the variational parameters $\alpha$, $\beta$, $m$ and $s^2$. Unfortunately a closed form solution for optimal variational parameters does not exist in this case, although we can used gradient based techniques for finding them.

We now turn to implementation where we use TensorFlow to maximize the ELBO. The file toy_example_1.py contains the function train_step() that takes as input the data $X$ and performs one iteration of update to the variational parameters. As before $100$ data points are generated using gm_generate.py and the data is used to find the optimal variational parameters using toy_example_1.py.


In [15]:
import numpy as np
import matplotlib.pyplot as plt
import random
from gm_generate import *
from toy_example_1 import *

# Generate the data
pz = [1]
mu = [np.asarray([0.0])]
sigma = [np.asarray([1]).reshape((1, 1))]
gmm = GMM(pz=pz, muks=mu, sigmaks=sigma, k=1, d=1)
x, _ = gmm.generate_points(n=100)
x = x.reshape((-1,))

# Hyperparameter
mu_p=0.0
sigma_p=1.0
alpha_p=3.5
beta_p=0.5
learning_rate=1e-2

# Compute variational inference estimate for the parameters
costs = []
toy_example = ToyExample1(mu_p, sigma_p, alpha_p, beta_p, learning_rate)
for i in range(1000):
    cost, mu_ex, sigma_ex, alpha, beta = toy_example.train_step(x)
    costs.append(cost)

# Compute mu_expected, sigma_expected
mu_expected = mu_ex # Expected value of mu using q_1 (normal)
sigma_expected = beta*alpha # Expected value of sigma using q_2 (gamma)

# Print results
print 'Optimal m:', mu_ex
print 'Optimal s^2:', sigma_ex
print 'Optimal alpha:', alpha
print 'Optimal beta:', beta
print 'Expected Value for mu:', mu_expected
print 'Optimal Value for sigma^2:', sigma_expected
    
# Plot cost vs iterations
plt.plot(costs)
plt.title('Iteration vs ELBO')
plt.xlabel('Iterations')
plt.ylabel('ELBO')
plt.show()

# Show the histogram, true distribution and estimated distribution
plt.hist(x, normed=True, color='#cccccc')
def true_dist(x):
   return (1.0/np.sqrt(2*np.pi*sigma[0])*np.exp(-0.5*((x-mu[0])/sigma[0])**2))
def estimated_dist(x):
    return (1.0/np.sqrt(2*np.pi*sigma_expected)*np.exp(-0.5*((x-mu_expected)/sigma_expected)**2))
x_axis = np.arange(np.min(x)-0.5, np.max(x)+0.5, 0.01)
plt.plot(x_axis, true_dist(x_axis).reshape((-1,)), 'ro', label='True Distribution')
plt.plot(x_axis, estimated_dist(x_axis).reshape((-1,)), 'co', label='Estimated Distribution')
plt.legend(bbox_to_anchor=(1.05, 1), loc=2, borderaxespad=0.)
plt.show()


Optimal m: 0.0223558
Optimal s^2: 0.0101681906669
Optimal alpha: 14.7924620757
Optimal beta: 0.0737221009091
Expected Value for mu: 0.0223558
Optimal Value for sigma^2: 1.09053138184

Note that unlike Maximum Likelihood Estimation, Variational Inference does not return point estimates for $\mu$ and $\sigma^2$. In this case we get the distributions describing $\mu$ and $\sigma^2$ which are $\mathcal{N}(m, s^2)$ and $Gamma(\alpha, \beta)$ respectively. In order to get a single value for $\mu$ and $\sigma^2$ we find the expected values of these variables under their respective distributions. The plot shows that these expected values are indeed very close to the ground truth values.

Another important issue involved here is the choice of hyperparameters. One has to select the prior carefully, although in the prescence of a lot of data - as it is commonly said - data overwhelms prior. Hence if we have a lot of data, the choice of hyperparameters do not significantly affect the final results. Next we demonstrate this. To avoid clutter we will call function run_toy_example_1() from toy_example_1.py which contains exactly the same code as the previous demo.


In [25]:
# Variational Inference with a bad prior using only 10 examples
from toy_example_1 import *

data_mean = 10.0
data_variance = 1.0
num_examples = 10
num_iters = 1000

# Hyperparameters
mu_p = -1.0 # Prior is peaked at mu = -1 but true mean is mu = 10
sigma_p = 1.0
alpha_p = 10 # The optimal value will be close to 4
beta_p = 10 # The optimal value will be close to 0
learning_rate=1e-1

# Run the demo
run_toy_example_1(num_examples, data_mean, data_variance, mu_p, sigma_p, alpha_p, beta_p, learning_rate, num_iters)


Optimal m: 9.83141
Optimal s^2: 4.75094301673
Optimal alpha: 6.93314044937
Optimal beta: 8.00713458598
Expected Value for mu: 9.83141
Optimal Value for sigma^2: 55.5145886816

It can be seen that when only 10 examples were used, although the mean was estimated properly, the biased prior for variance lead to incorrect estimation for variance. However, if we use 100 examples, even with the biased prior we get approximately the correct results.


In [26]:
# Variational Inference with a bad prior using only 10 examples
from toy_example_1 import *

data_mean = 10.0
data_variance = 1.0
num_examples = 100
num_iters = 1000

# Hyperparameters
mu_p = -1.0 # Prior is peaked at mu = -1 but true mean is mu = 10
sigma_p = 1.0
alpha_p = 10 # The optimal value will be close to 4
beta_p = 10 # The optimal value will be close to 0
learning_rate=1e-1

# Run the demo
run_toy_example_1(num_examples, data_mean, data_variance, mu_p, sigma_p, alpha_p, beta_p, learning_rate, num_iters)


Optimal m: 10.1042
Optimal s^2: 0.0132560920497
Optimal alpha: 11.7797199472
Optimal beta: 0.122972674671
Expected Value for mu: 10.1042
Optimal Value for sigma^2: 1.44858366877

Using Variational Inference for such a simple example was an overkill. It might seem that Variational Inference is very complicated as compared to Maximum Likelihood Estimation. On this particular example it actually is a lot more complicated without giving any apparent improvement in performance. However in general when distributions become very complex, Variational Inference is the way to go. We will demonstrate this using more complex examples in the upcoming sections.

Intermediate Example 1 - Gaussian Distrbution

Computing the ELBO requires an expression for $\mathbb{E}_{z \sim q(z)}[\log q(z)−\log p(x,z)]$. In most cases obtaining an analytical form for this expectation is very challenging. The usual approach followed in such cases is to approximate the expected value.

At the very extreme, one can approximate this expected value by using only one sample of $z \sim q(z)$ by setting $ELBO(q) = \log q(z) - \log p(x, z)$. One can also resort to a more general monte-carlo approximation by drawing some $K$ number of samples from $q(z)$. In this experiment we study the effect of $K$ on the rate of learning.

We first demonstrate the case with $K = 1$. The code for this experiment can found in the file intermediate_example_1.py. We sample $1000$ one dimensional gaussian points using gm_generate.py. The setting here is slightly different from the setting that we had for Toy Example 1. In this case we assume the prior over the variance to be a squared-normal distribution instead of a gamma distribution. The reason for this is that it is easy to reparameterise the gaussian distribution as compared to the gamma distribution.

More details about reparameterisation can be found in Kingma et al. (2014). The basic idea is that we need reparameterisation because the gradients can not flow through a random node. Another point worth mentioning is that for a gamma distribution reparameterisation is possible but not straightforward. There are different approaches to do it, but those are beyond the scope of this project. An interested reader can refer to Chaney (2015), Knowles (2015) and Ruiz et al. (2016).


In [8]:
# Approximating ELBO using a single sample
from intermediate_example_1 import *

data_mean = 0.0
data_variance = 1.0
num_examples = 1000
num_iters = 1000

# Hyperparameters
mu_p = 0.0
sigma_p = 1.0
mu_var_p = 0.0
sigma_var_p = 1.0
learning_rate=1e-2
K = 1

# Run the demo
run_intermediate_example_1(num_examples, data_mean, data_variance, mu_p, sigma_p, mu_var_p, sigma_var_p, learning_rate, num_iters, K, num_examples)


Optimal m: -0.0359369
Optimal s^2: 0.0386445196165
Optimal m var: 0.987701
Optimal s^2 var: 0.0654398348638
Expected Value for mu: -0.0359369
Optimal Value for sigma^2: 0.975552556538

ELBO exhibits an oscillatory behavior due to approximation of expectation using a single sample.It takes about 400 iterations to reach a stable value for ELBO. Next we experiment with K=10 keeping everything else same as before.


In [9]:
# Approximating ELBO using K=10
from intermediate_example_1 import *

data_mean = 0.0
data_variance = 1.0
num_examples = 1000
num_iters = 1000

# Hyperparameters
mu_p = 0.0
sigma_p = 1.0
mu_var_p = 0.0
sigma_var_p = 1.0
learning_rate=1e-2
K = 10

# Run the demo
run_intermediate_example_1(num_examples, data_mean, data_variance, mu_p, sigma_p, mu_var_p, sigma_var_p, learning_rate, num_iters, K, num_examples)


Optimal m: 0.0366572
Optimal s^2: 0.0117142236635
Optimal m var: 1.00175
Optimal s^2 var: 0.0567927684518
Expected Value for mu: 0.0366572
Optimal Value for sigma^2: 1.00350830161

It can be seen that this time we get a stable ELBO value in less that 200 iterations. One can play around with parameter $K$ in the code given above and empirically verify that as $K$ becomes larger the number of iterations taken to converge to a stable value of ELBO decreases before it gets plateaued out. It is expected since the monte-carlo estimates get better as the number of samples increase.

Intermediate Example 2 - Mixture of Gaussians

In this section we will see how Variational Inference can be used to model given data using a Gaussian Mixture Distribution. In the beginning of this notebook we discussed about Gaussian Mixture Distribution. For the purpose of this example the input is a set $X = \{x^{(1)}, x^{(2)}, ..., x^{(N)}\}$ where $\forall i\,\, x^{(i)} \in \mathbb{R}$. That is, for pedagogical purpose, we will start with scalar data. We model $X$ as a set of iid data points generated from a Gaussian Mixture with $M$ components ($M$ is a known constant).

We associate a latent random variable $z^{(i)}$ with each of the observed random variable $x^{(i)}$. For each $i$, $z^{(i)}$ denotes the mixture component to which $x^{(i)}$ belongs. We encode $z^{(i)}$ using $1-of-K$ encoding. Say for example that $x^{(i)}$ belongs to $k^{th}$ mixture component, then $z^{(i)}_k = 1$ and $z^{(i)}_j = 0, \,\, \forall j \neq k$. Let the mean and variance of the $j^{th}$ mixture component be given by $(\mu_j, \sigma_j^2)$. If we define $\mu = [\mu_1, \mu_2, ..., \mu_M]^\intercal$ and $\sigma^2 = [\sigma_1^2, \sigma_2^2, ..., \sigma_M^2]^\intercal$, then the probability of point $x^{(i)}$ conditioned on $z^{(i)}$ is given by:

\begin{equation} P(x^{(i)} | z^{(i)}, \mu, \sigma^2) = \mathcal{N}(x^{(i)} | z^{(i)^\intercal} \mu, z^{(i)^\intercal} \sigma^2) \end{equation}

Of course we do not know what $z^{(i)}$'s are. If that information were known then maximizing the log probability of observed data with respect to parameters $\mu$ and $\sigma^2$ will be a trivial task. Usually the Expectation Maximimation algorithm is employed to perform maximum likelihood estimation of parameters in models with latent variables. However the EM algorithm assumes that the expected value of $P(X, Z)$ under the true posterior distribution can be computed exactly in closed form. While this assumption holds for the case of our current example, in general computing the aforementioned closed form expression is not tractible. Variational Inference does not need expectation under the exact posterior and hence is more general than EM.

As before, we will proceed by assuming parameteric form for approximating the true posterior distribution in order to get an expression for ELBO. The true posterior is given by $P(Z, \mu, \sigma^2 | X)$ where $Z = \{z^{(1)}, z^{(2)}, ..., z^{(N)}\}$. We will approximate this distribution by $q(Z, \mu, \sigma^2)$. By mean field assumption, $q(.)$ factorises in to $q(Z, \mu, \sigma^2) = q_1(Z) q_2(\mu) q_3(\sigma^2)$. Further, since the data points are drawn iid $q_1(.)$ factorises into $q_1(Z) = \prod_{i=1}^N q_{1i}(z^{(i)})$. We make one more simplifying assumption that the mean and variance of each mixture component are independent of the mean and variance of other mixture components. Thus $q_2(.)$ and $q_3(.)$ also factorise to $q_2(\mu) = \prod_{j=1}^M q_{2j}(\mu_j)$ and $q_3(\sigma^2) = \prod_{j=1}^M q_{3j}(\sigma^2_j)$ respectively. Finally we get:

\begin{equation} q(Z, \mu, \sigma^2) = \big( \prod_{i=1}^N q_{1i}(z^{(i)}) \big) \big( \prod_{j=1}^M q_{2j}(\mu_j) \big) \big( \prod_{j=1}^M q_{3j}(\sigma^2_j) \big) \end{equation}

We model $q_{1i}(z^{(i)})$ by a multinomial distribution. More specifically $q_{1i}(z^{(i)} = e_j) = \pi_{ij}$ such that $\forall i, j \,\, \pi_{ij} \geq 0$ and $\forall i\,\, \sum_{j=1}^M \pi_{ij} = 1$. The distribution over mean variables $q_{2j}(\mu_j)$ is modelled by a normal distribution i.e. $q_{2j}(\mu_j) = \mathcal{N}(\mu_j | m_j, s_j^2)$. Finally, we model $q_{3j}(\sigma_j^2) = ln \mathcal{N}(\sigma_j^2 | \alpha_j, \beta^2_j)$, where $ln \mathcal{N}(\alpha, \beta^2)$ is a log-normal distribution with mean $\alpha$ and scale parameter $\beta^2$. Thus we have a total of $NM + 4M$ variational parameters that should be learned from data.

To complete the specification of the problem we must also specify the prior distribution over $z^{(i)}$s, $\mu_j$s and $\sigma^2_j$s. We assume a uniform prior over $z^{(i)}$s, a standard normal distribution prior over $\mu_j$s and a standard log-normal prior over $\sigma_j^2$s. So we have:

\begin{align} P(z^{(i)}) &= 1/M \\ P(\mu_j) &= \mathcal{N}(\mu_j| 0, \sigma_\mu^2); \,\,\,\, P(\mu) = \prod_{j=1}^M P(\mu_j) \\ P(\sigma^2_j) &= ln \mathcal{N}(\sigma^2_j | 0, \sigma_\sigma^2); \,\,\,\, P(\sigma^2) = \prod_{j=1}^M P(\sigma^2_j) \end{align}

Note that our prior also assumes that the data points are drawn iid and that the mean and variance of different mixture components are independent of the mean and variance of other mixture components.

Now that we have $P(x^{(i)} | z^{(i)}, \mu, \sigma^2)$, $P(z^{(i)})$, $P(\mu)$, $P(\sigma^2)$ and $q(Z, \mu, \sigma^2)$ we can obtain an expression for the term inside the expectation in ELBO. More specifically we can obtain the expression for $\log P(X, Z, \mu, \sigma^2) - \log q(Z, \mu, \sigma^2)$.

\begin{align} \log P(X, Z, \mu, \sigma^2) - \log q(Z, \mu, \sigma^2) &= K - \frac{1}{2\sigma_\mu^2} \sum_{j=1}^{M} \mu_j^2 - \frac{1}{2\sigma_\sigma^2} \sum_{j=1}^{M} (\log \sigma_j^2)^2 -\\ & \frac{1}{2} \sum_{i=1}^{N} \sum_{j=1}^{M} z_j^{(i)} \log \sigma_j^2 -\frac{1}{2} \sum_{i=1}^{N} \sum_{j=1}^{M} \pi_{ij} \frac{(x_i - \mu_j)^2}{\sigma_j^2} + \\ & \frac{1}{2} \sum_{j=1}^{M} \log s_j^2 + \frac{1}{2} \sum_{j=1}^{M} \frac{(\mu_j - m_j)^2}{s_j^2} + \frac{1}{2} \sum_{j=1}^{M} \log \beta_j^2 + \\ & \frac{1}{2} \sum_{j=1}^{M} \frac{(\log \sigma_j^2 - \alpha_j)^2}{\beta_j^2} - \sum_{i=1}^{N} \sum_{j=1}^{M} \pi_{ij} \log \pi_{ij} \end{align}

In the equation given above, $K$ contains all the terms that are independent of the variational parameters, $\mu_j$s and $\sigma_j$s. We can write the ELBO as:

\begin{equation} ELBO(q) = \mathbb{E}_{Z, \mu, \sigma^2 \sim q} \big[\log P(X, Z, \mu, \sigma^2) - \log q(Z, \mu, \sigma^2)\big] \end{equation}

We can either calculate an analytical expression for this expectation or we can use Monto-Carlo approximation as was done in Intermediate Example 1 to get an approximate value for ELBO.

The code which implements the minimization of ELBO using Monte-Carlo approximation to the expectation is present in intermediate_example_2.py. In the demo that follows we use a mixture of three 1D gaussians and use $1$ sample to approximate the expectation. The problem setup is same as the one used in Intermediate Example 1.


In [13]:
# Variational Inference for 1D Mixture of Gaussians
from intermediate_example_2 import *

data_mean = [-6.0, 2.0, 6.0]
data_variance = [3, 1, 2]
num_examples = 1000
num_iters = 1000

# Hyperparameters
sigma_mu = 1.0
sigma_sigma = 1.0
learning_rate=2*1e-2
K = 1
M = 3

# Run the demo
run_intermediate_example_2(data_mean, data_variance, sigma_mu, sigma_sigma, learning_rate, num_iters, K, num_examples, M)


Mean for mu: [-5.84040117  5.77850294  1.93057168]
Mean for sigma: [ 1.2342416   1.01328278 -0.14440267]
Variance for mu: [ 0.08176422  0.0710019   0.03328886]
Variance for sigma: [ 0.05227579  0.05723557  0.05811854]
Optimal mu: [-5.84040117  5.77850294  1.93057168]
Optimal s^2: [ 3.52675986  2.83459902  0.89106011]

One can again play around with the hyperparameters in the code given above. To the best of our knowledge this is the first time where a log-normal prior has been used for the variance term. Generally people use the Gamma distribution as a prior. The problem with Gamma distribution is that it is very hard to reparameterise it effeciently. Hence we can not use the Monte-Carlo approximation to expectation if a Gamma prior is used. On the other hand it is very easy to reparameterise the log-normal distribution.

For the sake of comparison we present the result of running the Expectation Maximization algorithm on the same type of data. Both EM and Variational Inference may lead to sub-optimal results because of being stuck at local minimas. These results which look good were obtained by running the code multiple number of times using random initialization. However it must be noted that it takes a lot more number of tries in the case of EM as compared to Variational Inference.


In [53]:
# EM for 1D Mixture of Gaussians
from em import *
from gm_generate import *

# Hyperparameters
data_mean = [-6.0, 2.0, 6.0]
data_variance = [3, 1, 2]
num_iters = 20
M = 3
N = 1000

# Run the example
run_em(data_mean, data_variance, num_iters, N, M)


Mu: [-5.82042636  6.00467566  2.00050195]
Sigma: [ 3.75729544  1.92194565  0.66294103]