Introduction to Statistics for Astronomers and Physicists

Section 3c: Posterior Analysis II

Dr Angus H Wright

2022-02-09

Section 1: Introduction

Section 3: Bayesian Statistics (Weeks 7-9)

Bayes theorem led to a revolution in statistics, via the concepts of prior and posterior evidence. In modern astronomy and physics, applications of Bayesian statistics are widespread. We begin with a study of Bayes theorem, and the fundamental differences between frequentest and Bayesian analysis. We explore applications of Bayesian statistics, through well studied statistical problems (both within and outside of physics).

Where we finished last time

Summarising Posterior Information

The posterior distribution is a complete summary of the inference about \(\theta\). In essence, the posterior distribution is the inference about \(\theta\). However, for many applications, we wish to summarise the information contained within the posterior into some digestible quantity.

Point Estimates

We can do this by specifying some form of Loss Function \(L(\theta,a)\), which measures the penalty in estimating a value of \(\theta\) given an individual datum at \(a\).

There are a range of potential loss functions that are common to use, and the particular choice of the loss function will often depend on the problem being analysed. The most common loss functions (some of which we have already seen!) are:

  1. Quadratic Loss (minimised by mean): \(L(\theta,a)=(\theta-a)^2\)
  2. Absolute Error Loss (minimised by median): \(L(\theta,a)=|\theta-a|\)
  3. \(0-1\) Loss (minimised by mode): \[ L(\theta,a)= \begin{cases} 0 & \textrm{if}\; |a-\theta|\leq \epsilon \\ 1 & \textrm{if}\; |a-\theta|> \epsilon \\ \end{cases} \]
  4. Linear Loss: for specified \(g,h>0\) \[ L(\theta,a)= \begin{cases} g(a-\theta) & \textrm{if}\; a> \theta \\ h(a-\theta) & \textrm{if}\; a\leq \theta \\ \end{cases} \]

Credability Intervals

The idea of credability intervals is analagous to confidence intervals in classical statistics. The reasoning is that point estimates give no measure of accuracy, so it is prefereable to specify the region in which a parameter is likely to reside.

Predictive Distributions

So far we have focussed on parameter estimation. We’ve shown how the Bayesian framework combines sample information and prior information to give parameter inference in the form of the posterior distribution. However commonly we are interested in making predictions about the future.

In making a prediction, on the basis of an estimated model, there are two sources of uncertainty:

  1. Uncertainty in the parameter values which have been assigned on the basis of previous data; and
  2. Uncertainty due to the fact that any future value of the parameter is itself a random event.

In classical statistics, it is standard practice to fit a model to the existant data and them make predictions assuming that the model is correct. There is only the second source of uncertainty included in this analysis. This leads to predictions that are believed to be more precise than they really are. There is simply, however, no satisfactory way around this problem in the framework of classical statistics, because parameters are considered constant in the likelihood.

In the Bayesian framework, however, it is straightforward to allow for both sources of uncertainty by averaging over the uncertainty in the parameter estimates themselves.

Suppose we have past observations \(x=\{x_1,\dots,x_n\}\) of a variable with the density function (or likelihood) \(p(x|\theta)\), and we wish to make inference about the distribution of any future value \(y\) from this process.

With a prior \(p(\theta)\), Bayes theorem gives us the posterior \(p(\theta|x)\), and then the predictive density function of \(y\) given \(x\) is: \[ p(y|x)=\int p(y|\theta)p(\theta|x)\textrm{d}\theta \] that is, the integral of the likelihood of a single point times the posterior.

Though simple in principle, computation of the posterior predictive distribution can again be difficult. However once again there are some conjugate families that make generating posterior predicitve distributions tractable. For others, these can be computed numerically (but at cost).

Predicting a coin toss

Consider our Binomial coin toss with unknown bias \(\theta\): \(x\sim\textrm{Bin}(n,\theta)\). The conjugate prior to the binomial likelihood is \(p(\theta)\sim \textrm{Be}(p,q)\), giving the posterior \(\theta|x\sim\textrm{Be}(p+x,q+n-x)\).

Suppose we intend to make \(N\) further observations in the future, and we let \(z\) be the number of successful throws out of the \(N\) trials.

\[ z|\theta\sim\textrm{Bin}(N,\theta) \\ \therefore p(z|\theta)={N\choose z} \theta^z(1-\theta)^{N-z} \] We can compute the predictive distribution of \(z\) given our observations using the posterior distribution: \[ \begin{align} p(z|x)&=\int p(z|\theta)p(\theta|x)\textrm{d}\theta\\ &=\int_0^1 {N \choose z}\theta^z(1-\theta)^{N-z} \frac{\theta^{p+x-1}(1-\theta)^{q+n-x-1}}{\textrm{Be}(p+q,q+n-x)}\textrm{d}\theta\\ &= {N\choose z}\frac{1}{\textrm{Be}(p+q,q+n-x)}\int_0^1\theta^{p+x-1}(1-\theta)^{q+n-x-1}\textrm{d}\theta \\ \\ \therefore p(z|x)&= {N \choose z} \frac{\textrm{Be}(z+p+q,q+2n-x-z)}{\textrm{Be}(p+q,q+n-x)}. \end{align} \]

Remember back to our first exploration into Bayes Theorem with our three scenarios (Section 3a). We can now predict the outcomes of a subsequent \(10\) trials for each scenario given the data (\(x=10,n=10\)) and our priors.

p_zx<-function(z=0:10,N=10,x=10,n=10,p,q) { 
  return=choose(N,z)*beta(z+p+q,q+2*n-x-z)/beta(p+q,q+n-x)
}
magplot(0:10,p_zx(p=0.1,q=100),xlim=c(0,10),ylim=c(0,1),type='l',lwd=2,col='red3',ylab=expression(p(z*"|"*x)),xlab='z')
lines(0:10,p_zx(p=0.7,q=1),col='green3',lwd=2,lty=2)
lines(0:10,p_zx(p=10,q=0.5),col='blue3',lwd=2)
legend('topleft',legend=c("Fortune Teller","Wine Critic","Classical Musician"),
       lty=c(1,2,1),lwd=2,col=c("red3","green3","blue3"),
       bty='n',inset=0.05)

Posterior Simulation

The last, arguably most popular/useful method of posterior interpretation is posterior simulation. This is related to the long discussion that we had on simulation and sampling from unknown distributions three lectures ago (Section \(2d\)).

Bayesian Posteriors and Sampling

As we’ve already seen, even simple combinations of priors and posteriors can lead to complex posteriors, if we can’t use a conjugate analysis.

As a result, using a posterior to do parameter inference requires some means of describing the (often complex) posterior in some digestible manner. This generally means using point statistics and/or credability intervals, both of which require us to understand the distribution that we want to analyse (that is, to know the PDF). But we’ve already demonstrated that knowing the analytic form of the posterior PDF can be difficult if not impossible in most cases. Fortunately, sampling comes to the rescue.

Posterior Simulation, and Markov Chain Monte Carlo

More than any other technique, Markov Chain Monte Carlo has been responsible for the current resurgence of Bayesian Statistics in the natural sciences. This is because MCMC allows us to estimate a vast array of Bayesian models with ease.

The idea of MCMC was first introduced as a method for the efficient simulation of energy levels of atoms in crystalline lattices. It was subsequently adapted for broader use within statistics.

The concept of MCMC is as follows:

Suppose we have some arbitrary “target distribution\(\pi\): \[ \pi(x),\;\; x\in \Omega\in \mathbb{R}^p. \]

If \(\pi\) is sufficiently complex that we are unable to sample from it directly, then an indirect method for sampling from it is to construct a Markov Chain within a state space \(\Omega\), whose “stationary distribution” is \(\pi(x)\). If we run the chain for long enough, simulated values from the chain can be treated as samples from the target distribution and used as a basis for summarising the important features of \(\pi\).

There is a lot of jargon in that last paragraph, but don’t fret. We will make this clear in the following slides.

The Markov Chain

What is a markov chain?

A sequence of random variables \(X_t\) is defined as a Markov Chain if it follows the conditional probability: \[ P(X_{t+1}=x_{t+1}|X_t=x_t,X_{t-1}=x_{t-1},\dots,X_0=x_0) = P(X_{t+1}=y|X_t=x_t) \] That is, it is a sequence of numbers that, given the current state, is independent of the past. The probability of transition from state \(x=x_{t}\) to \(y=x_{t+1}\) is given by some transition probability.

Transition Probability

The \(n^{\textrm{th}}\) step transition probability \(P^n(x,y)\) of a Markov Chain moving from state \(x\) to state \(y\) in precisely \(n\) steps is given by: \[ \begin{align} P^n(x,y) &= P(X_n=y|X_0=x) \\ &= \sum_{x_1}\dots\sum_{x_{n-1}} P(x,x_1)P(x_1,x_2)\dots P(x_{n-1},y) \end{align} \] More generally, the transition probability can be visualised, for a finite state space \(\Omega\) as an \(|\Omega|\times|\Omega|\) matrix of probabilities of transitioning from state \(x\) to state \(y\): \[ P(x,y) = P(X_1=y|X_0=x);\;\; x,y\in\Omega. \]

All entries in \(P(x,y)\) are therefore probabilities (i.e. \(P(x,y)\in[0,1]\)), and the rows of \(P(x,y)\) must sum to unity (because from every point \(X_0=x\), you have to go somewhere in \(\Omega\)).

Visualising the Markov Chain

You may already be familiar with the concept of the Markov chain, even if you don’t know it by that name specifically. This is because there is a special type of Markov Chain that is somewhat well known: the random walk.

A random walk is a Markov chain whereby the transition probability is uniform for the points adjacent to \(x\): \[ P(x,y)= \begin{pmatrix} 0 & 1 & 0 & 0 & \dots & 0 & 0 & 0 & 0\\ 0.5 & 0 & 0.5 & 0& \dots & 0 & 0 & 0 & 0\\ 0 & 0.5 & 0 & 0.5 & \dots & 0 & 0 & 0 & 0\\ \vdots & \vdots & \vdots & \vdots & \ddots & \vdots & \vdots & \vdots & \vdots\\ 0 & 0 & 0 & 0 & \dots & 0.5 & 0 & 0.5 & 0\\ 0 & 0 & 0 & 0 & \dots & 0 & 0.5 & 0 & 0.5\\ 0 & 0 & 0 & 0 & \dots & 0 & 0 & 1 & 0 \\ \end{pmatrix} \]

In a program, we can code up such a sequence trivially:

set.seed(1904)
step<-function(x) return(x+sample(c(-1,1),size=1))
step(0)
## [1] -1
step(0)
## [1] 1

A “random walk” is therefore just a sequence of steps:

rand.walk<-function(x0=0,n=100) {
  x<-rep(NA,n+1)
  x[1]<-x0
  for (i in 1:n) {
    x[i+1]<-step(x[i])
  }
  return(x)
}
plot(0:100,rand.walk(n=100),xlab='step number',ylab='location',type='s')

Simplifying this code into a (faster) implementation:

rand.walk<-function(x0=0,n=100) {
  x<-cumsum(c(x0,sample(c(-1,1),size=n,replace=TRUE)))
  return(x)
}
plot(0:1E5,rand.walk(n=1E5),xlab='step number',ylab='location',type='s')

We can also run many of the Markov Chains from the same starting point:

walkers<-replicate(30,rand.walk(n=1E5))
plot(0:1E5,ylim=range(walkers),xlab='step number',ylab='location',type='n')
for (i in 1:ncol(walkers)) { 
  lines(0:1E5,walkers[,i],type='s',col=hsv(v=0,a=0.4))
}

Using this random walk, we can now explore the transition probability we discussed earlier: \[ \begin{align} P^n(x,y) &= P(X_n=y|X_0=x) \\ &= \sum_{x_1}\dots\sum_{x_{n-1}} P(x,x_1)P(x_1,x_2)\dots P(x_{n-1},y) \end{align} \] Analytically, this probability requires us to sum over all possible values of \(x_1, x_2, \dots, x_{n-1}\). One question we might want to ask, though, is: what is the probability of returning to my starting point as a function of the number of steps?

We can use our walker script will allow us to calculate this probability relatively trivially:

walk.home<-function(x0=0,...) { 
  return(
    ifelse(any(rand.walk(x0=x0,...)[-1]==x0),1,0)
  )
}
arrived<-replicate(30,walk.home(n=1E2))
print(sum(arrived)/length(arrived))
## [1] 0.9333333

We can test this probability for a range of numbers of steps:

prob<-NULL
steps<-10^seq(0,5,len=20)
for (step_len in steps) { 
  arrived<-replicate(500,walk.home(n=step_len))
  prob<-c(prob,sum(arrived)/length(arrived))
}
magplot(steps,prob,ylim=c(0,1),log='x',xlab='step number',
ylab='probability of arriving home',type='l')

So with enough steps in \(1D\), a random walk is \(\sim\)guaranteed to make it home.

Higher Dimensions

What about in the more realistic case of 2D?

Do you think that:

  1. There’s a guarantee of returning home, similar to \(1D\)? or
  2. The probability will plateau at \(P(\textrm{arriving home})<1\)?

Higher Dimensional Random Walk

We can further generalise our random walk script to N dimensions fairly easily:

rand.walk.nD<-function(x0=0,nStep=100,nD=1) {
  #Define the nStep steps as ±1
  steps<-sample(c(-1,1),size=nStep,replace=TRUE)
  #Select the axis in which each step occurs
  dims<-sample(1:nD,size=nStep,replace=TRUE)
  #Setup the walk coordinates
  x<-matrix(0,ncol=nD,nrow=nStep)
  #Loop over each dimension
  for (i in 1:nD) { 
    #For this dimension, assign the steps
    x[dims==i,i]<-steps[dims==i]
    #Calculate the cumulative sum
    x[,i]<-cumsum(x[,i])
  }
  #Return
  return(x)
}

We can then use our new function to plot random walks in \(2D\). First, with few steps per chain:

#Plot a few walks with various steps
set.seed(666)
layout(cbind(1,2))
magplot(rand.walk.nD(nStep=1e2,nD=2),xlab='location x',ylab='location y',
     type='l',col=hsv(v=0,a=0.3),xlim=c(-15,15),ylim=c(-15,15),asp=1,lwd=2)
lines(rand.walk.nD(nStep=1E2,nD=2),col=hsv(h=2/3,v=0.8,a=0.3),lwd=2)
lines(rand.walk.nD(nStep=1E2,nD=2),col=hsv(h=0,v=0.8,a=0.3),lwd=2)
magplot(rand.walk.nD(nStep=1E2,nD=2),xlab='location x',ylab='location y',
     type='l',col=hsv(v=0,a=0.3),xlim=c(-15,15),ylim=c(-15,15),asp=1,lwd=2)
lines(rand.walk.nD(nStep=1E2,nD=2),col=hsv(h=2/3,v=0.8,a=0.3),lwd=2)
lines(rand.walk.nD(nStep=1E2,nD=2),col=hsv(h=0,v=0.8,a=0.3),lwd=2)

Adding more steps:

#Plot a few walks with various steps
set.seed(666)
layout(cbind(1,2))
magplot(rand.walk.nD(nStep=1E4,nD=2),xlab='location x',ylab='location y',
     type='l',col=hsv(v=0,a=0.3),xlim=c(-100,100),ylim=c(-100,100),asp=1,lwd=2)
lines(rand.walk.nD(nStep=1E4,nD=2),col=hsv(h=2/3,v=0.8,a=0.3),lwd=2)
lines(rand.walk.nD(nStep=1E4,nD=2),col=hsv(h=0,v=0.8,a=0.3))
magplot(rand.walk.nD(nStep=1E4,nD=2),xlab='location x',ylab='location y',
     type='l',col=hsv(v=0,a=0.3),xlim=c(-100,100),ylim=c(-100,100),asp=1,lwd=2)
lines(rand.walk.nD(nStep=1E4,nD=2),col=hsv(h=2/3,v=0.8,a=0.3),lwd=2)
lines(rand.walk.nD(nStep=1E4,nD=2),col=hsv(h=0,v=0.8,a=0.3),lwd=2)

And more again:

#Plot a few walks with various steps
set.seed(666)
layout(cbind(1,2))
magplot(rand.walk.nD(nStep=1E5,nD=2),xlab='location x',ylab='location y',lwd=2,
     type='l',col=hsv(v=0,a=0.3),xlim=c(-350,450),ylim=c(-300,200),asp=1)
lines(rand.walk.nD(nStep=1E5,nD=2),col=hsv(h=2/3,v=0.8,a=0.3),lwd=2)
lines(rand.walk.nD(nStep=1E5,nD=2),col=hsv(h=0,v=0.8,a=0.3),lwd=2)
magplot(rand.walk.nD(nStep=1E5,nD=2),xlab='location x',ylab='location y',
     type='l',col=hsv(v=0,a=0.3),xlim=c(-200,200),ylim=c(-300,300),asp=1,lwd=2)
lines(rand.walk.nD(nStep=1E5,nD=2),col=hsv(h=2/3,v=0.8,a=0.3),lwd=2)
lines(rand.walk.nD(nStep=1E5,nD=2),col=hsv(h=0,v=0.8,a=0.3),lwd=2)

We can see that these \(6\) random walks all meander around the plane.

And so now we can ask the question of whether we ever return to the origin in 2D with a random walk as well. We can generate many random walks, as before, and find the number that return to the origin:

library(matrixStats)
#Define our new walk home function
walk.home.nD<-function(x0=0,...) { 
  return(
    ifelse(any(rowAlls(rand.walk.nD(x0=x0,...)[-1,,drop=FALSE]==x0)),1,0) 
  )
}
#Compute the probability of returning home
steps<-10^seq(1,5,len=5)
prob<-matrix(NA,nrow=length(steps),ncol=3)
for (nD in 1:3) { 
  for (i in 1:length(steps)) { 
    arrived<-replicate(1e3,walk.home.nD(nStep=steps[i],nD=nD))
    prob[i,nD]<-sum(arrived)/length(arrived)
  }
}
magplot(steps,xlim=range(steps),ylim=c(0,1),log='x',xlab='step number',
ylab='probability of arriving home',type='n')
colours<-c("blue3","black","red3")
for (nD in 1:3) { 
  lines(steps,prob[,nD],lwd=2,col=colours[nD])
}
legend('bottom',legend=paste0(1:3,"D"),col=colours,lwd=2,bty='n',inset=0.02)

If you expand the figure out to much higher numbers of steps, it turns out that you do return home with certainty (as \(n_{\textrm{step}}\rightarrow\infty\)) in both one and two-dimensions. However in 3D, you only have a \(\sim35\%\) chance of making it home.

The drunken human is much more likely to chance upon their home than is the drunken fish!

Important Markov Chain Properties

There are two primary properties of a Markov chain that are important to be aware of, as they will become important later.

Aperiodicity

The period \(dx\) of a particular chain state \(x\) is formally the largest common divisor of the set \(\{n\geq 1: P^n(x,x)>0\}\); that is, the value that describes the rate at which a chain is able to revisit a state \(x\) given that it is currently at state \(x\).

The Markov Chain is aperiodic if \(dx=1\): the chain is able to return to t state \(x\) without any formal waiting/cycling period. Note that
\(P^n(x,x)>0\) implies that \(dx=1\). If a chain is periodic, then it is only able to revisit certain parts of the state space at finite intervals.

Irreducibility

A markov chain is said to be \(\pi\)-irreducible if, for every \(x,y\in\Omega\): \[ \pi(y)>0\rightarrow P(x_n=y|x_0=x)>0. \] Irreducibility means that the chain can move from anywhere in the state space (\(x\)) to any other point in the state space (\(y\)) in a finite number of steps (\(n\)).

Building an MCMC Sampler: The Gibbs Sampler

The first MCMC sampler that we are going to look at is the Gibbs Sampler. Consider a pair of variables \(X,Y\), whose joint distribution is denoted by \(\pi(x,y)\).

The Gibbs Sampler generates a sample from \(\pi(x)\), i.e. the marginal density of \(\pi\) with respect to x, by sampling (in turn) from the conditional distributions \(\pi(x|y)\) and \(\pi(y|x)\), which frequently known in statistical models of complex data. This is done by generating a “gibbs sequence” of random variables: \[ Y_0^\prime,X_0^\prime,Y_1^\prime,X_1^\prime,\dots,Y_N^\prime,X_N^\prime \] The initial value \(Y_0^\prime\) is chosen arbitrarily, and all other values are found iteratively by generating values from: \[ X^\prime_t \sim \pi(x|Y^\prime_t) \\ Y^\prime_{t+1} \sim \pi(x|X^\prime_t) \] This is known as Gibbs Sampling.

It turns out that, under reasonably general conditions, the distribution of \(X^\prime_N\) converges to \(\pi(x)\) as \(N\rightarrow\infty\). Said differently, provided \(N\) is large enough, the final observation of \(X^\prime_N\) is effectively a sample from \(\pi(x)\).

The Gibbs Sampler: Algorithm

  1. Initialise \(\vec{X}=\left(X_0^{(1)},\dots,X_0^{(d)}\right)\)
  2. Simulate \(X_1^{(1)}\) from the conditional distributions of \[ \pi(X^{(1)}|X_0^{(2)},\dots,X_0^{(d)}) \]
  3. Simulate \(X_1^{(2)}\) from the conditional distributions of \[ \pi(X^{(2)}|X_1^{(1)},X_0^{(3)},\dots,X_0^{(d)}) \]
  4. Iterate

Demonstration: Bivariate Gaussian

Consider a single observation \((y_1,y_2)\) from a bivariate normally distributed population with unknown mean \(\theta=(\theta_1,\theta_2)\) and known covariance matrix: \[ \begin{pmatrix} 1 & \rho \\ \rho & 1 \\ \end{pmatrix} \]

With a uniform prior distribution on \(\theta\), the posterior distribution is \[ {\theta_1 \choose \theta_2}|y\sim N\left({y_1 \choose y_2}, \begin{pmatrix} 1 & \rho \\ \rho & 1 \\ \end{pmatrix} \right) \] Although we can sample from this directly, let’s pretend that we cannot.

To apply the Gibbs sampler we must first know the form of the conditional posterior distributions. From the properties of the multivariate normal distribution these are: \[ \theta_1|\theta_2,y\sim N(y_1+\rho(\theta_2-y_2),1-\rho^2) \\ \theta_2|\theta_1,y\sim N(y_2+\rho(\theta_1-y_1),1-\rho^2) \\ \]

Let’s set \(\rho=0.8\), and \((y_1,y_2)=(0,0)\), and use the initial guess \(X_0^1,X_0^2=(\pm 2.5,\pm 2.5)\) (that is, we’re running 4 distinct chains).

#Gibbs Sampler in R 
Gibbs<-function(start,n=1e3,rho=0.8) {
  #Initialise the variables 
  X<-Y<-rep(NA,n)
  #Enter the initial guesses 
  X[1]<-start[1]
  Y[1]<-start[2]
  #Loop over the next n steps
  for (i in 2:n) { 
    #Generate the sample for X
    X[i]<-rnorm(1,rho*Y[i-1],sqrt(1-rho^2))
    Y[i]<-rnorm(1,rho*X[i],sqrt(1-rho^2))
  }
  return(cbind(X,Y))
}

Let’s start by plotting the first \(15\) samples from each of our Gibbs sequences:

#Define the initial guesses 
set.seed(666)
starts<-expand.grid(c(-2.5,2.5),c(-2.5,2.5))
#Construct the Gibbs Samples 
samps<-{}
for (j in 1:nrow(starts)) 
  samps[[j]]<-Gibbs(start=unlist(starts[j,]),n=15)
#Plot the samples 
magplot(starts,type='p',xlab=expression(theta_1),ylab=expression(theta_2),
        xlim=c(-5,5),ylim=c(-5,5),lwd=2,cex=2)
colours<-c("red3","green3","orange","blue3")
for (i in 1:nrow(starts)){ 
  lines(samps[[i]],lty=1,lwd=2,col=seqinr::col2alpha(colours[i],0.8))
}

You can see our four starting guesses as the black circles, and the subsequent samples of the sequence.

Let’s now crank it up to many samples:

#Define the initial guesses 
set.seed(666)
starts<-expand.grid(c(-2.5,2.5),c(-2.5,2.5))
#Construct the Gibbs Samples 
samps<-{}
for (j in 1:nrow(starts)) 
  samps[[j]]<-Gibbs(start=unlist(starts[j,]))
#Plot the samples 
magplot(starts,type='p',xlab=expression(theta_1),ylab=expression(theta_2),
        xlim=c(-5,5),ylim=c(-5,5),lwd=2,cex=2)
colours<-c("red3","green3","orange","blue3")
for (i in 1:nrow(starts)){ 
  lines(samps[[i]],lty=1,lwd=2,col=seqinr::col2alpha(colours[i],0.2))
}

With so many samples the individual points now become more useful to plot:

#Define the initial guesses 
#Plot the samples without consideration of the order
magplot(starts,type='p',xlab=expression(theta_1),ylab=expression(theta_2),
        xlim=c(-5,5),ylim=c(-5,5),lwd=2,cex=2)
colours<-c("red3","green3","orange","blue3")
for (i in 1:nrow(starts)){ 
  points(samps[[i]],pch=20,cex=1,col=seqinr::col2alpha(colours[i],0.3))
}

The Metropolis-Hastings Algorithm

An alternative, more general, way to construct an MCMC sampler is as a form of generalised rejection sampling, where the values are drawn from appropriate approximate distributions and are “corrected” in order that they, asyptotically, behave as random observations from the target distribution.

This is the motivation for the Metropolis-Hastings Algorithm, which sequentially draws candidate observations from a distribution, conditional only on the last observation, thereby inducing a markov chain.

The important aspect of the algorithm is that the approximating candidate distribution is improved at each step in the simulation, if we chose to let it depend on th previous observations. This is distinct from rejection sampling (which we’ve already discussed) where the candidate distribution must remain the same.

The MH Algorithm

Set \(t=1\) and while \(t\leq N\):

  1. Generate \(y\) from \(Q(x_t,.)\) and \(U\) from \(U(0,1)\);
  2. let \(x_{t+1}=y\) if \(U\leq \alpha(x_t,y)\), otherwise let \(x_{t+1}=x_t\).
  3. \(t=t+1\); iterate.

where

This process produces samples from the target distribution.

Special Cases of \(Q(x,y)\)

Gibbs

The Gibbs Sampler is therefore a special case of the MH algorithm, where the candidate generating function \(Q(x,y)\) is the conditional distribution: \[ Q(x,y)=\pi(x|y) \] and the acceptance probability is unity for all \(x,y\): \[ \alpha(x,y)=1\;\;\forall (x,y). \]

Random-Walk Metropolis-Hastings

If we use a candidate generation function of the form: \[ Q(x,y)=f(|y-x|) \] for some arbitrary density function \(f\), the kernal driving the proposal is therefore a random walk, like what we explored previously. This is because the observation is of the form \(y_{t+1}=x_{t}+z\), where \(z\sim f\).

There are many common choices for \(f\), some of which you are probably already thinking about: the uniform distribution, multivariate normal, and Student t-distributions are all of this form, simple, and analytic, and so are very useful in this application.

Note also that, if \(f\) is symmetric: \[ f(z)=f(-z), \] then the acceptance probability is of the form: \[ \alpha(x,y)=\min\left(1,\frac{\pi(x)}{\pi(y)}\right), \] meaning that acceptance is determined simply by the ratio of target distribution densities at the current and proposed positions.

The Independence Sampler

If \(Q(x,y)=f(y)\), then the candidate location is drawn independent of the current state of the chain. In this case the acceptance probability can be written as: \[ \alpha(x,y)=\min\left(1,\frac{w(y)}{w(x)}\right), \] where \(w(x)=\pi(x)/f(x)\) is the importance weight function that would be used in importance sampling, given observations generated from \(f\). In this way importance sampling and independence sampling are not dissimilar: importance sampling builds up mass by frequently returning to the state \(x\), given a large \(w(x)\). The Independence sampler, however, builds up mass on \(x\) by not leaving.

MH Demonstration: Bivariate Gaussian

The target density is a bivariate unit normal distribution \(\pi(\theta|y)=N(\theta|0,I)\) where \(I\) is the \(2\times 2\) identity matrix. The proposal distribution is also a bivariate Gaussian, which is centred at the current iteration: \[ Q(\theta^t|\theta^{t-1})=N(\theta^\star|\theta^{t-1},0.2^I). \] This is therefore a random walk proposal with symmetric probability.

#Example of writing our own MH algorithm in R
#Load the "multivariate Normal" package
library(mvtnorm)
#Write our Bivariate Normal MH function 
bvnm.MH<-function(n,start) { 
  #Initialise the variables
  X<-Y<-rep(NA,n)
  X[1]<-start[1]
  Y[1]<-start[2]
  Id<-diag(2) # The identity matrix
  #Loop over our iterations 
  for (i in 2:n) { 
    #Generate a proposal point
    prop<-rmvnorm(1,c(X[i-1],Y[i-1]),0.2*Id)
    #Compute the acceptance probability pi(y)/pi(x)
    accept<-dmvnorm(prop,c(0,0),Id)/
            dmvnorm(c(X[i-1],Y[i-1]),c(0,0),Id)
    #Draw a random univariate 
    u<-runif(1)
    #Test the acceptance criteria
    if (u<accept) { 
      #Accept the new data point
      X[i]<-prop[1]
      Y[i]<-prop[2]
    } else { 
      X[i]<-X[i-1]
      Y[i]<-Y[i-1]
    }
  }
  return(cbind(X,Y))
}

Now that we have our MH algorithm set up, we can run a chain:

#Running our MH Code
chain<-bvnm.MH(1e3,start=c(-10,10))
magplot(chain,type='l',col='blue3',lwd=2,xlab=expression(theta[1]),
        ylab=expression(theta[2]))

Important Questions in MCMC

Iterative simulation from the posterior adds two primary difficulties to simulation inference. Namely:

  1. How long does a chain need to run before it is “long enough”?
  2. How many samples do we need to perform unbiased inference?

There are two primary ways of dealing with these questions.

Burn in

If the chain has not been run for a sufficiently large number of iterations, the simulated samples may be grossly unrepresentative of the posterior that we wish to infer. The long trail of samples that travel from the starting guess into the main cloud of samples is a principle example. These samples are known as the burn in, and should be discarded.

#Burn in 
layout(cbind(1,2))
magplot(chain[,1],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[1]))
magplot(chain[,2],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[2]))

Particularly for small numbers of samples (i.e. short chains) these burn-in samples can be pathological. However in all chains they clearly do not sample the target distribution, so should be discarded.

#Removing Burn in 
chain<-chain[-200:0,]
layout(cbind(1,2))
magplot(chain[,1],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[1]))
magplot(chain[,2],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[2]))

Thinning

We want our final posterior samples to be i.i.d observations from the target distribution. However, the iterative nature of the MCMC means that the sequence of draws is not independent.

#Correlation in Samples 
layout(cbind(1,2))
magplot(chain[-1,1],chain[-nrow(chain),1],pch=20,col='blue3',lwd=2,
        ylab=expression(theta[1]^"i+1"),xlab=expression(theta[1]^i),asp=1)
magplot(chain[-1,2],chain[-nrow(chain),2],pch=20,col='blue3',lwd=2,
        ylab=expression(theta[2]^"i+1"),xlab=expression(theta[2]^i),asp=1)

However we can circumvent this by thinning the samples to every \(n^{\textrm{th}\) sample (but of course this necessitates a longer chain).

#Independence in thinning
chain<-bvnm.MH(5e4,start=c(-10,10))
chain<-chain[-200:0,]
thin<-function(x,by) x[seq(1,nrow(x),by=by),]
layout(matrix(1:4,2,2,byrow=TRUE))
par(mar=c(3,4,1,1))
chain.fifth<-thin(chain,by=5)
chain.twent<-thin(chain,by=20)
chain.hundr<-thin(chain,by=100)
colour<-seqinr::col2alpha('blue3',0.3)
magplot(chain[-1,1],chain[-nrow(chain),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+1"),xlab=expression(theta[1]^i),asp=1)
magplot(chain.fifth[-1,1],chain.fifth[-nrow(chain.fifth),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+5"),xlab=expression(theta[1]^i),asp=1)
magplot(chain.twent[-1,1],chain.twent[-nrow(chain.twent),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+20"),xlab=expression(theta[1]^i),asp=1)
magplot(chain.hundr[-1,1],chain.hundr[-nrow(chain.hundr),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+100"),xlab=expression(theta[1]^i),asp=1)

So the thinning to every \(\sim 10^{\textrm{th}}\) sample is reasonably independent, without producing a significant decrease in the accepted number of samples (as with the \(100^{\textrm{th}}\) sample thinning.

A Bayesian MH Problem

Now that we have the MH algorithm formalised, we can try tackling a proper Bayesian problem. Let’s go back to our old standard problem, coin tosses, but with a bit of a twist.

We are given a bag of coins, and are told that a certain fraction of the coins are biased. We are not told how many coins are biased, nor what their biasing is. We draw a random coin from the bag \(1000\) times, toss it \(3\) times, and record the number of observed heads.

#Generate and Plot some data
set.seed(42)
data<-simulate.throws(1e3)
maghist(data,xlab='Number of Heads in 3 tosses',ylab='Count',verbose=FALSE)

We want to infer, given the observations, the likely values of the fraction of coins in the bag which are biased \(\theta_1\), and the amount of biasing that affects each biased coin \(\theta_2\). That is, we want to know the posterior distribution \(P(\theta_1,\theta_2|x)\).

Recall the procedure for performing Bayesian Inference:

  1. Specification of the likelihood model \(p(x|\theta)\);
  2. Determination of the prior \(p(\theta)\);
  3. Calculation of the posterior distribution \(p(\theta|x)\); and
  4. Draw inference from the posterior distribution.

The likelihood model

In our case, the likelihood model consists of \(3\) parts:

  1. The probability of drawing a biased coin from the bag
  2. The probability of a biased coin showing \(x\) heads in \(3\) trials
  3. The probability of a fair coin showing \(x\) heads in \(3\) trails

All of these probabilities can be described using binomials:

  1. \(\textrm{biased coin}|\theta_1 \sim \textrm{Bin}(1,\theta_1)\)
  2. \(x\;\textrm{heads}|\textrm{biased coin},\theta_2 \sim \textrm{Bin}(3,\theta_2)\)
  3. \(x\;\textrm{heads}|\textrm{fair coin} \sim \textrm{Bin}(3,0.5)\)

So we can construct our likelihood from these probabilities: \[ \begin{align} P(X|\theta_1,\theta_2) &= P(X|n,0.5)(1-\theta_1)+P(X|n,\theta_2)\theta_1 \\ &=\left[\textrm{Bin}(3,0.5)(1-\theta_1)+\textrm{Bin}(3,\theta_2)\theta_1\right] \end{align} \] for \(\theta_1,\theta_2\in[0,1]\).

Which in R code is simply:

#Define a handy "interval" function 
interval<-function(x,min,max) ifelse(x>=min & x<=max,x,0)
likelihood<-function(x,n,theta1,theta2) { 
  #Ensure that theta's are in 0,1
  theta1=interval(theta1,0,1)
  theta2=interval(theta2,0,1)
  #Calculate the value of the likelihood at this theta1,theta2
  val=dbinom(x=x,size=n,prob=0.5)*(1-theta1)+
      dbinom(x=x,size=n,prob=theta2)*theta1
  #Return the value
  return(val)
}

The Priors

We have no idea what the values of the biases on the coins are, nor about what fraction of coins in the bag are biased. Therefore, we ought to impose an uninformative uniform prior on all possible values of \(\theta_1,\theta_2\). \[ P(\theta_i)= \begin{cases} 1 & 0\leq \theta_i\leq 1\\ 0 & \textrm{otherwise} \end{cases} \]

prior<-function(theta1,theta2) { 
  #Calculate the value of the prior at theta1,theta2
  val=dunif(x=theta1,min=0,max=1)*
      dunif(x=theta2,min=0,max=1)
  #Return the value
  return(val)
}

Calculate the posterior distribution

This is where our sampler comes in. We know from Bayes theorem that the posterior is proportional to the likelihood times the prior, and we want to sample from this unknown distribution to produce points that we can use to perform inference about the properties of the posterior.

coins.MH<-function(n,start) { 
  #Initialise the variables
  X<-Y<-rep(NA,n)
  X[1]<-start[1]
  Y[1]<-start[2]
  #Loop over our iterations 
  Id<-diag(2)
  #Define the posterior evaluation
  log.posterior<-function(x,n,theta1,theta2) 
    return(log(prior(theta1,theta2))+sum(log(likelihood(x=x,n=n,theta1=theta1,theta2=theta2))))
  #Loop over the steps in the chain  
  for (i in 2:n) { 
    #Generate a proposal point 
    prop<-rmvnorm(1,c(X[i-1],Y[i-1]),0.05*Id)
    #Compute the acceptance probability pi(y)/pi(x)
    #Use logarithms to avoid rounding errors
    accept<-exp(log.posterior(x=data,n=3,theta1=prop[1],theta2=prop[2])-
                log.posterior(x=data,n=3,theta1=X[i-1] ,theta2=Y[i-1] ))
    if (!is.finite(accept)) { accept<-0 } 
    #Draw a random univariate 
    u<-runif(1)
    #Test the acceptance criteria
    if (u<accept) { 
      #Accept the new data point
      X[i]<-prop[1]
      Y[i]<-prop[2]
    } else { 
      X[i]<-X[i-1]
      Y[i]<-Y[i-1]
    }
  }
  return(cbind(X,Y))
}

We now run our chain to construct our posterior samples:

#Running our MH Code
set.seed(42)
data<-simulate.throws(1e3)
chain<-coins.MH(1e4,start=c(0.8,0.1))
magplot(chain,type='l',col=seqinr::col2alpha('blue3',0.3),lwd=2,
        xlab=expression(theta[1]),ylab=expression(theta[2]))

What are the three things that we’ve forgotten?

  1. Burn in removal
  2. Trimming
  3. Multiple starting points

So let’s do those. First:

#Building more samples 
set.seed(42)
library(foreach)
library(doParallel)
## Loading required package: iterators
## Loading required package: parallel
registerDoParallel(cores=4)
chain<-foreach(i=1:8,.combine='rbind',.inorder=TRUE)%dopar%{
  return(
    cbind(coins.MH(1e4,start=c(runif(1),runif(1))),i)
  )
}
magplot(chain[,1:2],type='n',pch=20,col=seqinr::col2alpha('blue3',0.3),lwd=2,
        xlab=expression(theta[1]),ylab=expression(theta[2]),asp=1)
colours<-RColorBrewer::brewer.pal(8,"Set2")
for (i in 1:max(chain[,3])) { 
  lines(chain[which(chain[,3]==i),1:2],
        col=seqinr::col2alpha(colours[i],0.3),lwd=2)
}

#Cleaning up our posterior samples: Burn in removal
layout(cbind(1,2))
magplot(chain[,1],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[1]))
magplot(chain[,2],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[2]))

chain<-chain[-1e2:0,]
magplot(chain[,1],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[1]))
magplot(chain[,2],type='l',col='blue3',lwd=2,xlab="Step Number",
        ylab=expression(theta[2]))

And now our thinning:

#Cleaning up our posterior samples: Thinning 
layout(matrix(1:4,2,2,byrow=TRUE))
par(mar=c(3,4,1,1))
chain.fifth<-thin(chain,by=5)
chain.twent<-thin(chain,by=20)
chain.hundr<-thin(chain,by=100)
magplot(chain[-1,1],chain[-nrow(chain),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+1"),xlab=expression(theta[1]^i),asp=1)
magplot(chain.fifth[-1,1],chain.fifth[-nrow(chain.fifth),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+5"),xlab=expression(theta[1]^i),asp=1)
magplot(chain.twent[-1,1],chain.twent[-nrow(chain.twent),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+20"),xlab=expression(theta[1]^i),asp=1)
magplot(chain.hundr[-1,1],chain.hundr[-nrow(chain.hundr),1],pch=20,col=colour,lwd=2,
        ylab=expression(theta[1]^"i+100"),xlab=expression(theta[1]^i),asp=1)

With our cleaned samples we can now do our posterior inference:

#Final posterior samples 
chain<-thin(chain,10)
magplot(chain[,-3],type='p',pch=20,col=seqinr::col2alpha('blue3',0.3),lwd=2,
        xlab=expression(theta[1]),ylab=expression(theta[2]),asp=1)

Recall that these posterior samples allow us to trivially compute the marginal distributions:

#Final posterior samples 
layout(cbind(1,2))
magplot(density(chain[,1],bw=0.05),col='blue3',lwd=2,
        xlab=expression(theta[1]),ylab="PDF")
magplot(density(chain[,2],bw=0.05),col='blue3',lwd=2,
        xlab=expression(theta[2]),ylab="PDF")