Bayesian inference for a logistic regression model (Part 6)

Part 6: Hamiltonian Monte Carlo (HMC)

Introduction

This is the sixth part in a series of posts on MCMC-based Bayesian inference for a logistic regression model. If you are new to this series, please go back to Part 1.

In the previous post we saw how to construct an MCMC algorithm utilising gradient information by considering a Langevin equation having our target distribution of interest as its equilibrium. This equation has a physical interpretation in terms of the stochastic dynamics of a particle in a potential equal to minus the log of the target density. It turns out that thinking about the deterministic dynamics of a particle in such a potential can lead to more efficient MCMC algorithms.

Hamiltonian dynamics

Hamiltonian dynamics is often presented as an extension of a fairly general version of Lagrangian dynamics. However, for our purposes a rather simple version is quite sufficient, based on basic concepts from Newtonian dynamics, familiar from school. Inspired by our Langevin example, we will consider the dynamics of a particle in a potential function V(q). We will see later why we want V(q) = -\log \pi(q) for our target of interest, \pi(\cdot). In the context of Hamiltonian (and Lagrangian) dynamics we typically use q as our position variable, rather than x.

The potential function induces a (conservative) force on the particle equal to -\nabla V(q) when the particle is at position q. Then Newton’s second law of motion, "F=ma", takes the form

\displaystyle \nabla V(q) + m \ddot{q} = 0.

In Newtonian mechanics, we often consider the position vector q as 3-dimensional. Here it will be n-dimensional, where n is the number of variables in our target. We can then think of our second law as governing a single n-dimensional particle of mass m, or n one-dimensional particles all of mass m. But in this latter case, there is no need to assume that all particles have the same mass, and we could instead write our law of motion as

\displaystyle \nabla V(q) + M \ddot{q} = 0,

where M is a diagonal matrix. But in fact, since we could change coordinates, there’s no reason to require that M is diagonal. All we need is that M is positive definite, so that we don’t have negative mass in any coordinate direction.

We will take the above equation as the fundamental law governing our dynamical system of interest. The motivation from Newtonian dynamics is interesting, but not required. What is important is that the dynamics of such a system are conservative, in a way that we will shortly make precise.

Our law of motion is a second-order differential equation, since it involves the second derivative of q wrt time. If you’ve ever studied differential equations, you’ll know that there is an easy way to turn a second order equation into a first order equation with twice the dimension by augmenting the system with the velocities. Here, it is more convenient to augment the system with "momentum" variables, p, which we define as p = M\dot{q}. Then we can write our second order system as a pair of first order equations

\displaystyle \dot{q} = M^{-1}p

\displaystyle \dot{p} = -\nabla V(q)

These are, in fact, Hamilton’s equations for this system, though this isn’t how they are typically written.

If we define the kinetic energy as

\displaystyle T(p) = \frac{1}{2}p^\text{T}M^{-1}p,

then the Hamiltonian

\displaystyle H(q,p) = V(q) + T(p),

representing the total energy in the system, is conserved, since

\displaystyle \dot{H} = \nabla V\cdot \dot{q} + \dot{p}^\text{T}M^{-1}p = \nabla V\cdot \dot{q} + \dot{p}^\text{T}\dot{q} = [\nabla V + \dot{p}]\cdot\dot{q} = 0.

So, if we obey our Hamiltonian dynamics, our trajectory in (q,p)-space will follow contours of the Hamiltonian. It’s also clear that the system is time-reversible, so flipping the sign of the momentum p and integrating will exactly reverse the direction in which the contours are traversed. Another quite important property of Hamiltonian dynamics is that they are volume preserving. This can be verified by checking that the divergence of the flow is zero.

\displaystyle \nabla\cdot(\dot{q},\dot{p}) = \nabla_q\cdot\dot{q} + \nabla_p\cdot\dot{p} = 0,

since \dot{q} is a function of p only and \dot{p} is a function of q only.

Hamiltonian Monte Carlo (HMC)

In Hamiltonian Monte Carlo we introduce an augmented target distribution,

\displaystyle \tilde \pi(q,p) \propto \exp[-H(q,p)]

It is clear from this definition that moves leaving the Hamiltonian invariant will also leave the augmented target density unchanged. By following the Hamiltonian dynamics, we will be able to make big (reversible) moves in the space that will be accepted with high probability. Also, our target factorises into two independent components as

\displaystyle \tilde \pi(q,p) \propto \exp[-V(q)]\exp[-T(p)],

and so choosing V(q)=-\log \pi(q) will ensure that the q-marginal is our real target of interest, \pi(\cdot). It’s also clear that our p-marginal is \mathcal N(0,M). This is also the full-conditional for p, so re-sampling p from this distribution and leaving q unchanged is a Gibbs move that will leave the augmented target invariant. Re-sampling p will be necessary to properly explore our augmented target, since this will move us to a different contour of H.

So, an idealised version of HMC would proceed as follows: First, update p by sampling from its known tractable marginal. Second, update p and q jointly by following the Hamiltonian dynamics. If this second move is regarded as a (deterministic) reversible M-H proposal, it will be accepted with probability one since it leaves the augmented target density unchanged. If we could exactly integrate Hamilton’s equations, this would be fine. But in practice, we will need to use some imperfect numerical method for the integration step. But just as for MALA, we can regard the numerical method as a M-H proposal and correct for the fact that it is imperfect, preserving the exact augmented target distribution.

Hamiltonian systems admit nice numerical integration schemes called symplectic integrators. In HMC a simple alternating Euler method is typically used, known as the leap-frog algorithm. The component updates are all shear transformations, and therefore volume preserving, and exact reversibility is ensured by starting and ending with a half-step update of the momentum variables. In principle, to ensure reversibility of the proposal the momentum variables should be sign-flipped (reversed) to finish, but in practice this doesn’t matter since it doesn’t affect the evaluation of the Hamiltonian and it will then get refreshed, anyway.

So, advancing our system by a time step \epsilon can be done with

\displaystyle p(t+\epsilon/2) := p(t) - \frac{\epsilon}{2}\nabla V(q(t))

\displaystyle q(t+\epsilon) := q(t) + \epsilon M^{-1}p(t+\epsilon/2)

\displaystyle p(t+\epsilon) := p(t+\epsilon/2) - \frac{\epsilon}{2}\nabla V(q(t+\epsilon))

It is clear that if many such updates are chained together, adjacent momentum updates can be collapsed together, giving rise to the "leap-frog" nature of the algorithm, and therefore requiring roughly one gradient evaluation per \epsilon update, rather than two. Since this integrator is volume preserving and exactly reversible, for reasonably small \epsilon it follows the Hamiltonian dynamics reasonably well, but not exactly, and so it does not exactly preserve the Hamiltonian. However, it does make a good M-H proposal, and reasonable acceptance probabilities can often be obtained by chaining together l updates to advance the time of the system by T=l\epsilon. The "optimal" value of l and \epsilon will be highly problem dependent, but values of l=20 or l=50 are not unusual. There are various more-or-less standard methods for tuning these, but we will not consider them here.

Note that since our HMC update on the augmented space consists of a Gibbs move and a M-H update, it is important that our M-H kernel does not keep or thread through the old log target density from the previous M-H update, since the Gibbs move will have changed it in the meantime.

Implementations

R

We need a M-H kernel that does not thread through the old log density.

mhKernel = function(logPost, rprop)
    function(x) {
        prop = rprop(x)
        a = logPost(prop) - logPost(x)
        if (log(runif(1)) < a)
            prop
        else
            x
    }

We can then use this to construct a M-H move as part of our HMC update.

hmcKernel = function(lpi, glpi, eps = 1e-4, l=10, dmm = 1) {
    sdmm = sqrt(dmm)
    leapf = function(q, p) {
        p = p + 0.5*eps*glpi(q)
        for (i in 1:l) {
            q = q + eps*p/dmm
            if (i < l)
                p = p + eps*glpi(q)
            else
                p = p + 0.5*eps*glpi(q)
        }
        list(q=q, p=-p)
    }
    alpi = function(x)
        lpi(x$q) - 0.5*sum((x$p^2)/dmm)
    rprop = function(x)
        leapf(x$q, x$p)
    mhk = mhKernel(alpi, rprop)
    function(q) {
        d = length(q)
        x = list(q=q, p=rnorm(d, 0, sdmm))
        mhk(x)$q
    }
}

See the full runnable script for further details.

Python

First a M-H kernel,

def mhKernel(lpost, rprop):
    def kernel(x):
        prop = rprop(x)
        a = lpost(prop) - lpost(x)
        if (np.log(np.random.rand()) < a):
            x = prop
        return x
    return kernel

and then an HMC kernel.

def hmcKernel(lpi, glpi, eps = 1e-4, l=10, dmm = 1):
    sdmm = np.sqrt(dmm)
    def leapf(q, p):    
        p = p + 0.5*eps*glpi(q)
        for i in range(l):
            q = q + eps*p/dmm
            if (i < l-1):
                p = p + eps*glpi(q)
            else:
                p = p + 0.5*eps*glpi(q)
        return (q, -p)
    def alpi(x):
        (q, p) = x
        return lpi(q) - 0.5*np.sum((p**2)/dmm)
    def rprop(x):
        (q, p) = x
        return leapf(q, p)
    mhk = mhKernel(alpi, rprop)
    def kern(q):
        d = len(q)
        p = np.random.randn(d)*sdmm
        return mhk((q, p))[0]
    return kern

See the full runnable script for further details.

JAX

Again, we want an appropriate M-H kernel,

def mhKernel(lpost, rprop, dprop = jit(lambda new, old: 1.)):
    @jit
    def kernel(key, x):
        key0, key1 = jax.random.split(key)
        prop = rprop(key0, x)
        ll = lpost(x)
        lp = lpost(prop)
        a = lp - ll + dprop(x, prop) - dprop(prop, x)
        accept = (jnp.log(jax.random.uniform(key1)) < a)
        return jnp.where(accept, prop, x)
    return kernel

and then an HMC kernel.

def hmcKernel(lpi, glpi, eps = 1e-4, l = 10, dmm = 1):
    sdmm = jnp.sqrt(dmm)
    @jit
    def leapf(q, p):    
        p = p + 0.5*eps*glpi(q)
        for i in range(l):
            q = q + eps*p/dmm
            if (i < l-1):
                p = p + eps*glpi(q)
            else:
                p = p + 0.5*eps*glpi(q)
        return jnp.concatenate((q, -p))
    @jit
    def alpi(x):
        d = len(x) // 2
        return lpi(x[jnp.array(range(d))]) - 0.5*jnp.sum((x[jnp.array(range(d,2*d))]**2)/dmm)
    @jit
    def rprop(k, x):
        d = len(x) // 2
        return leapf(x[jnp.array(range(d))], x[jnp.array(range(d, 2*d))])
    mhk = mhKernel(alpi, rprop)
    @jit
    def kern(k, q):
        key0, key1 = jax.random.split(k)
        d = len(q)
        x = jnp.concatenate((q, jax.random.normal(key0, [d])*sdmm))
        return mhk(key1, x)[jnp.array(range(d))]
    return kern

There is something a little bit strange about this implementation, since the proposal for the M-H move is deterministic, the function rprop just ignores the RNG key that is passed to it. We could tidy this up by making a M-H function especially for deterministic proposals. We won’t pursue this here, but this issue will crop up again later in some of the other functional languages.

See the full runnable script for further details.

Scala

A M-H kernel,

def mhKern[S](
    logPost: S => Double, rprop: S => S,
    dprop: (S, S) => Double = (n: S, o: S) => 1.0
  ): (S) => S =
    val r = Uniform(0.0,1.0)
    x0 =>
      val x = rprop(x0)
      val ll0 = logPost(x0)
      val ll = logPost(x)
      val a = ll - ll0 + dprop(x0, x) - dprop(x, x0)
      if (math.log(r.draw()) < a) x else x0

and a HMC kernel.

def hmcKernel(lpi: DVD => Double, glpi: DVD => DVD, dmm: DVD,
  eps: Double = 1e-4, l: Int = 10) =
  val sdmm = sqrt(dmm)
  def leapf(q: DVD, p: DVD): (DVD, DVD) = 
    @tailrec def go(q0: DVD, p0: DVD, l: Int): (DVD, DVD) =
      val q = q0 + eps*(p0/:/dmm)
      val p = if (l > 1)
        p0 + eps*glpi(q)
      else
        p0 + 0.5*eps*glpi(q)
      if (l == 1)
        (q, -p)
      else
        go(q, p, l-1)
    go(q, p + 0.5*eps*glpi(q), l)
  def alpi(x: (DVD, DVD)): Double =
    val (q, p) = x
    lpi(q) - 0.5*sum(pow(p,2) /:/ dmm)
  def rprop(x: (DVD, DVD)): (DVD, DVD) =
    val (q, p) = x
    leapf(q, p)
  val mhk = mhKern(alpi, rprop)
  (q: DVD) =>
    val d = q.length
    val p = sdmm map (sd => Gaussian(0,sd).draw())
    mhk((q, p))._1

See the full runnable script for further details.

Haskell

A M-H kernel:

mdKernel :: (StatefulGen g m) => (s -> Double) -> (s -> s) -> g -> s -> m s
mdKernel logPost prop g x0 = do
  let x = prop x0
  let ll0 = logPost x0
  let ll = logPost x
  let a = ll - ll0
  u <- (genContVar (uniformDistr 0.0 1.0)) g
  let next = if ((log u) < a)
        then x
        else x0
  return next

Note that here we are using a M-H kernel specifically for deterministic proposals, since there is no non-determinism signalled in the type signature of prop. We can then use this to construct our HMC kernel.

hmcKernel :: (StatefulGen g m) =>
  (Vector Double -> Double) -> (Vector Double -> Vector Double) -> Vector Double ->
  Double -> Int -> g ->
  Vector Double -> m (Vector Double)
hmcKernel lpi glpi dmm eps l g = let
  sdmm = cmap sqrt dmm
  leapf q p = let
    go q0 p0 l = let
      q = q0 + (scalar eps)*p0/dmm
      p = if (l > 1)
        then p0 + (scalar eps)*(glpi q)
        else p0 + (scalar (eps/2))*(glpi q)
      in if (l == 1)
      then (q, -p)
      else go q p (l - 1)
    in go q (p + (scalar (eps/2))*(glpi q)) l
  alpi x = let
    (q, p) = x
    in (lpi q) - 0.5*(sumElements (p*p/dmm))
  prop x = let
    (q, p) = x
    in leapf q p
  mk = mdKernel alpi prop g
  in (\q0 -> do
         let d = size q0
         zl <- (replicateM d . genContVar (normalDistr 0.0 1.0)) g
         let z = fromList zl
         let p0 = sdmm * z
         (q, p) <- mk (q0, p0)
         return q)

See the full runnable script for further details.

Dex

Again we can use a M-H kernel specific to deterministic proposals.

def mdKernel {s} (lpost: s -> Float) (prop: s -> s)
    (x0: s) (k: Key) : s =
  x = prop x0
  ll0 = lpost x0
  ll = lpost x
  a = ll - ll0
  u = rand k
  select (log u < a) x x0

and use this to construct an HMC kernel.

def hmcKernel {n} (lpi: (Fin n)=>Float -> Float)
    (dmm: (Fin n)=>Float) (eps: Float) (l: Nat)
    (q0: (Fin n)=>Float) (k: Key) : (Fin n)=>Float =
  sdmm = sqrt dmm
  idmm = map (\x. 1.0/x) dmm
  glpi = grad lpi
  def leapf (q0: (Fin n)=>Float) (p0: (Fin n)=>Float) :
      ((Fin n)=>Float & (Fin n)=>Float) =
    p1 = p0 + (eps/2) .* (glpi q0)
    q1 = q0 + eps .* (p1*idmm)
    (q, p) = apply_n l (q1, p1) \(qo, po).
      pn = po + eps .* (glpi qo)
      qn = qo + eps .* (pn*idmm)
      (qn, pn)
    pf = p + (eps/2) .* (glpi q)
    (q, -pf)
  def alpi (qp: ((Fin n)=>Float & (Fin n)=>Float)) : Float =
    (q, p) = qp
    (lpi q) - 0.5*(sum (p*p*idmm))
  def prop (qp: ((Fin n)=>Float & (Fin n)=>Float)) :
      ((Fin n)=>Float & (Fin n)=>Float) =
    (q, p) = qp
    leapf q p
  mk = mdKernel alpi prop
  [k1, k2] = split_key k
  z = randn_vec k1
  p0 = sdmm * z
  (q, p) = mk (q0, p0) k2
  q

Note that the gradient is obtained via automatic differentiation. See the full runnable script for details.

Next steps

This was the main place that I was trying to get to when I started this series of posts. For differentiable log-posteriors (as we have in the case of Bayesian logistic regression), HMC is a pretty good algorithm for reasonably efficient posterior exploration. But there are lots of places we could go from here. We could explore the tuning of MCMC algorithms, or HMC extensions such as NUTS. We could look at MCMC algorithms that are specifically tailored to the logistic regression problem, or we could look at new MCMC algorithms for differentiable targets based on piecewise deterministic Markov processes. Alternatively, we could temporarily abandon MCMC and look at SMC or ABC approaches. Another possibility would be to abandon this multi-language approach and have a bit of a deep dive into Dex, which I think has the potential to be a great programming language for statistical computing. All of these are possibilities for the future, but I’ve a busy few weeks coming up, so the frequency of these posts is likely to substantially decrease.

Remember that all of the code associated with this series of posts is available from this github repo.

Bayesian inference for a logistic regression model (Part 5)

Part 5: the Metropolis-adjusted Langevin algorithm (MALA)

Introduction

This is the fifth part in a series of posts on MCMC-based Bayesian inference for a logistic regression model. If you are new to this series, please go back to Part 1.

In the previous post we saw how to use Langevin dynamics to construct an approximate MCMC scheme using the gradient of the log target distribution. Each step of the algorithm involved simulating from the Euler-Maruyama approximation to the transition kernel of the process, based on some pre-specified step size, \Delta t. We can improve the accuracy of this approximation by making the step size smaller, but this will come at the expense of a more slowly mixing MCMC chain.

Fortunately, there is an easy way to make the algorithm "exact" (in the sense that the equilibrium distribution of the Markov chain will be the exact target distribution), for any finite step size, \Delta t, simply by using the Euler-Maruyama approximation as the proposal distribution in a Metropolis-Hastings algorithm. This is the Metropolis-adjusted Langevin algorithm (MALA). There are various ways this could be coded up, but here, for clarity, a HoF for generating a MALA kernel will be used, and this function will in turn call on a HoF for generating a Metropolis-Hastings kernel.

Implementations

R

First we need a function to generate a M-H kernel.

mhKernel = function(logPost, rprop, dprop = function(new, old, ...) { 1 })
    function(x, ll) {
        prop = rprop(x)
        llprop = logPost(prop)
        a = llprop - ll + dprop(x, prop) - dprop(prop, x)
        if (log(runif(1)) < a)
            list(x=prop, ll=llprop)
        else
            list(x=x, ll=ll)
    }

Then we can easily write a function for returning a MALA kernel that makes use of this M-H function.

malaKernel = function(lpi, glpi, dt = 1e-4, pre = 1) {
    sdt = sqrt(dt)
    spre = sqrt(pre)
    advance = function(x) x + 0.5*pre*glpi(x)*dt
    mhKernel(lpi, function(x) rnorm(p, advance(x), spre*sdt),
             function(new, old) sum(dnorm(new, advance(old), spre*sdt, log=TRUE)))
}

Notice that our MALA function requires as input both the gradient of the log posterior (for the proposal) and the log posterior itself (for the M-H correction). Other details are as we have already seen – see the full runnable script.

Python

Again, we need a M-H kernel

def mhKernel(lpost, rprop, dprop = lambda new, old: 1.):
    def kernel(x, ll):
        prop = rprop(x)
        lp = lpost(prop)
        a = lp - ll + dprop(x, prop) - dprop(prop, x)
        if (np.log(np.random.rand()) < a):
            x = prop
            ll = lp
        return x, ll
    return kernel

and then a MALA kernel

def malaKernel(lpi, glpi, dt = 1e-4, pre = 1):
    p = len(init)
    sdt = np.sqrt(dt)
    spre = np.sqrt(pre)
    advance = lambda x: x + 0.5*pre*glpi(x)*dt
    return mhKernel(lpi, lambda x: advance(x) + np.random.randn(p)*spre*sdt,
            lambda new, old: np.sum(sp.stats.norm.logpdf(new, loc=advance(old), scale=spre*sdt)))

See the full runnable script for further details.

JAX

If we want our algorithm to run fast, and if we want to exploit automatic differentiation to avoid the need to manually compute gradients, then we can easily convert the above code to use JAX.

def mhKernel(lpost, rprop, dprop = jit(lambda new, old: 1.)):
    @jit
    def kernel(key, x, ll):
        key0, key1 = jax.random.split(key)
        prop = rprop(key0, x)
        lp = lpost(prop)
        a = lp - ll + dprop(x, prop) - dprop(prop, x)
        accept = (jnp.log(jax.random.uniform(key1)) < a)
        return jnp.where(accept, prop, x), jnp.where(accept, lp, ll)
    return kernel

def malaKernel(lpi, dt = 1e-4, pre = 1):
    p = len(init)
    glpi = jit(grad(lpost))
    sdt = jnp.sqrt(dt)
    spre = jnp.sqrt(pre)
    advance = jit(lambda x: x + 0.5*pre*glpi(x)*dt)
    return mhKernel(lpi, jit(lambda k, x: advance(x) +
                             jax.random.normal(k, [p])*spre*sdt),
            jit(lambda new, old:
                jnp.sum(jsp.stats.norm.logpdf(new,
                      loc=advance(old), scale=spre*sdt))))

See the full runnable script for further details.

Scala

def mhKernel[S](
    logPost: S => Double, rprop: S => S,
    dprop: (S, S) => Double = (n: S, o: S) => 1.0
  ): ((S, Double)) => (S, Double) =
    val r = Uniform(0.0,1.0)
    state =>
      val (x0, ll0) = state
      val x = rprop(x0)
      val ll = logPost(x)
      val a = ll - ll0 + dprop(x0, x) - dprop(x, x0)
      if (math.log(r.draw()) < a)
        (x, ll)
      else
        (x0, ll0)

def malaKernel(lpi: DVD => Double, glpi: DVD => DVD, pre: DVD, dt: Double = 1e-4) =
  val sdt = math.sqrt(dt)
  val spre = sqrt(pre)
  val p = pre.length
  def advance(beta: DVD): DVD =
    beta + (0.5*dt)*(pre*:*glpi(beta))
  def rprop(beta: DVD): DVD =
    advance(beta) + sdt*spre.map(Gaussian(0,_).sample())
  def dprop(n: DVD, o: DVD): Double = 
    val ao = advance(o)
    (0 until p).map(i => Gaussian(ao(i), spre(i)*sdt).logPdf(n(i))).sum
  mhKernel(lpi, rprop, dprop)

See the full runnable script for further details.

Haskell

mhKernel :: (StatefulGen g m) => (s -> Double) -> (s -> g -> m s) ->
  (s -> s -> Double) -> g -> (s, Double) -> m (s, Double)
mhKernel logPost rprop dprop g (x0, ll0) = do
  x <- rprop x0 g
  let ll = logPost(x)
  let a = ll - ll0 + (dprop x0 x) - (dprop x x0)
  u <- (genContVar (uniformDistr 0.0 1.0)) g
  let next = if ((log u) < a)
        then (x, ll)
        else (x0, ll0)
  return next

malaKernel :: (StatefulGen g m) =>
  (Vector Double -> Double) -> (Vector Double -> Vector Double) -> 
  Vector Double -> Double -> g ->
  (Vector Double, Double) -> m (Vector Double, Double)
malaKernel lpi glpi pre dt g = let
  sdt = sqrt dt
  spre = cmap sqrt pre
  p = size pre
  advance beta = beta + (scalar (0.5*dt))*pre*(glpi beta)
  rprop beta g = do
    zl <- (replicateM p . genContVar (normalDistr 0.0 1.0)) g
    let z = fromList zl
    return $ advance(beta) + (scalar sdt)*spre*z
  dprop n o = let
    ao = advance o
    in sum $ (\i -> logDensity (normalDistr (ao!i) 
      ((spre!i)*sdt)) (n!i)) <$> [0..(p-1)]
  in mhKernel lpi rprop dprop g

See the full runnable script for further details.

Dex

Recall that Dex is differentiable, so we don’t need to provide gradients.

def mhKernel {s} (lpost: s -> Float) (rprop: s -> Key -> s) (dprop: s -> s -> Float)
    (sll: (s & Float)) (k: Key) : (s & Float) =
  (x0, ll0) = sll
  [k1, k2] = split_key k
  x = rprop x0 k1
  ll = lpost x
  a = ll - ll0 + (dprop x0 x) - (dprop x x0)
  u = rand k2
  select (log u < a) (x, ll) (x0, ll0)

def malaKernel {n} (lpi: (Fin n)=>Float -> Float)
    (pre: (Fin n)=>Float) (dt: Float) :
    ((Fin n)=>Float & Float) -> Key -> ((Fin n)=>Float & Float) =
  sdt = sqrt dt
  spre = sqrt pre
  glp = grad lpi
  v = dt .* pre
  vinv = map (\ x. 1.0/x) v
  def advance (beta: (Fin n)=>Float) : (Fin n)=>Float =
    beta + (0.5*dt) .* (pre*(glp beta))
  def rprop (beta: (Fin n)=>Float) (k: Key) : (Fin n)=>Float =
    (advance beta) + sdt .* (spre*(randn_vec k))
  def dprop (new: (Fin n)=>Float) (old: (Fin n)=>Float) : Float =
    ao = advance old
    diff = new - ao
    -0.5 * sum ((log v) + diff*diff*vinv)
  mhKernel lpi rprop dprop

See the full runnable script for further details.

Next steps

MALA gives us an MCMC algorithm that exploits gradient information to generate "informed" M-H proposals. But it still has a rather "diffusive" character, making it difficult to tune in such a way that large moves are likely to be accepted in challenging high-dimensional situations.

The Langevin dynamics on which MALA is based can be interpreted as the (over-damped) stochastic dynamics of a particle moving in a potential energy field corresponding to minus the log posterior. It turns out that the corresponding deterministic dynamics can be exploited to generate proposals better able to make large moves while still having a high probability of acceptance. This is the idea behind Hamiltonian Monte Carlo (HMC), which we’ll look at next.

Bayesian inference for a logistic regression model (Part 4)

Part 4: Gradients and the Langevin algorithm

Introduction

This is the fourth part in a series of posts on MCMC-based Bayesian inference for a logistic regression model. If you are new to this series, please go back to Part 1.

In the previous post we saw how the Metropolis algorithm could be used to generate a Markov chain targeting our posterior distribution. In high dimensions the diffusive nature of the Metropolis random walk proposal becomes increasingly inefficient. It is therefore natural to try and develop algorithms that use additional information about the target distribution. In the case of a differentiable log posterior target, a natural first step in this direction is to try and make use of gradient information.

Gradient of a logistic regression model

There are various ways to derive the gradient of our logistic regression model, but it might be simplest to start from the first form of the log likelihood that we deduced in Part 2:

\displaystyle l(\beta;y) = y^\textsf{T}X\beta - \mathbf{1}^\textsf{T}\log(\mathbf{1}+\exp[X\beta])

We can write this out in component form as

\displaystyle l(\beta;y) = \sum_j\sum_j y_iX_{ij}\beta_j - \sum_i\log\left(1+\exp\left[\sum_jX_{ij}\beta_j\right]\right).

Differentiating wrt \beta_k gives

\displaystyle \frac{\partial l}{\partial \beta_k} = \sum_i y_iX_{ik} - \sum_i \frac{\exp\left[\sum_j X_{ij}\beta_j\right]X_{ik}}{1+\exp\left[\sum_j X_{ij}\beta_j\right]}.

It’s then reasonably clear that stitching all of the partial derivatives together will give the gradient vector

\displaystyle \nabla l = X^\textsf{T}\left[ y - \frac{\mathbf{1}}{\mathbf{1}+\exp[-X\beta]} \right].

This is the gradient of the log likelihood, but we also need the gradient of the log prior. Since we are assuming independent \beta_i \sim N(0,v_i) priors, it is easy to see that the gradient of the log prior is just -\beta\circ v^{-1}. It is the sum of these two terms that gives the gradient of the log posterior.

R

In R we can implement our gradient function as

glp = function(beta) {
    glpr = -beta/(pscale*pscale)
    gll = as.vector(t(X) %*% (y - 1/(1 + exp(-X %*% beta))))
    glpr + gll
}

Python

In Python we could use

def glp(beta):
    glpr = -beta/(pscale*pscale)
    gll = (X.T).dot(y - 1/(1 + np.exp(-X.dot(beta))))
    return (glpr + gll)

We don’t really need a JAX version, since JAX can auto-diff the log posterior for us.

Scala

  def glp(beta: DVD): DVD =
    val glpr = -beta /:/ pvar
    val gll = (X.t)*(y - ones/:/(ones + exp(-X*beta)))
    glpr + gll

Haskell

Using hmatrix we could use something like

glp :: Matrix Double -> Vector Double -> Vector Double -> Vector Double
glp x y b = let
  glpr = -b / (fromList [100.0, 1, 1, 1, 1, 1, 1, 1])
  gll = (tr x) #> (y - (scalar 1)/((scalar 1) + (cmap exp (-x #> b))))
  in glpr + gll

There’s something interesting to say about Haskell and auto-diff, but getting into this now will be too much of a distraction. I may come back to it in some future post.

Dex

Dex is differentiable, so we don’t need a gradient function – we can just use grad lpost. However, for interest and comparison purposes we could nevertheless implement it directly with something like

prscale = map (\ x. 1.0/x) pscale

def glp (b: (Fin 8)=>Float) : (Fin 8)=>Float =
  glpr = -b*prscale*prscale
  gll = (transpose x) **. (y - (map (\eta. 1.0/(1.0 + eta)) (exp (-x **. b))))
  glpr + gll

Langevin diffusions

Now that we have a way of computing the gradient of the log of our target density we need some MCMC algorithms that can make good use of it. In this post we will look at a simple approximate MCMC algorithm derived from an overdamped Langevin diffusion model. In subsequent posts we’ll look at more sophisticated, exact MCMC algorithms.

The multivariate stochastic differential equation (SDE)

\displaystyle dX_t = \frac{1}{2}\nabla\log\pi(X_t)dt + dW_t

has \pi(\cdot) as its equilibrium distribution. Informally, an SDE of this form is a continuous time process with infinitesimal transition kernel

\displaystyle X_{t+dt}|(X_t=x_t) \sim N\left(x_t+\frac{1}{2}\nabla\log\pi(x_t)dt,\mathbf{I}dt\right).

There are various more-or-less formal ways to see that \pi(\cdot) is stationary. A good way is to check it satisfies the Fokker–Planck equation with zero LHS. A less formal approach would be to see that the infinitesimal transition kernel for the process satisfies detailed balance with \pi(\cdot).

Similar arguments show that for any fixed positive definite matrix A, the SDE

\displaystyle dX_t = \frac{1}{2}A\nabla\log\pi(X_t)dt + A^{1/2}dW_t

also has \pi(\cdot) as a stationary distribution. It is quite common to choose a diagonal matrix A to put the components of X_t on a common scale.

The unadjusted Langevin algorithm

Simulating exact sample paths from SDEs such as the overdamped Langevin diffusion model is typically difficult (though not necessarily impossible), so we instead want something simple and tractable as the basis of our MCMC algorithms. Here we will just simulate from the Euler–Maruyama approximation of the process by choosing a small but finite time step \Delta t and using the transition kernel

\displaystyle X_{t+\Delta t}|(X_t=x_t) \sim N\left(x_t+\frac{1}{2}A\nabla\log\pi(x_t)\Delta t, A\Delta t\right)

as the basis of our MCMC method. For sufficiently small \Delta t this should accurately approximate the Langevin dynamics, leading to an equilibrium distribution very close to \pi(\cdot). That said, we would like to choose \Delta t as large as we can get away with, since that will lead to a more rapidly mixing MCMC chain. Below are some implementations of this kernel for a diagonal pre-conditioning matrix.

Implementation

R

We can create a kernel for the unadjusted Langevin algorithm in R with the following function.

ulKernel = function(glpi, dt = 1e-4, pre = 1) {
    sdt = sqrt(dt)
    spre = sqrt(pre)
    advance = function(x) x + 0.5*pre*glpi(x)*dt
    function(x, ll) rnorm(p, advance(x), spre*sdt)
}

Here, we can pass in pre, which is expected to be a vector representing the diagonal of the pre-conditioning matrix, A. We can then use this kernel to generate an MCMC chain as we have seen previously. See the full runnable script for further details.

Python

def ulKernel(glpi, dt = 1e-4, pre = 1):
    p = len(init)
    sdt = np.sqrt(dt)
    spre = np.sqrt(pre)
    advance = lambda x: x + 0.5*pre*glpi(x)*dt
    def kernel(x):
        return advance(x) + np.random.randn(p)*spre*sdt
    return kernel

See the full runnable script for further details.

JAX

def ulKernel(lpi, dt = 1e-4, pre = 1):
    p = len(init)
    glpi = jit(grad(lpi))
    sdt = jnp.sqrt(dt)
    spre = jnp.sqrt(pre)
    advance = jit(lambda x: x + 0.5*pre*glpi(x)*dt)
    @jit
    def kernel(key, x):
        return advance(x) + jax.random.normal(key, [p])*spre*sdt
    return kernel

Note how for JAX we can just pass in the log posterior, and the gradient function can be obtained by automatic differentiation. See the full runnable script for further details.

Scala

def ulKernel(glp: DVD => DVD, pre: DVD, dt: Double): DVD => DVD =
  val sdt = math.sqrt(dt)
  val spre = sqrt(pre)
  def advance(beta: DVD): DVD =
    beta + (0.5*dt)*(pre*:*glp(beta))
  beta => advance(beta) + sdt*spre.map(Gaussian(0,_).sample())

See the full runnable script for further details.

Haskell

ulKernel :: (StatefulGen g m) =>
  (Vector Double -> Vector Double) -> Vector Double -> Double -> g ->
  Vector Double -> m (Vector Double)
ulKernel glpi pre dt g beta = do
  let sdt = sqrt dt
  let spre = cmap sqrt pre
  let p = size pre
  let advance beta = beta + (scalar (0.5*dt))*pre*(glpi beta)
  zl <- (replicateM p . genContVar (normalDistr 0.0 1.0)) g
  let z = fromList zl
  return $latex  advance(beta) + (scalar sdt)*spre*z

See the full runnable script for further details.

Dex

In Dex we can write a function that accepts a gradient function

def ulKernel {n} (glpi: (Fin n)=>Float -> (Fin n)=>Float)
    (pre: (Fin n)=>Float) (dt: Float)
    (b: (Fin n)=>Float) (k: Key) : (Fin n)=>Float =
  sdt = sqrt dt
  spre = sqrt pre
  b + (((0.5)*dt) .* (pre*(glpi b))) +
    (sdt .* (spre*(randn_vec k)))

or we can write a function that accepts a log posterior, and uses auto-diff to construct the gradient

def ulKernel {n} (lpi: (Fin n)=>Float -> Float)
    (pre: (Fin n)=>Float) (dt: Float)
    (b: (Fin n)=>Float) (k: Key) : (Fin n)=>Float =
  glpi = grad lpi
  sdt = sqrt dt
  spre = sqrt pre
  b + ((0.5)*dt) .* (pre*(glpi b)) +
    sdt .* (spre*(randn_vec k))

and since Dex is statically typed, we can’t easily mix these functions up.

See the full runnable scripts, without and with auto-diff.

Next steps

In this post we have seen how to construct an MCMC algorithm that makes use of gradient information. But this algorithm is approximate. In the next post we’ll see how to correct for the approximation by using the Langevin updates as proposals within a Metropolis-Hastings algorithm.

Bayesian inference for a logistic regression model (Part 3)

Part 3: The Metropolis algorithm

Introduction

This is the third part in a series of posts on MCMC-based Bayesian inference for a logistic regression model. If you are new to this series, please go back to Part 1.

In the previous post we derived the log posterior for the model and implemented it in a variety of programming languages and libraries. In this post we will construct a Markov chain having the posterior as its equilibrium.

MCMC

Detailed balance

A homogeneous Markov chain with transition kernel p(\theta_{n+1}|\theta_n) is said to satisfy detailed balance for some target distribution \pi(\theta) if

\displaystyle \pi(\theta)p(\theta'|\theta) = \pi(\theta')p(\theta|\theta'), \quad \forall \theta, \theta'

Integrating both sides wrt \theta gives

\displaystyle \int_\Theta \pi(\theta)p(\theta'|\theta)d\theta = \pi(\theta'),

from which it is clear that \pi(\cdot) is a stationary distribution of the chain (and the chain is reversible). Under fairly mild regularity conditions we expect \pi(\cdot) to be the equilibrium distribution of the chain.

For a given target \pi(\cdot) we would like to find an easy-to-sample-from transition kernel p(\cdot|\cdot) that satisfies detailed balance. This will then give us a way to (asymptotically) generate samples from our target.

In the context of Bayesian inference, the target \pi(\theta) will typically be the posterior distribution, which in the previous post we wrote as \pi(\theta|y). Here we drop the notational dependence on y, since MCMC can be used for any target distribution of interest.

Metropolis-Hastings

Suppose we have a fairly arbitrary easy-to-sample-from transition kernel q(\theta_{n+1}|\theta_n) and a target of interest, \pi(\cdot). Metropolis-Hastings (M-H) is a strategy for using q(\cdot|\cdot) to construct a new transition kernel p(\cdot|\cdot) satisfying detailed balance for \pi(\cdot).

The kernel p(\theta'|\theta) can be described algorithmically as follows:

  1. Call the current state of the chain \theta. Generate a proposal \theta^\star by simulating from q(\theta^\star|\theta).
  2. Compute the acceptance probability

\displaystyle \alpha(\theta^\star|\theta) = \min\left[1,\frac{\pi(\theta^\star)q(\theta|\theta^\star)}{\pi(\theta)q(\theta^\star|\theta)}\right].

  1. With probability \alpha(\theta^\star|\theta) return new state \theta'=\theta^\star, otherwise return \theta'=\theta.

It is clear from the algorithmic description that this kernel will have a point mass at \theta'=\theta, but that for \theta'\not=\theta the transition kernel will be p(\theta'|\theta)=q(\theta'|\theta)\alpha(\theta'|\theta). But then

\displaystyle \pi(\theta)p(\theta'|\theta) = \min[\pi(\theta)q(\theta'|\theta),\pi(\theta')q(\theta|\theta')]

is symmetric in \theta and \theta', and so detailed balance is satisfied. Since detailed balance is trivial at the point mass at \theta=\theta' we are done.

Metropolis algorithm

It is often convenient to generate proposals perturbatively, using a distribution that is symmetric about the current state of the chain. But then q(\theta'|\theta)=q(\theta|\theta'), and so q(\cdot|\cdot) drops out of the acceptance probability. This is the Metropolis algorithm.

Some computational tricks

To generate an event with probability \alpha(\theta^\star|\theta), we can generate a u\sim U(0,1) and accept if u < \alpha(\theta^\star|\theta). This is convenient for several reasons. First, it means that we can ignore the "min", and just accept if

\displaystyle u < \frac{\pi(\theta^\star)q(\theta|\theta^\star)}{\pi(\theta)q(\theta^\star|\theta)}

since u\leq 1 regardless. Better still, we can take logs, and accept if

\displaystyle \log u < \log\pi(\theta^\star) - \log\pi(\theta) + \log q(\theta|\theta^\star) - \log q(\theta^\star|\theta),

so there is no need to evaluate any raw densities. Again, in the case of a symmetric proposal distribution, the q(\cdot|\cdot) terms can be dropped.

Another trick worth noting is that in the case of the simple M-H algorithm described, using a single update for the entire state space (and not multiple component-wise updates, for example), and assuming that the same M-H kernel is used repeatedly to generate successive states of a Markov chain, then the \log \pi(\theta) term (which in the context of Bayesian inference will typically be the log posterior) will have been computed at the previous update (irrespective of whether or not the previous move was accepted). So if we are careful about how we pass on that old value to the next iteration, we can avoid recomputing the log posterior, and our algorithm will only require one log posterior evaluation per iteration rather than two. In functional programming languages it is often convenient to pass around this current log posterior density evaluation explicitly, effectively augmenting the state space of the Markov chain to include the log posterior density.

HoF for a M-H kernel

Since I’m a fan of functional programming, we will adopt a functional style throughout, and start by creating a higher-order function (HoF) that accepts a log-posterior and proposal kernel as input and returns a Metropolis kernel as output.

R

In R we can write a function to create a M-H kernel as follows.

mhKernel = function(logPost, rprop, dprop = function(new, old, ...) { 1 })
    function(x, ll) {
        prop = rprop(x)
        llprop = logPost(prop)
        a = llprop - ll + dprop(x, prop) - dprop(prop, x)
        if (log(runif(1)) < a)
            list(x=prop, ll=llprop)
        else
            list(x=x, ll=ll)
    }

Note that the kernel returned requires as input both a current state x and its associated log-posterior, ll. The new state and log-posterior densities are returned.

We need to use this transition kernel to simulate a Markov chain by successive substitution of newly simulated values back into the kernel. In more sophisticated programming languages we will use streams for this, but in R we can just use a for loop to sample values and write the states into the rows of a matrix.

mcmc = function(init, kernel, iters = 10000, thin = 10, verb = TRUE) {
    p = length(init)
    ll = -Inf
    mat = matrix(0, nrow = iters, ncol = p)
    colnames(mat) = names(init)
    x = init
    if (verb) 
        message(paste(iters, "iterations"))
    for (i in 1:iters) {
        if (verb) 
            message(paste(i, ""), appendLF = FALSE)
        for (j in 1:thin) {
            pair = kernel(x, ll)
            x = pair$x
            ll = pair$ll
            }
        mat[i, ] = x
        }
    if (verb) 
        message("Done.")
    mat
}

Then, in the context of our running logistic regression example, and using the log-posterior from the previous post, we can construct our kernel and run it as follows.

pre = c(10.0,1,1,1,1,1,5,1)
out = mcmc(init, mhKernel(lpost,
          function(x) x + pre*rnorm(p, 0, 0.02)), thin=1000)

Note the use of a symmetric proposal, so the proposal density is not required. Also note the use of a larger proposal variance for the intercept term and the second last covariate. See the full runnable script for further details.

Python

We can do something very similar to R in Python using NumPy. Our HoF for constructing a M-H kernel is

def mhKernel(lpost, rprop, dprop = lambda new, old: 1.):
    def kernel(x, ll):
        prop = rprop(x)
        lp = lpost(prop)
        a = lp - ll + dprop(x, prop) - dprop(prop, x)
        if (np.log(np.random.rand()) < a):
            x = prop
            ll = lp
        return x, ll
    return kernel

Our Markov chain runner function is

def mcmc(init, kernel, thin = 10, iters = 10000, verb = True):
    p = len(init)
    ll = -np.inf
    mat = np.zeros((iters, p))
    x = init
    if (verb):
        print(str(iters) + " iterations")
    for i in range(iters):
        if (verb):
            print(str(i), end=" ", flush=True)
        for j in range(thin):
            x, ll = kernel(x, ll)
        mat[i,:] = x
    if (verb):
        print("\nDone.", flush=True)
    return mat

We can use this code in the context of our logistic regression example as follows.

pre = np.array([10.,1.,1.,1.,1.,1.,5.,1.])

def rprop(beta):
    return beta + 0.02*pre*np.random.randn(p)

out = mcmc(init, mhKernel(lpost, rprop), thin=1000)

See the full runnable script for further details.

JAX

The above R and Python scripts are fine, but both languages are rather slow for this kind of workload. Fortunately it’s rather straightforward to convert the Python code to JAX to obtain quite amazing speed-up. We can write our M-H kernel as

def mhKernel(lpost, rprop, dprop = jit(lambda new, old: 1.)):
    @jit
    def kernel(key, x, ll):
        key0, key1 = jax.random.split(key)
        prop = rprop(key0, x)
        lp = lpost(prop)
        a = lp - ll + dprop(x, prop) - dprop(prop, x)
        accept = (jnp.log(jax.random.uniform(key1)) < a)
        return jnp.where(accept, prop, x), jnp.where(accept, lp, ll)
    return kernel

and our MCMC runner function as

def mcmc(init, kernel, thin = 10, iters = 10000):
    key = jax.random.PRNGKey(42)
    keys = jax.random.split(key, iters)
    @jit
    def step(s, k):
        [x, ll] = s
        x, ll = kernel(k, x, ll)
        s = [x, ll]
        return s, s
    @jit
    def iter(s, k):
        keys = jax.random.split(k, thin)
        _, states = jax.lax.scan(step, s, keys)
        final = [states[0][thin-1], states[1][thin-1]]
        return final, final
    ll = -np.inf
    x = init
    _, states = jax.lax.scan(iter, [x, ll], keys)
    return states[0]

There are really only two slightly tricky things about this code.

The first relates to the way JAX handles pseudo-random numbers. Since JAX is a pure functional eDSL, it can’t be used in conjunction with the typical pseudo-random number generators often used in imperative programming languages which rely on a global mutable state. This can be dealt with reasonably straightforwardly by explicitly passing around the random number state. There is a standard way of doing this that has been common practice in functional programming languages for decades. However, this standard approach is very sequential, and so doesn’t work so well in a parallel context. JAX therefore uses a splittable random number generator, where new states are created by splitting the current state into two (or more). We’ll come back to this when we get to the Haskell examples.

The second thing that might be unfamiliar to imperative programmers is the use of the scan operation (jax.lax.scan) to generate the Markov chain rather than a "for" loop. But scans are standard operations in most functional programming languages.

We can then call this code for our logistic regression example with

pre = jnp.array([10.,1.,1.,1.,1.,1.,5.,1.]).astype(jnp.float32)

@jit
def rprop(key, beta):
    return beta + 0.02*pre*jax.random.normal(key, [p])

out = mcmc(init, mhKernel(lpost, rprop), thin=1000)

See the full runnable script for further details.

Scala

In Scala we can use a similar approach to that already seen for defining a HoF to return a M-H kernel.

def mhKernel[S](
    logPost: S => Double, rprop: S => S,
    dprop: (S, S) => Double = (n: S, o: S) => 1.0
  ): ((S, Double)) => (S, Double) =
    val r = Uniform(0.0,1.0)
    state =>
      val (x0, ll0) = state
      val x = rprop(x0)
      val ll = logPost(x)
      val a = ll - ll0 + dprop(x0, x) - dprop(x, x0)
      if (math.log(r.draw()) < a)
        (x, ll)
      else
        (x0, ll0)

Note that Scala’s static typing does not prevent us from defining a function that is polymorphic in the type of the chain state, which we here call S. Also note that we are adopting a pragmatic approach to random number generation, exploiting the fact that Scala is not a pure functional language, using a mutable generator, and omitting to capture the non-determinism of the rprop function (and the returned kernel) in its type signature. In Scala this is a choice, and we could adopt a purer approach if preferred. We’ll see what such an approach will look like in Haskell, coming up next.

Now that we have the kernel, we don’t need to write an explicit runner function since Scala has good support for streaming data. There are many more-or-less sophisticated ways that we can work with data streams in Scala, and the choice depends partly on how pure one is being about tracking effects (such as non-determinism), but here I’ll just use the simple LazyList from the standard library for unfolding the kernel into an infinite MCMC chain before thinning and truncating appropriately.

  val pre = DenseVector(10.0,1.0,1.0,1.0,1.0,1.0,5.0,1.0)
  def rprop(beta: DVD): DVD = beta + pre *:* (DenseVector(Gaussian(0.0,0.02).sample(p).toArray))
  val kern = mhKernel(lpost, rprop)
  val s = LazyList.iterate((init, -Inf))(kern) map (_._1)
  val out = s.drop(150).thin(1000).take(10000)

See the full runnable script for further details.

Haskell

Since Haskell is a pure functional language, we need to have some convention regarding pseudo-random number generation. Haskell supports several styles. The most commonly adopted approach wraps a mutable generator up in a monad. The typical alternative is to use a pure functional generator and either explicitly thread the state through code or hide this in a monad similar to the standard approach. However, Haskell also supports the use of splittable generators, so we can consider all three approaches for comparative purposes. The approach taken does affect the code and the type signatures, and even the streaming data abstractions most appropriate for chain generation.

Starting with a HoF for producing a Metropolis kernel, an approach using the standard monadic generators could like like

mKernel :: (StatefulGen g m) => (s -> Double) -> (s -> g -> m s) -> 
           g -> (s, Double) -> m (s, Double)
mKernel logPost rprop g (x0, ll0) = do
  x <- rprop x0 g
  let ll = logPost(x)
  let a = ll - ll0
  u <- (genContVar (uniformDistr 0.0 1.0)) g
  let next = if ((log u) < a)
        then (x, ll)
        else (x0, ll0)
  return next

Note how non-determinism is captured in the type signatures by the monad m. The explicit pure approach is to thread the generator through non-deterministic functions.

mKernelP :: (RandomGen g) => (s -> Double) -> (s -> g -> (s, g)) -> 
            g -> (s, Double) -> ((s, Double), g)
mKernelP logPost rprop g (x0, ll0) = let
  (x, g1) = rprop x0 g
  ll = logPost(x)
  a = ll - ll0
  (u, g2) = uniformR (0, 1) g1
  next = if ((log u) < a)
        then (x, ll)
        else (x0, ll0)
  in (next, g2)

Here the updated random number generator state is returned from each non-deterministic function for passing on to subsequent non-deterministic functions. This explicit sequencing of operations makes it possible to wrap the generator state in a state monad giving code very similar to the stateful monadic generator approach, but as already discussed, the sequential nature of this approach makes it unattractive in parallel and concurrent settings.

Fortunately the standard Haskell pure generator is splittable, meaning that we can adopt a splitting approach similar to JAX if we prefer, since this is much more parallel-friendly.

mKernelP :: (RandomGen g) => (s -> Double) -> (s -> g -> s) -> 
            g -> (s, Double) -> (s, Double)
mKernelP logPost rprop g (x0, ll0) = let
  (g1, g2) = split g
  x = rprop x0 g1
  ll = logPost(x)
  a = ll - ll0
  u = unif g2
  next = if ((log u) < a)
        then (x, ll)
        else (x0, ll0)
  in next

Here non-determinism is signalled by passing a generator state (often called a "key" in the context of splittable generators) into a function. Functions receiving a key are responsible for splitting it to ensure that no key is ever used more than once.

Once we have a kernel, we need to unfold our Markov chain. When using the monadic generator approach, it is most natural to unfold using a monadic stream

mcmc :: (StatefulGen g m) =>
  Int -> Int -> s -> (g -> s -> m s) -> g -> MS.Stream m s
mcmc it th x0 kern g = MS.iterateNM it (stepN th (kern g)) x0

stepN :: (Monad m) => Int -> (a -> m a) -> (a -> m a)
stepN n fa = if (n == 1)
  then fa
  else (\x -> (fa x) >>= (stepN (n-1) fa))

whereas for the explicit approaches it is more natural to unfold into a regular infinite data stream. So, for the explicit sequential approach we could use

mcmcP :: (RandomGen g) => s -> (g -> s -> (s, g)) -> g -> DS.Stream s
mcmcP x0 kern g = DS.unfold stepUf (x0, g)
  where
    stepUf xg = let
      (x1, g1) = kern (snd xg) (fst xg)
      in (x1, (x1, g1))

and with the splittable approach we could use

mcmcP :: (RandomGen g) =>
  s -> (g -> s -> s) -> g -> DS.Stream s
mcmcP x0 kern g = DS.unfold stepUf (x0, g)
  where
    stepUf xg = let
      (x1, g1) = xg
      x2 = kern g1 x1
      (g2, _) = split g1
      in (x2, (x2, g2))

Calling these functions for our logistic regression example is similar to what we have seen before, but again there are minor syntactic differences depending on the approach. For further details see the full runnable scripts for the monadic approach, the pure sequential approach, and the splittable approach.

Dex

Dex is a pure functional language and uses a splittable random number generator, so the style we use is similar to JAX (or Haskell using a splittable generator). We can generate a Metropolis kernel with

def mKernel {s} (lpost: s -> Float) (rprop: Key -> s -> s) : 
    Key -> (s & Float) -> (s & Float) =
  def kern (k: Key) (sll: (s & Float)) : (s & Float) =
    (x0, ll0) = sll
    [k1, k2] = split_key k
    x = rprop k1 x0
    ll = lpost x
    a = ll - ll0
    u = rand k2
    select (log u < a) (x, ll) (x0, ll0)
  kern

We can then unfold our Markov chain with

def markov_chain {s} (k: Key) (init: s) (kern: Key -> s -> s) (its: Nat) :
    Fin its => s =
  with_state init \st.
    for i:(Fin its).
      x = kern (ixkey k i) (get st)
      st := x
      x

Here we combine Dex’s state effect with a for loop to unfold the stream. See the full runnable script for further details.

Next steps

As previously discussed, none of these codes are optimised, so care should be taken not to over-interpret running times. However, JAX and Dex are noticeably faster than the alternatives, even running on a single CPU core. Another interesting feature of both JAX and Dex is that they are differentiable. This makes it very easy to develop algorithms using gradient information. In subsequent posts we will think about the gradient of our example log-posterior and how we can use gradient information to develop "better" sampling algorithms.

The complete runnable scripts are all available from this public github repo.

Bayesian inference for a logistic regression model (Part 2)

Part 2: The log posterior

Introduction

This is the second part in a series of posts on MCMC-based Bayesian inference for a logistic regression model. If you are new to this series, please go back to Part 1.

In the previous post we looked at the basic modelling concepts, and how to fit the model using a variety of PPLs. In this post we will prepare for doing MCMC by considering the problem of computing the unnormalised log posterior for the model. We will then see how this posterior can be implemented in several different languages and libraries.

Derivation

Basic structure

In Bayesian inference the posterior distribution is just the conditional distribution of the model parameters given the data, and therefore proportional to the joint distribution of the model and data. We often write this as

\displaystyle \pi(\theta|y) \propto \pi(\theta,y) = \pi(\theta)\pi(y|\theta).

Taking logs we have

\displaystyle \log \pi(\theta, y) = \log \pi(\theta) + \log \pi(y|\theta).

So (up to an additive constant) the log posterior is just the sum of the log prior and log likelihood. There are many good (numerical) reasons why we try to work exclusively with the log posterior and try to avoid ever evaluating the raw posterior density.

For our example logistic regression model, the parameter vector \theta is just the vector of regression coefficients, \beta. We assumed independent mean zero normal priors for the components of this vector, so the log prior is just the sum of logs of normal densities. Many scientific libraries will have built-in functions for returning the log-pdf of standard distributions, but if an explicit form is required, the log of the density of a N(0,\sigma^2) at x is just

\displaystyle -\log(2\pi)/2 - \log|\sigma| - x^2/(2\sigma^2),

and the initial constant term normalising the density can often be dropped.

Log-likelihood (first attempt)

Information from the data comes into the log posterior via the log-likelihood. The typical way to derive the likelihood for problems of this type is to assume the usual binary encoding of the data (success 1, failure 0). Then, for a Bernoulli observation with probability p_i,\ i=1,\ldots,n, the likelihood associated with observation y_i is

\displaystyle f(y_i|p_i) = \left[ \hphantom{1-}p_i \quad :\ y_i=1 \atop 1-p_i \quad :\ y_i=0 \right. \quad = \quad p_i^{y_i}(1-p_i)^{1-y_i}.

Taking logs and then switching to parameter \eta_i=\text{logit}(p_i) we have

\displaystyle \log f(y_i|\eta_i) = y_i\eta_i - \log(1+e^{\eta_i}),

and summing over n observations gives the log likelihood

\displaystyle \log\pi(y|\eta) \equiv \ell(\eta;y) = y\cdot \eta - \mathbf{1}\cdot\log(\mathbf{1}+\exp{\eta}).

In the context of logistic regression, \eta is the linear predictor, so \eta=X\beta, giving

\displaystyle \ell(\beta;y) = y^\textsf{T}X\beta - \mathbf{1}^\textsf{T}\log(\mathbf{1}+\exp[X\beta]).

This is a perfectly good way of expressing the log-likelihood, and we will come back to it later when we want the gradient of the log-likelihood, but it turns out that there is a similar-but-different way of deriving it that results in an expression that is equivalent but slightly cheaper to evaluate.

Log-likelihood (second attempt)

For our second attempt, we will assume that the data is coded in a different way. Instead of the usual binary encoding, we will assume that the observation \tilde y_i is 1 for success and -1 for failure. This isn’t really a problem, since the two encodings are related by \tilde y_i = 2y_i-1. This new encoding is convenient in the context of a logit parameterisation since then

\displaystyle f(y_i|\eta_i) = \left[ p_i \ :\ \tilde y_i=1\atop 1-p_i\ :\ \tilde y_i=-1 \right. \ = \ \left[ (1+e^{-\eta_i})^{-1} \ :\ \tilde y_i=1\atop (1+e^{\eta_i})^{-1} \ :\ \tilde y_i=-1 \right. \ = \ (1+e^{-\tilde y_i\eta_i})^{-1} ,

and hence

\displaystyle \log f(y_i|\eta_i) = -\log(1+e^{-\tilde y_i\eta_i}).

Summing over observations gives

\displaystyle \ell(\eta;\tilde y) = -\mathbf{1}\cdot \log(\mathbf{1}+\exp[-\tilde y\circ \eta]),

where \circ denotes the Hadamard product. Substituting \eta=X\beta gives the log-likelihood

\displaystyle \ell(\beta;\tilde y) = -\mathbf{1}^\textsf{T} \log(\mathbf{1}+\exp[-\tilde y\circ X\beta]).

This likelihood is a bit cheaper to evaluate that the one previously derived. If we prefer to write in terms of the original data encoding, we can obviously do so as

\displaystyle \ell(\beta; y) = -\mathbf{1}^\textsf{T} \log(\mathbf{1}+\exp[-(2y-\mathbf{1})\circ (X\beta)]),

and in practice, it is this version that is typically used. To be clear, as an algebraic function of \beta and y the two functions are different. But they coincide for binary vectors y which is all that matters.

Implementation

R

In R we can create functions for evaluating the log-likelihood, log-prior and log-posterior as follows (assuming that X and y are in scope).

ll = function(beta)
    sum(-log(1 + exp(-(2*y - 1)*(X %*% beta))))

lprior = function(beta)
    dnorm(beta[1], 0, 10, log=TRUE) + sum(dnorm(beta[-1], 0, 1, log=TRUE))

lpost = function(beta) ll(beta) + lprior(beta)

Python

In Python (with NumPy and SciPy) we can define equivalent functions with

def ll(beta):
    return np.sum(-np.log(1 + np.exp(-(2*y - 1)*(X.dot(beta)))))

def lprior(beta):
    return (sp.stats.norm.logpdf(beta[0], loc=0, scale=10) + 
            np.sum(sp.stats.norm.logpdf(beta[range(1,p)], loc=0, scale=1)))

def lpost(beta):
    return ll(beta) + lprior(beta)

JAX

Python, like R, is a dynamic language, and relatively slow for MCMC algorithms. JAX is a tensor computation framework for Python that embeds a pure functional differentiable array processing language inside Python. JAX can JIT-compile high-performance code for both CPU and GPU, and has good support for parallelism. It is rapidly becoming the preferred way to develop high-performance sampling algorithms within the Python ecosystem. We can encode our log-posterior in JAX as follows.

@jit
def ll(beta):
    return jnp.sum(-jnp.log(1 + jnp.exp(-(2*y - 1)*jnp.dot(X, beta))))

@jit
def lprior(beta):
    return (jsp.stats.norm.logpdf(beta[0], loc=0, scale=10) + 
            jnp.sum(jsp.stats.norm.logpdf(beta[jnp.array(range(1,p))], loc=0, scale=1)))

@jit
def lpost(beta):
    return ll(beta) + lprior(beta)

Scala

JAX is a pure functional programming language embedded in Python. Pure functional programming languages are intrinsically more scalable and compositional than imperative languages such as R and Python, and are much better suited to exploit concurrency and parallelism. I’ve given a bunch of talks about this recently, so if you are interested in this, perhaps start with the materials for my Laplace’s Demon talk. Scala and Haskell are arguably the current best popular general purpose functional programming languages, so it is possibly interesting to consider the use of these languages for the development of scalable statistical inference codes. Since both languages are statically typed compiled functional languages with powerful type systems, they can be highly performant. However, neither is optimised for numerical (tensor) computation, so you should not expect that they will have performance comparable with optimised tensor computation frameworks such as JAX. We can encode our log-posterior in Scala (with Breeze) as follows:

  def ll(beta: DVD): Double =
      sum(-log(ones + exp(-1.0*(2.0*y - ones)*:*(X * beta))))

  def lprior(beta: DVD): Double =
    Gaussian(0,10).logPdf(beta(0)) + 
      sum(beta(1 until p).map(Gaussian(0,1).logPdf(_)))

  def lpost(beta: DVD): Double = ll(beta) + lprior(beta)

Spark

Apache Spark is a Scala library for distributed "big data" processing on clusters of machines. Despite fundamental differences, there is a sense in which Spark for Scala is a bit analogous to JAX for Python: both Spark and JAX are concerned with scalability, but they are targeting rather different aspects of scalability: JAX is concerned with getting regular sized data processing algorithms to run very fast (on GPUs), whereas Spark is concerned with running huge data processing tasks quickly by distributing work over clusters of machines. Despite obvious differences, the fundamental pure functional computational model adopted by both systems is interestingly similar: both systems are based on lazy transformations of immutable data structures using pure functions. This is a fundamental pattern for scalable data processing transcending any particular language, library or framework. We can encode our log posterior in Spark as follows.

    def ll(beta: DVD): Double = 
      df.map{row =>
        val y = row.getAs[Double](0)
        val x = BDV.vertcat(BDV(1.0),toBDV(row.getAs[DenseVector](1)))
        -math.log(1.0 + math.exp(-1.0*(2.0*y-1.0)*(x.dot(beta))))}.reduce(_+_)
    def lprior(beta: DVD): Double =
      Gaussian(0,10).logPdf(beta(0)) +
        sum(beta(1 until p).map(Gaussian(0,1).logPdf(_)))
    def lpost(beta: DVD): Double =
      ll(beta) + lprior(beta)

Haskell

Haskell is an old, lazy pure functional programming language with an advanced type system, and remains the preferred language for the majority of functional programming language researchers. Hmatrix is the standard high performance numerical linear algebra library for Haskell, so we can use it to encode our log-posterior as follows.

ll :: Matrix Double -> Vector Double -> Vector Double -> Double
ll x y b = (negate) (vsum (cmap log (
                              (scalar 1) + (cmap exp (cmap (negate) (
                                                         (((scalar 2) * y) - (scalar 1)) * (x #> b)
                                                         )
                                                     )))))

pscale :: [Double] -- prior standard deviations
pscale = [10.0, 1, 1, 1, 1, 1, 1, 1]
lprior :: Vector Double -> Double
lprior b = sum $  (\x -> logDensity (normalDistr 0.0 (snd x)) (fst x)) <$> (zip (toList b) pscale)
           
lpost :: Matrix Double -> Vector Double -> Vector Double -> Double
lpost x y b = (ll x y b) + (lprior b)

Again, a reminder that, here and elsewhere, there are various optimisations could be done that I’m not bothering with. This is all just proof-of-concept code.

Dex

JAX proves that a pure functional DSL for tensor computation can be extremely powerful and useful. But embedding such a language in a dynamic imperative language like Python has a number of drawbacks. Dex is an experimental statically typed stand-alone DSL for differentiable array and tensor programming that attempts to combine some of the correctness and composability benefits of powerful statically typed functional languages like Scala and Haskell with the performance benefits of tensor computation systems like JAX. It is currently rather early its development, but seems very interesting, and is already quite useable. We can encode our log-posterior in Dex as follows.

def ll (b: (Fin 8)=>Float) : Float =
  neg $  sum (log (map (\ x. (exp x) + 1) ((map (\ yi. 1 - 2*yi) y)*(x **. b))))

pscale = [10.0, 1, 1, 1, 1, 1, 1, 1] -- prior SDs
prscale = map (\ x. 1.0/x) pscale

def lprior (b: (Fin 8)=>Float) : Float =
  bs = b*prscale
  neg $  sum ((log pscale) + (0.5 .* (bs*bs)))

def lpost (b: (Fin 8)=>Float) : Float =
  (ll b) + (lprior b)

Next steps

Now that we have a way of evaluating the log posterior, we can think about constructing Markov chains having the posterior as their equilibrium distribution. In the next post we will look at one of the simplest ways of doing this: the Metropolis algorithm.

Complete runnable scripts are available from this public github repo.

Bayesian inference for a logistic regression model (Part 1)

Part 1: The basics

Introduction

This is the first in a series of posts on MCMC-based fully Bayesian inference for a logistic regression model. In this series we will look at the model, and see how the posterior distribution can be sampled using a variety of different programming languages and libraries.

Logistic regression

Logistic regression is concerned with predicting a binary outcome based on some covariate information. The probability of "success" is modelled via a logistic transformation of a linear predictor constructed from the covariate vector.

This is a very simple model, but is a convenient toy example since it is arguably the simplest interesting example of an intractable (nonlinear) statistical model requiring some kind of iterative numerical fitting method, even in the non-Bayesian setting. In a Bayesian context, the posterior distribution is intractable, necessitating either approximate or computationally intensive numerical methods of "solution". In this series of posts, we will mainly concentrate on MCMC algortithms for sampling the full posterior distribution of the model parameters given some observed data.

We assume n observations and p covariates (including an intercept term that is always 1). The binary observations y_i,\ i=1,\ldots,n are 1 for a "success" and 0 for a "failure". The covariate p-vectors x_i, i=1,\ldots,n all have 1 as the first element. The statistical model is

\displaystyle \text{logit}\left(\text{Pr}[Y_i = 1]\right) = x_i \cdot \beta,\quad i=1,\ldots,n,

where \beta is a p-vector of parameters, a\cdot b = a^\textsf{T} b, and

\displaystyle \text{logit}(q) \equiv \log\left(\frac{q}{1-q}\right),\quad \forall\ q\in (0,1).

Equivalently,

\displaystyle \text{Pr}[Y_i = 1] = \text{expit}(x_i \cdot \beta),\quad i=1,\ldots,n,

where

\displaystyle \text{expit}(\theta) \equiv \frac{1}{1+e^{-\theta}},\quad \forall\ \theta\in\mathbb{R}.

Note that the expit function is sometimes called the logistic or sigmoid function, but expit is slightly less ambiguous. The statistical problem is to choose the parameter vector \beta to provide the "best" model for the probability of success. In the Bayesian setting, a prior distribution (typically multivariate normal) is specified for \beta, and then the posterior distribution after conditioning on the data is the object of inferential interest.

Example problem

In order to illustrate the ideas, it is useful to have a small running example. Here we will use the (infamous) Pima training dataset (MASS::Pima.tr in R). Here there are n=200 observations and 7 predictors. Adding an intercept gives p=8 covariates. For the Bayesian analysis, we need a prior on \beta. We will assume independent mean zero normal distributions for each component. The prior standard deviation for the intercept will be 10 and for the other covariates will be 1.

Describing the model in some PPLs

In this first post in the series, we will use probabilistic programming languages (PPLs) to describe the model and sample the posterior distribution.

JAGS

JAGS is a stand-alone domain specific language (DSL) for probabilistic programming. It can be used independently of general purpose programming languages, or called from popular languages for data science such as Python and R. We can describe our model in JAGS with the following code.

  model {
    for (i in 1:n) {
      y[i] ~ dbern(pr[i])
      logit(pr[i]) <- inprod(X[i,], beta)
    }
    beta[1] ~ dnorm(0, 0.01)
    for (j in 2:p) {
      beta[j] ~ dnorm(0, 1)
    }
  }

Note that JAGS uses precision as the second parameter of a normal distribution. See the full runnable R script for further details. Given this model description, JAGS can construct an MCMC sampler for the posterior distribution of the model parameters given the data. See the full script for how to feed in the data, run the sampler, and analyse the output.

Stan

Stan is another stand-alone DSL for probabilistic programming, and has a very sophisticated sampling algorithm, making it a popular choice for non-trivial models. It uses gradient information for sampling, and therefore requires a differentiable log-posterior. We could encode our logistic regression model as follows.

data {
  int<lower=1> n;
  int<lower=1> p;
  int<lower=0, upper=1> y[n];
  real X[n,p];
}
parameters {
  real beta[p];
}
model {
  for (i in 1:n) {
    real eta = dot_product(beta, X[i,]);
    real pr = 1/(1+exp(-eta));
    y[i] ~ binomial(1, pr);
  }
  beta[1] ~ normal(0, 10);
  for (j in 2:p) {
    beta[j] ~ normal(0, 1);
  }
}

Note that Stan uses standard deviation as the second parameter of the normal distribution. See the full runnable R script for further details.

PyMC

JAGS and Stan are stand-alone DSLs for probabilistic programming. This has the advantage of making them independent of any particular host (general purpose) programming language. But it also means that they are not able to take advantage of the language and tool support of an existing programming language. An alternative to stand-alone DSLs are embedded DSLs (eDSLs). Here, a DSL is embedded as a library or package within an existing (general purpose) programming language. Then, ideally, in the context of PPLs, probabilistic programs can become ordinary values within the host language, and this can have some advantages, especially if the host language is sophisticated. A number of probabilistic programming languages have been implemented as eDSLs in Python. Python is not a particularly sophisticated language, so the advantages here are limited, but not negligible.

PyMC is probably the most popular eDSL PPL in Python. We can encode our model in PyMC as follows.

pscale = np.array([10.,1.,1.,1.,1.,1.,1.,1.])
with pm.Model() as model:
    beta = pm.Normal('beta', 0, pscale, shape=p)
    eta = pmm.matrix_dot(X, beta)
    pm.Bernoulli('y', logit_p=eta, observed=y)
    traces = pm.sample(2500, tune=1000, init="adapt_diag", return_inferencedata=True)

See the full runnable script for further details.

NumPyro

NumPyro is a fork of Pyro for NumPy and JAX (of which more later). We can encode our model with NumPyro as follows.

pscale = jnp.array([10.,1.,1.,1.,1.,1.,1.,1.]).astype(jnp.float32)
def model(X, y):
    coefs = numpyro.sample("beta", dist.Normal(jnp.zeros(p), pscale))
    logits = jnp.dot(X, coefs)
    return numpyro.sample("obs", dist.Bernoulli(logits=logits), obs=y)

See the full runnable script for further details.

Please note that none of the above examples have been optimised, or are even necessarily expressed idiomatically within each PPL. I’ve just tried to express the model in as simple and similar way across each PPL. For example, I know that there is a function bernoulli_logit_glm in Stan which would simplify the model and improve sampling efficiency, but I’ve deliberately not used it in order to try and keep the implementations as basic as possible. The same will be true for all of the code examples in this series of blog posts. The code has not been optimised and should therefore not be used for serious benchmarking.

Next steps

PPLs are convenient, and are becoming increasingly sophisticated. Each of the above PPLs provides a simple way to pass in observed data, and automatically construct an MCMC algorithm for sampling from the implied posterior distribution – see the full scripts for details. All of the PPLs work well for this problem, and all produce identical results up to Monte Carlo error. Each PPL has its own approach to sampler construction, and some PPLs offer multiple choices. However, more challenging problems often require highly customised samplers. Such samplers will often need to be created from scratch, and will require (at least) the ability to compute the (unnormalised log) posterior density at proposed parameter values, so in the next post we will look at how this can be derived for this model (in a couple of different ways) and coded up from scratch in a variety of programming languages.

All of the complete, runnable code associated with this series of blog posts can be obtained from this public github repo.

Unbiased MCMC with couplings

Yesterday there was an RSS Read Paper meeting for the paper Unbiased Markov chain Monte Carlo with couplings by Pierre Jacob, John O’Leary and Yves F. AtchadĂŠ. The paper addresses the bias in MCMC estimates due to lack of convergence to equilibrium (the “burn-in” problem), and shows how it is possible to modify MCMC algorithms in order to construct estimates which exactly remove this bias. The requirement is to couple a pair of MCMC chains so that they will at some point meet exactly and thereafter remain coupled. This turns out to be easier to do that one might naively expect. There are many reasons why we might want to remove bias from MCMC estimates, but the primary motivation in the paper was the application to parallel MCMC computation. The idea here is that many pairs of chains can be run independently on any available processors, and the unbiased estimates from the different pairs can be safely averaged to get an (improved) unbiased estimate based on all of the chains. As a discussant of the paper, I’ve spent a bit of time thinking about this idea, and have created a small repository of materials relating to the paper which may be useful for others interested in understanding the method and how to use it in practice.

The repo includes a page of links to related papers, blog posts, software and other resources relating to unbiased MCMC that I’ve noticed on-line.

Earlier in the year I gave an internal seminar at Newcastle giving a tutorial introduction to the main ideas from the paper, including runnable R code implementations of the examples. The talk was prepared as an executable R Markdown document. The R Markdown source code is available in the repo, but for the convenience of casual browsers I’ve also included a pre-built set of PDF slides. Code examples include code for maximal coupling of two (univariate) distributions, coupling Metropolis-Hastings chains, and coupling a Gibbs sampler for an AR(1) process.

I haven’t yet finalised my written discussion contribution, but the slides I presented at the Read Paper meeting are also available. Again, there is source code and pre-built PDF slides. My discussion focused on seeing how well the technique works for Gibbs samplers applied to high-dimensional latent process models (an AR(1) process and a Gaussian Markov random field), and reflecting on the extent to which the technique really solves the burn-in/parallel MCMC problem.

The repo also contains a few stand-alone code examples. There are some simple tutorial examples in R (largely derived from my tutorial introduction), including implementation of (univariate) independent and reflection maximal couplings, and a coupled AR(1) process example.

The more substantial example concerns a coupled Gibbs sampler for a GMRF. This example is written in the Scala programming language. There are a couple of notable features of this implementation. First, the code illustrates monadic coupling of probability distributions, based on the Rand type in the Breeze scientific library. This provides an elegant way to max couple arbitrary (continuous) random variables, and to create coupled Metropolis(-Hastings) kernels. For example, a coupling of two distributions can be constructed as

  def couple[T](p: ContinuousDistr[T], q: ContinuousDistr[T]): Rand[(T, T)] = {
    def ys: Rand[T] =
      for {
        y  <- q
        w  <- Uniform(0, 1)
        ay <- if (math.log(w) > p.logPdf(y) - q.logPdf(y)) Rand.always(y) else ys
      } yield ay
    val pair = for {
      x <- p
      w <- Uniform(0, 1)
    } yield (math.log(w) <= q.logPdf(x) - p.logPdf(x), x)
    pair flatMap {
      case (b, x) => if (b) Rand.always((x, x)) else (ys map (y => (x, y)))
    }
  }

and then draws can be sampled from the resulting Rand[(T, T)] polymorphic type as required. Incidentally, this also illustrates how to construct an independent maximal coupling without evaluating any raw likelihoods.
The other notable feature of the code is the use of a parallel comonadic image type for parallel Gibbs sampling of the GMRF, producing a (lazy) Stream of coupled MCMC samples.

Write your own general-purpose monadic probabilistic programming language from scratch in 50 lines of (Scala) code

Background

In May I attended a great workshop on advances and challenges in machine learning languages at the CMS in Cambridge. There was an a good mix of people from different disciplines, and a bit of a theme around probabilistic programming. The workshop schedule includes links to many of the presentations, and is generally worth browsing. In particular, it includes a link to the slides for my presentation on a compositional approach to scalable Bayesian computation and probabilistic programming. I’ve given a few talks on this kind of thing over the last couple of years, at Newcastle, at the Isaac Newton Institute in Cambridge (twice), and at CIRM in France. But I think I explained things best at this workshop at the CMS, though my impression could partly have been a reflection of the more interested and relevant audience. In the talk I started with a basic explanation of why ideas from category theory and functional programming can help to solve problems in statistical computing in a more composable and scalable way, before moving on to discuss probability monads and their fundamental connection to probabilistic programming. The take home message from the talk is that if you have a generic inference algorithm, expressing the logic in the context of probability monads can give you an embedded probabilistic programming language (PPL) for that inference algorithm essentially “for free”.

So, during my talk I said something a little fool-hardy. I can’t remember my exact words, but while presenting the idea behind an SMC-based probability monad I said something along the lines of “one day I will write a blog post on how to write a probabilistic programming language from scratch in 50 lines of code, and this is how I’ll do it“! Rather predictably (with hindsight), immediately after my talk about half a dozen people all pleaded with me to urgently write the post! I’ve been a little busy since then, but now that things have settled down a little for the summer, I’ve some time to think and code, so here is that post.

Introduction

The idea behind this post is to show that, if you think about the problem in the right way, and use a programming language with syntactic support for monadic composition, then producing a flexible, general, compositional, embedded domain specific language (DSL) for probabilistic programming based on a given generic inference algorithm is no more effort than hard-coding two or three illustrative examples. You would need to code up two or three examples for a paper anyway, but providing a PPL is way more useful. There is also an interesting converse to this, which is that if you can’t easily produce a PPL for your “general” inference algorithm, then perhaps it isn’t quite as “general” as you thought. I’ll try to resist exploring that here…

To illustrate these principles I want to develop a fairly minimal PPL, so that the complexities of the inference algorithm don’t hide the simplicity of the PPL embedding. Importance sampling with resampling is probably the simplest useful generic Bayesian inference algorithm to implement, so that’s what I’ll use. Note that there are many limitations of the approach that I will adopt, which will make it completely unsuitable for “real” problems. In particular, this implementation is: inefficient, in terms of both compute time and memory usage, statistically inefficient for deep nesting and repeated conditioning, due to the particle degeneracy problem, specific to a particular probability monad, strictly evaluated, impure (due to mutation of global random number state), etc. All of these things are easily fixed, but all at the expense of greater abstraction, complexity and lines of code. I’ll probably discuss some of these generalisations and improvements in future posts, but for this post I want to keep everything as short and simple as practical. It’s also worth mentioning that there is nothing particularly original here. Many people have written about monadic embedded PPLs, and several have used an SMC-based monad for illustration. I’ll give some pointers to useful further reading at the end.

The language, in 50 lines of code

Without further ado, let’s just write the PPL. I’m using plain Scala, with just a dependency on the Breeze scientific library, which I’m going to use for simulating random numbers from standard distributions, and evaluation of their log densities. I have a directory of materials associated with this post in a git repo. This post is derived from an executable tut document (so you know it works), which can be found here. If you just want to follow along copying code at the command prompt, just run sbt from an empty or temp directory, and copy the following to spin up a Scala console with the Breeze dependency:

set libraryDependencies += "org.scalanlp" %% "breeze" % "1.0-RC4"
set libraryDependencies += "org.scalanlp" %% "breeze-natives" % "1.0-RC4"
set scalaVersion := "2.13.0"
console

We start with a couple of Breeze imports

import breeze.stats.{distributions => bdist}
import breeze.linalg.DenseVector

which are not strictly necessary, but clean up the subsequent code. We are going to use a set of weighted particles to represent a probability distribution empirically, so we’ll start by defining an appropriate ADT for these:

implicit val numParticles = 300

case class Particle[T](v: T, lw: Double) { // value and log-weight
  def map[S](f: T => S): Particle[S] = Particle(f(v), lw)
}

We also include a map method for pushing a particle through a transformation, and a default number of particles for sampling and resampling. 300 particles are enough for illustrative purposes. Ideally it would be good to increase this for more realistic experiments. We can use this particle type to build our main probability monad as follows.

trait Prob[T] {
  val particles: Vector[Particle[T]]
  def map[S](f: T => S): Prob[S] = Empirical(particles map (_ map f))
  def flatMap[S](f: T => Prob[S]): Prob[S] = {
    Empirical((particles map (p => {
      f(p.v).particles.map(psi => Particle(psi.v, p.lw + psi.lw))
    })).flatten).resample
  }
  def resample(implicit N: Int): Prob[T] = {
    val lw = particles map (_.lw)
    val mx = lw reduce (math.max(_,_))
    val rw = lw map (lwi => math.exp(lwi - mx))
    val law = mx + math.log(rw.sum/(rw.length))
    val ind = bdist.Multinomial(DenseVector(rw.toArray)).sample(N)
    val newParticles = ind map (i => particles(i))
    Empirical(newParticles.toVector map (pi => Particle(pi.v, law)))
  }
  def cond(ll: T => Double): Prob[T] =
    Empirical(particles map (p => Particle(p.v, p.lw + ll(p.v))))
  def empirical: Vector[T] = resample.particles.map(_.v)
}

case class Empirical[T](particles: Vector[Particle[T]]) extends Prob[T]

Note that if you are pasting into the Scala REPL you will need to use :paste mode for this. So Prob[_] is our base probability monad trait, and Empirical[_] is our simplest implementation, which is just a collection of weighted particles. The method flatMap forms the naive product of empirical measures and then resamples in order to stop an explosion in the number of particles. There are two things worth noting about the resample method. The first is that the log-sum-exp trick is being used to avoid overflow and underflow when the log weights are exponentiated. The second is that although the method returns an equally weighted set of particles, the log weights are all set in order that the average raw weight of the output set matches the average raw weight of the input set. This is a little tricky to explain, but it turns out to be necessary in order to correctly propagate conditioning information back through multiple monadic binds (flatMaps). The cond method allows conditioning of a distribution using an arbitrary log-likelihood. It is included for comparison with some other implementations I will refer to later, but we won’t actually be using it, so we could save two lines of code here if necessary. The empirical method just extracts an unweighted set of values from a distribution for subsequent analysis.

It will be handy to have a function to turn a bunch of unweighted particles into a set of particles with equal weights (a sort-of inverse of the empirical method just described), so we can define that as follows.

def unweighted[T](ts: Vector[T], lw: Double = 0.0): Prob[T] =
  Empirical(ts map (Particle(_, lw)))

Probabilistic programming is essentially trivial if we only care about forward sampling. But interesting PPLs allow us to condition on observed values of random variables. In the context of SMC, this is simplest when the distribution being conditioned has a tractable log-likelihood. So we can now define an extension of our probability monad for distributions with a tractable log-likelihood, and define a bunch of convenient conditioning (or “fitting”) methods using it.

trait Dist[T] extends Prob[T] {
  def ll(obs: T): Double
  def ll(obs: Seq[T]): Double = obs map (ll) reduce (_+_)
  def fit(obs: Seq[T]): Prob[T] =
    Empirical(particles map (p => Particle(p.v, p.lw + ll(obs))))
  def fitQ(obs: Seq[T]): Prob[T] = Empirical(Vector(Particle(obs.head, ll(obs))))
  def fit(obs: T): Prob[T] = fit(List(obs))
  def fitQ(obs: T): Prob[T] = fitQ(List(obs))
}

The only unimplemented method is ll(). The fit method re-weights a particle set according to the observed log-likelihood. For convenience, it also returns a particle cloud representing the posterior-predictive distribution of an iid value from the same distribution. This is handy, but comes at the expense of introducing an additional particle cloud. So, if you aren’t interested in the posterior predictive, you can avoid this cost by using the fitQ method (for “fit quick”), which doesn’t return anything useful. We’ll see examples of this in practice, shortly. Note that the fitQ methods aren’t strictly required for our “minimal” PPL, so we can save a couple of lines by omitting them if necessary. Similarly for the variants which allow conditioning on a collection of iid observations from the same distribution.

At this point we are essentially done. But for convenience, we can define a few standard distributions to help get new users of our PPL started. Of course, since the PPL is embedded, it is trivial to add our own additional distributions later.

case class Normal(mu: Double, v: Double)(implicit N: Int) extends Dist[Double] {
  lazy val particles = unweighted(bdist.Gaussian(mu, math.sqrt(v)).sample(N).toVector).particles
  def ll(obs: Double) = bdist.Gaussian(mu, math.sqrt(v)).logPdf(obs) }

case class Gamma(a: Double, b: Double)(implicit N: Int) extends Dist[Double] {
  lazy val particles = unweighted(bdist.Gamma(a, 1.0/b).sample(N).toVector).particles
  def ll(obs: Double) = bdist.Gamma(a, 1.0/b).logPdf(obs) }

case class Poisson(mu: Double)(implicit N: Int) extends Dist[Int] {
  lazy val particles = unweighted(bdist.Poisson(mu).sample(N).toVector).particles
  def ll(obs: Int) = bdist.Poisson(mu).logProbabilityOf(obs) }

Note that I’ve parameterised the Normal and Gamma the way that statisticians usually do, and not the way they are usually parameterised in scientific computing libraries (such as Breeze).

That’s it! This is a complete, general-purpose, composable, monadic PPL, in 50 (actually, 48, and fewer still if you discount trailing braces) lines of code. Let’s now see how it works in practice.

Examples

Normal random sample

We’ll start off with just about the simplest slightly interesting example I can think of: Bayesian inference for the mean and variance of a normal distribution from a random sample.

import breeze.stats.{meanAndVariance => meanVar}
// import breeze.stats.{meanAndVariance=>meanVar}

val mod = for {
  mu <- Normal(0, 100)
  tau <- Gamma(1, 0.1)
  _ <- Normal(mu, 1.0/tau).fitQ(List(8.0,9,7,7,8,10))
} yield (mu,tau)
// mod: Wrapped.Prob[(Double, Double)] = Empirical(Vector(Particle((8.718127116254472,0.93059589932682),-15.21683812389373), Particle((7.977706390420308,1.1575288208065433),-15.21683812389373), Particle((7.977706390420308,1.1744750937611985),-15.21683812389373), Particle((7.328100552769214,1.1181787982959164),-15.21683812389373), Particle((7.977706390420308,0.8283737237370494),-15.21683812389373), Particle((8.592847414557049,2.2934836446009026),-15.21683812389373), Particle((8.718127116254472,1.498741032928539),-15.21683812389373), Particle((8.592847414557049,0.2506065368748732),-15.21683812389373), Particle((8.543283880264225,1.127386759627675),-15.21683812389373), Particle((7.977706390420308,1.3508728798704925),-15.21683812389373), Particle((7.977706390420308,1.1134430556990933),-15.2168...

val modEmp = mod.empirical
// modEmp: Vector[(Double, Double)] = Vector((7.977706390420308,0.8748006833362748), (6.292345096890432,0.20108091703626174), (9.15330820843396,0.7654238730107492), (8.960935105658741,1.027712984079369), (7.455292602273359,0.49495749079351836), (6.911716909394562,0.7739749058662421), (6.911716909394562,0.6353785792877397), (7.977706390420308,1.1744750937611985), (7.977706390420308,1.1134430556990933), (8.718127116254472,1.166399872049532), (8.763777227034538,1.0468304705769353), (8.718127116254472,0.93059589932682), (7.328100552769214,1.6166695922250236), (8.543283880264225,0.4689300351248357), (8.543283880264225,2.0028918490755094), (7.536025958690963,0.6282318170458533), (7.328100552769214,1.6166695922250236), (7.049843463553113,0.20149378088848635), (7.536025958690963,2.3565657669819897...

meanVar(modEmp map (_._1)) // mu
// res0: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(8.311171010932343,0.4617800639333532,300)

meanVar(modEmp map (_._2)) // tau
// res1: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(0.940762723934599,0.23641881704888842,300)

Note the use of the empirical method to turn the distribution into an unweighted set of particles for Monte Carlo analysis. Anyway, the main point is that the syntactic sugar for monadic binds (flatMaps) provided by Scala’s for-expressions (similar to do-notation in Haskell) leads to readable code not so different to that in well-known general-purpose PPLs such as BUGS, JAGS, or Stan. There are some important differences, however. In particular, the embedded DSL has probabilistic programs as regular values in the host language. These may be manipulated and composed like other values. This makes this probabilistic programming language more composable than the aforementioned languages, which makes it much simpler to build large, complex probabilistic programs from simpler, well-tested, components, in a scalable way. That is, this PPL we have obtained “for free” is actually in many ways better than most well-known PPLs.

Noisy measurements of a count

Here we’ll look at the problem of inference for a discrete count given some noisy iid continuous measurements of it.

val mod = for {
  count <- Poisson(10)
  tau <- Gamma(1, 0.1)
  _ <- Normal(count, 1.0/tau).fitQ(List(4.2,5.1,4.6,3.3,4.7,5.3))
} yield (count, tau)
// mod: Wrapped.Prob[(Int, Double)] = Empirical(Vector(Particle((5,4.488795220669575),-11.591037521513753), Particle((5,1.7792314573063672),-11.591037521513753), Particle((5,2.5238021156137673),-11.591037521513753), Particle((4,3.280754333896923),-11.591037521513753), Particle((5,2.768438569482849),-11.591037521513753), Particle((4,1.3399975573518912),-11.591037521513753), Particle((5,1.1792835858615431),-11.591037521513753), Particle((5,1.989491156206883),-11.591037521513753), Particle((4,0.7825254987152054),-11.591037521513753), Particle((5,2.7113936834028793),-11.591037521513753), Particle((5,3.7615196800240387),-11.591037521513753), Particle((4,1.6833300961124709),-11.591037521513753), Particle((5,2.749183220798113),-11.591037521513753), Particle((5,2.1074062883430202),-11.591037521513...

val modEmp = mod.empirical
// modEmp: Vector[(Int, Double)] = Vector((4,3.243786594839479), (4,1.5090869158886693), (4,1.280656912383482), (5,2.0616356908358195), (5,3.475433097869503), (5,1.887582611202514), (5,2.8268877720514745), (5,0.9193261688050818), (4,1.7063629502805908), (5,2.116414832864841), (5,3.775508828984636), (5,2.6774941123762814), (5,2.937859946593459), (5,1.2047689975166402), (5,2.5658806161572656), (5,1.925890364268593), (4,1.0194093176888832), (5,1.883288825936725), (5,4.9503779454422965), (5,0.9045613180858916), (4,1.5795027943928661), (5,1.925890364268593), (5,2.198539449287062), (5,1.791363956348445), (5,0.9853760689818026), (4,1.6541388923071607), (5,2.599899960899971), (4,1.8904423810277957), (5,3.8983183765907836), (5,1.9242319515895554), (5,2.8268877720514745), (4,1.772120802027519), (5,2...

meanVar(modEmp map (_._1.toDouble)) // count
// res2: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(4.670000000000004,0.23521739130434777,300)

meanVar(modEmp map (_._2)) // tau
// res3: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(1.9678279101913874,0.9603971613375548,300)

I’ve included this mainly as an example of inference for a discrete-valued parameter. There are people out there who will tell you that discrete parameters are bad/evil/impossible. This isn’t true – discrete parameters are cool!

Linear model

Because our PPL is embedded, we can take full advantage of the power of the host programming language to build our models. Let’s explore this in the context of Bayesian estimation of a linear model. We’ll start with some data.

val x = List(1.0,2,3,4,5,6)
// x: List[Double] = List(1.0, 2.0, 3.0, 4.0, 5.0, 6.0)

val y = List(3.0,2,4,5,5,6)
// y: List[Double] = List(3.0, 2.0, 4.0, 5.0, 5.0, 6.0)

val xy = x zip y
// xy: List[(Double, Double)] = List((1.0,3.0), (2.0,2.0), (3.0,4.0), (4.0,5.0), (5.0,5.0), (6.0,6.0))

Now, our (simple) linear regression model will be parameterised by an intercept, alpha, a slope, beta, and a residual variance, v. So, for convenience, let’s define an ADT representing a particular linear model.

case class Param(alpha: Double, beta: Double, v: Double)
// defined class Param

Now we can define a prior distribution over models as follows.

val prior = for {
  alpha <- Normal(0,10)
  beta <- Normal(0,4)
  v <- Gamma(1,0.1)
} yield Param(alpha, beta, v)
// prior: Wrapped.Prob[Param] = Empirical(Vector(Particle(Param(-2.392517550699654,-3.7516090283880095,1.724680963054379),0.0), Particle(Param(7.60982717067903,-1.4318199629361292,2.9436745225038545),0.0), Particle(Param(-1.0281832158124837,-0.2799562317845073,4.05125312048092),0.0), Particle(Param(-1.0509321093485073,-2.4733837587060448,0.5856868459456287),0.0), Particle(Param(7.678898742733517,0.15616204936412104,5.064540017623097),0.0), Particle(Param(-3.392028985658713,-0.694412176170572,7.452625596437611),0.0), Particle(Param(3.0310535934425324,-2.97938526497514,2.138446100857938),0.0), Particle(Param(3.016959696424399,1.3370878561954143,6.18957854813488),0.0), Particle(Param(2.6956505371497066,1.058845844793446,5.257973123790336),0.0), Particle(Param(1.496225540527873,-1.573936445746...

Since our language doesn’t include any direct syntactic support for fitting regression models, we can define our own function for conditioning a distribution over models on a data point, which we can then apply to our prior as a fold over the available data.

def addPoint(current: Prob[Param], obs: (Double, Double)): Prob[Param] = for {
    p <- current
    (x, y) = obs
    _ <- Normal(p.alpha + p.beta * x, p.v).fitQ(y)
  } yield p
// addPoint: (current: Wrapped.Prob[Param], obs: (Double, Double))Wrapped.Prob[Param]

val mod = xy.foldLeft(prior)(addPoint(_,_)).empirical
// mod: Vector[Param] = Vector(Param(1.4386051853067798,0.8900831186754122,4.185564696221981), Param(0.5530582357040271,1.1296886766045509,3.468527573093037), Param(0.6302560079049571,0.9396563485293532,3.7044543917875927), Param(3.68291303096638,0.4781372802435529,5.151665328789926), Param(3.016959696424399,0.4438016738989412,1.9988914122633519), Param(3.016959696424399,0.4438016738989412,1.9988914122633519), Param(0.6302560079049571,0.9396563485293532,3.7044543917875927), Param(0.6302560079049571,0.9396563485293532,3.7044543917875927), Param(3.68291303096638,0.4781372802435529,5.151665328789926), Param(3.016959696424399,0.4438016738989412,1.9988914122633519), Param(0.6302560079049571,0.9396563485293532,3.7044543917875927), Param(0.6302560079049571,0.9396563485293532,3.7044543917875927), ...

meanVar(mod map (_.alpha))
// res4: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(1.5740812481283812,1.893684802867127,300)

meanVar(mod map (_.beta))
// res5: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(0.7690238868623273,0.1054479268115053,300)

meanVar(mod map (_.v))
// res6: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(3.5240853748668695,2.793386340338213,300)

We could easily add syntactic support to our language to enable the fitting of regression-style models, as is done in Rainier, of which more later.

Dynamic generalised linear model

The previous examples have been fairly simple, so let’s finish with something a bit less trivial. Our language is quite flexible enough to allow the analysis of a dynamic generalised linear model (DGLM). Here we’ll fit a Poisson DGLM with a log-link and a simple Brownian state evolution. More complex models are more-or-less similarly straightforward. The model is parameterised by an initial state, state0, and and evolution variance, w.

val data = List(2,1,0,2,3,4,5,4,3,2,1)
// data: List[Int] = List(2, 1, 0, 2, 3, 4, 5, 4, 3, 2, 1)

val prior = for {
  w <- Gamma(1, 1)
  state0 <- Normal(0.0, 2.0)
} yield (w, List(state0))
// prior: Wrapped.Prob[(Double, List[Double])] = Empirical(Vector(Particle((0.12864918092587044,List(-2.862479260552014)),0.0), Particle((1.1706344622093179,List(1.6138397233532091)),0.0), Particle((0.757288087950638,List(-0.3683499919402798)),0.0), Particle((2.755201217523856,List(-0.6527488751780317)),0.0), Particle((0.7535085397802043,List(0.5135562407906502)),0.0), Particle((1.1630726564525629,List(0.9703146201262348)),0.0), Particle((1.0080345715326213,List(-0.375686732266234)),0.0), Particle((4.603723117526974,List(-1.6977366375222938)),0.0), Particle((0.2870669117815037,List(2.2732160435099433)),0.0), Particle((2.454675218313211,List(-0.4148287542786906)),0.0), Particle((0.3612534201761152,List(-1.0099270904161748)),0.0), Particle((0.29578453393473114,List(-2.4938128878051966)),0.0)...

We can define a function to create a new hidden state, prepend it to the list of hidden states, and condition on the observed value at that time point as follows.

def addTimePoint(current: Prob[(Double, List[Double])],
  obs: Int): Prob[(Double, List[Double])] = for {
  tup <- current
  (w, states) = tup
  os = states.head
  ns <- Normal(os, w)
  _ <- Poisson(math.exp(ns)).fitQ(obs)
} yield (w, ns :: states)
// addTimePoint: (current: Wrapped.Prob[(Double, List[Double])], obs: Int)Wrapped.Prob[(Double, List[Double])]

We then run our (augmented state) particle filter as a fold over the time series.

val mod = data.foldLeft(prior)(addTimePoint(_,_)).empirical
// mod: Vector[(Double, List[Double])] = Vector((0.053073252551193446,List(0.8693030057529023, 1.2746526177834938, 1.020307245610461, 1.106341696651584, 1.070777529635013, 0.8749041525303247, 0.9866999164354662, 0.4082577920509255, 0.06903234462140699, -0.018835642776197814, -0.16841912034400547, -0.08919045681401294)), (0.0988871875952762,List(-0.24241948109998607, 0.09321618969352086, 0.9650532206325375, 1.1738734442767293, 1.2272325310228442, 0.9791695328246326, 0.5576319082578128, -0.0054280215024367084, 0.4256621012454391, 0.7486862644576158, 0.8193517409118243, 0.5928750312493785)), (0.16128799384962295,List(-0.30371187329667104, -0.3976854602292066, 0.5869357473774455, 0.9881090696832543, 1.2095181380307558, 0.7211231597865506, 0.8085486452269925, 0.2664373341459165, -0.627344024142...

meanVar(mod map (_._1)) // w
// res7: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(0.29497487517435844,0.0831412016262515,300)

meanVar(mod map (_._2.reverse.head)) // state0 (initial state)
// res8: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(0.04617218427664018,0.372844704533101,300)

meanVar(mod map (_._2.head)) // stateN (final state)
// res9: breeze.stats.meanAndVariance.MeanAndVariance = MeanAndVariance(0.4937178761565612,0.2889287607470016,300)

Summary, conclusions, and further reading

So, we’ve seen how we can build a fully functional, general-purpose, compositional, monadic PPL from scratch in 50 lines of code, and we’ve seen how we can use it to solve real, analytically intractable Bayesian inference problems of non-trivial complexity. Of course, there are many limitations to using exactly this PPL implementation in practice. The algorithm becomes intolerably slow for deeply nested models, and uses unreasonably large amounts of RAM for large numbers of particles. It also suffers from a particle degeneracy problem if there are too many conditioning events. But it is important to understand that these are all deficiencies of the naive inference algorithm used, not the PPL itself. The PPL is flexible and compositional and can be used to build models of arbitrary size and complexity – it just needs to be underpinned by a better, more efficient, inference algorithm. Rainier is a Scala library I’ve blogged about previously which uses a very similar PPL to the one described here, but is instead underpinned by a fast, efficient, HMC algorithm. With my student Jonny Law, we have recently arXived a paper on Functional probabilistic programming for scalable Bayesian modelling, discussing some of these issues, and exploring the compositional nature of monadic PPLs (somewhat glossed over in this post).

Since the same PPL can be underpinned by different inference algorithms encapsulated as probability monads, an obvious question is whether it is possible to abstract the PPL away from the inference algorithm implementation. Of course, the answer is “yes”, and this has been explored to great effect in papers such as Practical probabilistic programming with monads and Functional programming for modular Bayesian inference. Note that they use the cond approach to conditioning, which looks a bit unwieldy, but is equivalent to fitting. As well as allowing alternative inference algorithms to be applied to the same probabilistic program, it also enables the composing of inference algorithms – for example, composing a MH algorithm with an SMC algorithm in order to get a PMMH algorithm. The ideas are implemented in an embedded DSL for Haskell, monad-bayes. If you are not used to Haskell, the syntax will probably seem a bit more intimidating than Scala’s, but the semantics are actually quite similar, with the main semantic difference being that Scala is strictly evaluated by default, whereas Haskell is lazily evaluated by default. Both languages support both lazy and strict evaluation – the difference relates simply to default behaviour, but is important nevertheless.

Papers

Software

  • min-ppl – code associated with this blog post
  • Rainier – a more efficient PPL with similar syntax
  • monad-bayes – a Haskell library exploring related ideas

The scala-smfsb library

In the previous post I gave a very quick introduction to the smfsb R package. As mentioned in that post, although good for teaching and learning, R isn’t a great language for serious scientific computing or computational statistics. So for the publication of the third edition of my textbook, Stochastic modelling for systems biology, I have created a library in the Scala programming language replicating the functionality provided by the R package. Here I will give a very quick introduction to the scala-smfsb library. Some familiarity with both Scala and the smfsb R package will be helpful, but is not strictly necessary. Note that the library relies on the Scala Breeze library for linear algebra and probability distributions, so some familiarity with that library can also be helpful.

Setup

To follow the along you need to have Sbt installed, and this in turn requires a recent JDK. If you are new to Scala, you may find the setup page for my Scala course to be useful, but note that on many Linux systems it can be as simple as installing the packages openjdk-8-jdk and sbt.

Once you have Sbt installed, you should be able to run it by entering sbt at your OS command line. You now need to use Sbt to create a Scala REPL with a dependency on the scala-smfsb library. There are many ways to do this, but if you are new to Scala, the simplest way is probably to start up Sbt from an empty or temporary directory (which doesn’t contain any Scala code), and then paste the following into the Sbt prompt:

set libraryDependencies += "com.github.darrenjw" %% "scala-smfsb" % "0.6"
set libraryDependencies += "org.scalanlp" %% "breeze-viz" % "0.13.2"
set scalaVersion := "2.12.6"
set scalacOptions += "-Yrepl-class-based"
console

The first time you run this it will take a little while to download and cache various library dependencies. But everything is cached, so it should be much quicker in future. When it is finished, you should have a Scala REPL ready to enter Scala code.

An introduction to scala-smfsb

It should be possible to type or copy-and-paste the commands below one-at-a-time into the Scala REPL. We need to start with a few imports.

import smfsb._
import breeze.linalg.{Vector => BVec, _}
import breeze.numerics._
import breeze.plot._

Note that I’ve renamed Breeze’s Vector type to BVec to avoid clashing with that in the Scala standard library. We are now ready to go.

Simulating models

Let’s begin by instantiating a Lotka-Volterra model, simulating a single realisation of the process, and then plotting it.

// Simulate LV with Gillespie
val model = SpnModels.lv[IntState]()
val step = Step.gillespie(model)
val ts = Sim.ts(DenseVector(50, 100), 0.0, 20.0, 0.05, step)
Sim.plotTs(ts, "Gillespie simulation of LV model with default parameters")

The library comes with a few other models. There’s a Michaelis-Menten enzyme kinetics model:

// Simulate other models with Gillespie
val stepMM = Step.gillespie(SpnModels.mm[IntState]())
val tsMM = Sim.ts(DenseVector(301,120,0,0), 0.0, 100.0, 0.5, stepMM)
Sim.plotTs(tsMM, "Gillespie simulation of the MM model")

and an auto-regulatory genetic network model, for example.

val stepAR = Step.gillespie(SpnModels.ar[IntState]())
val tsAR = Sim.ts(DenseVector(10, 0, 0, 0, 0), 0.0, 500.0, 0.5, stepAR)
Sim.plotTs(tsAR, "Gillespie simulation of the AR model")

If you know the book and/or the R package, these models should all be familiar.
We are not restricted to exact stochastic simulation using the Gillespie algorithm. We can use an approximate Poisson time-stepping algorithm.

// Simulate LV with other algorithms
val stepPts = Step.pts(model)
val tsPts = Sim.ts(DenseVector(50, 100), 0.0, 20.0, 0.05, stepPts)
Sim.plotTs(tsPts, "Poisson time-step simulation of the LV model")

Alternatively, we can instantiate the example models using a continuous state rather than a discrete state, and then simulate using algorithms based on continous approximations, such as Euler-Maruyama simulation of a chemical Langevin equation (CLE) approximation.

val stepCle = Step.cle(SpnModels.lv[DoubleState]())
val tsCle = Sim.ts(DenseVector(50.0, 100.0), 0.0, 20.0, 0.05, stepCle)
Sim.plotTs(tsCle, "Euler-Maruyama/CLE simulation of the LV model")

If we want to ignore noise temporarily, there’s also a simple continuous deterministic Euler integrator built-in.

val stepE = Step.euler(SpnModels.lv[DoubleState]())
val tsE = Sim.ts(DenseVector(50.0, 100.0), 0.0, 20.0, 0.05, stepE)
Sim.plotTs(tsE, "Continuous-deterministic Euler simulation of the LV model")

Spatial stochastic reaction-diffusion simulation

We can do 1d reaction-diffusion simulation with something like:

val N = 50; val T = 40.0
val model = SpnModels.lv[IntState]()
val step = Spatial.gillespie1d(model,DenseVector(0.8, 0.8))
val x00 = DenseVector(0, 0)
val x0 = DenseVector(50, 100)
val xx00 = Vector.fill(N)(x00)
val xx0 = xx00.updated(N/2,x0)
val output = Sim.ts(xx0, 0.0, T, 0.2, step)
Spatial.plotTs1d(output)

For 2d simulation, we use PMatrix, a comonadic matrix/image type defined within the library, with parallelised map and coflatMap (cobind) operations. See my post on comonads for scientific computing for further details on the concepts underpinning this, though note that it isn’t necessary to understand comonads to use the library.

val r = 20; val c = 30
val model = SpnModels.lv[DoubleState]()
val step = Spatial.cle2d(model, DenseVector(0.6, 0.6), 0.05)
val x00 = DenseVector(0.0, 0.0)
val x0 = DenseVector(50.0, 100.0)
val xx00 = PMatrix(r, c, Vector.fill(r*c)(x00))
val xx0 = xx00.updated(c/2, r/2, x0)
val output = step(xx0, 0.0, 8.0)
val f = Figure("2d LV reaction-diffusion simulation")
val p0 = f.subplot(2, 1, 0)
p0 += image(PMatrix.toBDM(output map (_.data(0))))
val p1 = f.subplot(2, 1, 1)
p1 += image(PMatrix.toBDM(output map (_.data(1))))

Bayesian parameter inference

The library also includes functions for carrying out parameter inference for stochastic dynamical systems models, using particle MCMC, ABC and ABC-SMC. See the examples directory for further details.

Next steps

Having worked through this post, the next step is to work through the tutorial. There is some overlap of content with this blog post, but the tutorial goes into more detail regarding the basics. It also finishes with suggestions for how to proceed further.

Source

This post started out as a tut document (the Scala equivalent of an RMarkdown document). The source can be found here.

Monadic probabilistic programming in Scala with Rainier

Introduction

Rainier is an interesting new probabilistic programming library for Scala recently open-sourced by Stripe. Probabilistic programming languages provide a computational framework for building and fitting Bayesian models to data. There are many interesting probabilistic programming languages, and there is currently a lot of interesting innovation happening with probabilistic programming languages embedded in strongly typed functional programming languages such as Scala and Haskell. However, most such languages tend to be developed by people lacking expertise in statistics and numerics, leading to elegant, composable languages which work well for toy problems, but don’t scale well to the kinds of practical problems that applied statisticians are interested in. Conversely, there are a few well-known probabilistic programming languages developed by and for statisticians which have efficient inference engines, but are hampered by inflexible, inelegant languages and APIs. Rainier is interesting because it is an attempt to bridge the gap between these two worlds: it has a functional, composable, extensible, monadic API, yet is backed by a very efficient, high-performance scalable inference engine, using HMC and a static compute graph for reverse-mode AD. Clearly there will be some loss of generality associated with choosing an efficient inference algorithm (eg. for HMC, there needs to be a fixed number of parameters and they must all be continuous), but it still covers a large proportion of the class of hierarchical models commonly used in applied statistical modelling.

In this post I’ll give a quick introduction to Rainier using an interactive session requiring only that SBT is installed and the Rainier repo is downloaded or cloned.

Interactive session

To follow along with this post just clone, or download and unpack, the Rainier repo, and run SBT from the top-level Rainier directory and paste commands. First start a Scala REPL.

project rainierPlot
console

Before we start building models, we need some data. For this post we will focus on a simple logistic regression model, and so we will begin by simulating some synthetic data consistent with such a model.

val r = new scala.util.Random(0)
val N = 1000
val beta0 = 0.1
val beta1 = 0.3
val x = (1 to N) map { i =>
  3.0 * r.nextGaussian
}
val theta = x map { xi =>
  beta0 + beta1 * xi
}
def expit(x: Double): Double = 1.0 / (1.0 + math.exp(-x))
val p = theta map expit
val y = p map (pi => (r.nextDouble < pi))

Now we have some synthetic data, we can fit the model and see if we are able to recover the “true” parameters used to generate the synthetic data. In Rainier, we build models by declaring probabilistic programs for the model and the data, and then run an inference engine to generate samples from the posterior distribution.

Start with a bunch of Rainier imports:

import com.stripe.rainier.compute._
import com.stripe.rainier.core._
import com.stripe.rainier.sampler._
import com.stripe.rainier.repl._

Now we want to build a model. We do so by describing the joint distribution of parameters and data. Rainier has a few built-in distributions, and these can be combined using standard functional monadic combinators such as map, zip, flatMap, etc., to create a probabilistic program representing a probability monad for the model. Due to the monadic nature of such probabilistic programs, it is often most natural to declare them using a for-expression.

val model = for {
  beta0 <- Normal(0, 5).param
  beta1 <- Normal(0, 5).param
  _ <- Predictor.from{x: Double =>
      {
        val theta = beta0 + beta1 * x
        val p = Real(1.0) / (Real(1.0) + (Real(0.0) - theta).exp)
        Categorical.boolean(p)
      }
    }.fit(x zip y)
} yield Map("b0"->beta0, "b1"->beta1)

This kind of construction is very natural for anyone familiar with monadic programming in Scala, but will no doubt be a little mysterious otherwise. RandomVariable is the probability monad used for HMC sampling, and these can be constructed from Distributions using .param (for unobserved parameters) and .fit (for variables with associated observations). Predictor is just a convenience for observations corresponding to covariate information. model is therefore a RandomVariable over beta0 and beta1, the two unobserved parameters of interest. Note that I briefly discussed this kind of pure functional approach to describing probabilistic programs (using Rand from Breeze) in my post on MCMC as a stream.

Now we have our probabilistic program, we can sample from it using HMC as follows.

implicit val rng = ScalaRNG(3)
val its = 10000
val thin = 5
val out = model.sample(HMC(5), 10000, its*thin, thin)
println(out.take(10))

The argument to HMC() is the number of leapfrog steps to take per iteration.

Finally, we can use EvilPlot to look at the HMC output and check that we have managed to reasonably recover the true parameters associated with our synthetic data.

import com.cibo.evilplot.geometry.Extent
import com.stripe.rainier.plot.EvilTracePlot._

render(traces(out, truth = Map("b0" -> beta0, "b1" -> beta1)),
  "traceplots.png", Extent(1200, 1000))
render(pairs(out, truth = Map("b0" -> beta0, "b1" -> beta1)), "pairs.png")

Everything looks good, and the sampling is very fast!

Further reading

For further information, see the Rainier repo. In particular, start with the tour of Rainier’s core, which gives a more detailed introduction to how Rainier works than this post. Those interested in how the efficient AD works may want to read about the compute graph, and the implementation notes explain how it all fits together. There is some basic ScalaDoc for the core package, and also some examples (including this one), and there’s a gitter channel for asking questions. This is a very new project, so there are a few minor bugs and wrinkles in the initial release, but development is progressing rapidly, so I fully expect the library to get properly battle-hardened over the next few months.

For those unfamiliar with the monadic approach to probabilistic programming, then Ścibior et al (2015) is probably a good starting point.