A Brief Introduction to Learning Bayesian Networks

This page is based on the following paper, which focusses on learning the structure of DBNs, using biosequence data as an example application domain. For a more thorough treatment of learning, see

The learning problem can be summarized as follows.

Structure   Observability    Method
---------------------------------------------
Known       full             Sample statistics 
Known       partial          EM or gradient ascent
Unknown     full             Search through model space 
Unknown     partial          Structural EM
Full observability means that the values of all variables are known; partial observability means that we do not know the values of some of the variables. This might be because they cannot, in principle, be measured (in which case they are usually called hidden variables), or because they just happen to be unmeasured in the training data (in which case they are called missing variables). Note that a variable can be observed intermittently.

Unknown structure means we do not know the complete topology of the graph. Typically we will know some parts of it, or at least know some properties the graph is likely to have, e.g., it is common to assume a bound, K, on the maximum number of parents (fan-in) that a node can take on, and we may know that nodes of a certain ``type'' only connect to other nodes of the same type. Such prior knowledge can either be a ``hard'' constraint, or a ``soft'' prior.

To exploit such soft prior knowledge, we must use Bayesian methods, which have the additional advantage that they return a distribution over possible models instead of a single best model. Handling priors on model structures, however, is quite complicated, so we do not discuss this issue here (see \cite{Heckerman96} for a review). Instead, we assume that the goal is to find a single model which maximizes some scoring function (discussed in Section~\ref{sec:cost_fn}). We will, however, consider priors on {\em parameters}, which are much simpler. In Section~\ref{sec:learn_cts}, when we consider DBNs in which all the variables are continuous, we will see that we can use numerical priors as a proxy for structural priors.

An interesting approximation to the full-blown Bayesian approach is to learn a mixture of models, each of which has different structure, depending on the value of a hidden, discrete, mixture node. This has been done for Gaussian networks \cite{Thiesson98} and discrete, undirected trees \cite{Meila98}. Unfortunately, there are severe computational difficulties in generalizing this technique to arbitrary, discrete networks.

In this paper, we restrict our attention to learning the structure of the inter-slice connectivity matrix of the DBN. This has the advantage that we do not need to worry about introducing (directed) cycles, and also that we can use the ``arrow of time'' to disambiguate the direction of causal relations, i.e., we know an arc must go from X_{t-1} to X_t and not vice versa, because the network models the temporal evolution of a physical system. Of course, we still cannot completely overcome the problem that ``correlation is not causation'': it is possible that the correlation between X_{t-1} and X_t is due to one or more hidden common causes (or observed common effects, because of the explaining away phenomenon). Hence the models learned from data should be subject to experimental verification. (\cite{Heckerman97} discusses Bayesian techniques for learning causal (static) networks, and \cite{Spirtes93} discusses constraint based techniques --- but see \cite{Freeman97} for a critique of this approach. These techniques have mostly been used in the social sciences, where experimental verification is difficult.)

When the structure of the model is known, the learning task becomes one of parameter estimation. Although the focus of this paper is on learning structure, this calls parameter learning as a subroutine, so we start by discussing this case.

Finally, we mention that learning BNs is a huge subject, and in this paper, we only touch on the aspects that we consider most relevant to learning genetic networks from time series data. For further details, see the review papers \cite{Buntine94,Buntine96,Heckerman96}.

Known structure, full observability

We assume that the goal of learning in this case is to find the values of the parameters of each CPD which maximizes the likelihood of the training data, which contains S sequences (assumed to be independent), each of which has the observed values of all n nodes per slice for each of T slices. For notational simplicity, we assume each sequence is of the same length. Thus we can imagine ``unrolling'' a two-slice DBN to produce a (static) BN with T slices.

We assume the parameter values for all nodes are constant (tied) across time (in contrast to the assumption made in HMM protein alignment), so that for a time series of length T, we get one data point (sample) for each CPD in the initial slice, and T-1 data points for each of the other CPDs. If S=1, we cannot reliably estimate the parameters of the nodes in the first slice, so we usually assume these are fixed apriori. That leaves us with N=S(T-1) samples for each of the remaining CPDs. In cases where N is small compared to the number of parameters that require fitting, we can use a numerical prior to regularize the problem (see Section~\ref{sec:param_est}). In this case, we call the estimates Maximum A Posterori (MAP) estimates, as opposed to Maximum Likelihood (ML) estimates.

Using the Bayes Ball algorithm (see Figure~\ref{fig:bayes_ball}), it is easy to see that each node is conditionally independent of its non-descendants given its parents (indeed, this is often taken as the {\em definition} of a BN), and hence, by the chain rule of probability, we find that the joint probability of all the nodes in the graph is

where m=n(T-1) is the number of nodes in the unrolled network (excluding the first slice), Pa(X_i) are the parents of node i, and nodes are numbered in topological order (i.e., parents before children). The normalized log-likelihood of the training set L = \frac{1}{N} \log \Pr(D|G), where D = \{D_1, \ldots, D_S\}, is a sum of terms, one for each node:

We see that the log-likelihood scoring function {\em decomposes} according to the structure of the graph, and hence we can maximize the contribution to the log-likelihood of each node independently (assuming the parameters in each node are independent of the other nodes).

Parameter Estimation

For discrete nodes with CPTs, we can compute the ML estimates by simple counting. For example, in the water sprinkler network, we have

where N(W=w,S=s,R=r) is the number of times this event occurs in the training set. As is well known from the HMM literature, however, ML estimates of CPTs are prone to sparse data problems, which can be solved by using (mixtures of) Dirichlet priors (pseudo counts).

Estimating the parameters of noisy-OR distributions (and its relatives) is harder, essentially because of the presence of hidden variables (see Figure~\ref{fig:noisy_or}), so we must use the techniques discussed in Section~\ref{sec:known_struct_partial_obs}.

The key advantage of BNs is that they have a small number of parameters, and hence require less data to learn. For example, we can specify the joint distribution over n nodes using only O(n * 2^k) parameters (where k is the maximum fan-in of any node), instead of the O(2^n) which would be required for an explicit representation.

Choosing the form of the CPDs

In the above discussion, we assumed that we knew the form of the CPD for each node. If we are uncertain, one approach is to use a mixture distribution, although this introduces a hidden variable, which makes things more complicated (see Section~\ref{sec:known_struct_partial_obs}). Alternatively, we can use a decision tree \cite{Boutilier96}, or a table of parent values along with their associated non-zero probabilities \cite{Friedman96}, to represent the CPD. This can increase the number of free parameters gradually, from 1 to 2^k, where k is the number of parents.

Known structure, partial observability

When some of the variables are not observed, the likelihood surface becomes multimodal, and we must use iterative methods, such as EM \cite{Lauritzen95} or gradient ascent \cite{Binder97}, to find a local maximum of the ML/MAP function. These algorithms need to use an inference algorithm to compute the expected sufficient statistics (or related quantity) for each node. For example, using EM, the above ML estimate becomes

P(W=w|S=s,R=r) = EN(W=w,S=s,R=r) / EN(S=s,R=r)

where EN(.) is the expected number of times event . occurs in the whole training set. These expected counts can be computed as follows

EN(.) = E sum_l I(.|D(l)) = sum_l P(. | D(l))

where I(.|D(l)) is an indicator function which is 1 if event . occurs in training case l, and 0 otherwise.

Unknown structure, full observability

We start by discussing the scoring function which we use to select models; we then discuss algorithms which attempt to optimize this function over the space of models, and finally examine their computational and sample complexity.

The objective function used for model selection

It is commonly assumed that the goal is to find the model with maximum likelihood. Often (e.g., as in the REVEAL algorithm \cite{Liang98}, and as in the Reconstructability Analysis (RA) or General Systems Theory community \cite{Klir85,Krippendorff86}) this is stated as finding the model in which the sum of the mutual information (MI) \cite{Cover91} between each node and its parents is maximal; in Appendix~\ref{app:MI}, we prove that these objective functions are equivalent, in the sense that they rank models in the same order.

The trouble is, the ML model will be a complete graph, since this has the largest number of parameters, and hence can fit the data the best. A well-principled way to avoid this kind of over-fitting is to put a prior on models, specifying that we prefer sparse models. Then, by Bayes' rule, the MAP model is the one that maximizes

Taking logs, we find
where c = - \log \Pr(D) is a constant independent of G. Thus the effect of the prior is equivalent to penalizing overly complex models, as in the Minimum Description Length (MDL) approach.

An exact Bayesian approach to model selection is usually unfeasible, since it involves computing the marginal likelihood P(D) = \sum_G P(D,G), which is a sum over an exponential number of models (see Section~\ref{sec:algo}). However, we can use an asymptotic approximation to the posterior called BIC (Bayesian Information Criterion), which is defined as follows:

where N is the number of samples, #G is the dimension of the model (In the fully observable case, the dimension of a model is the number of free parameters. In a model with hidden variables, it might be less than this \cite{Geiger96}), and \hat{\Theta}_G is the ML estimate of the parameters. The first term is computed using Equation~\ref{eqn:loglik}. The second term is a penalty for model complexity. Since the number of parameters in the model is the sum of the number of parameters in each node\footnote{ Hence nodes with compact representations of their CPDs will encur a lower penalty, which can allow connections to form which might otherwise have been rejected \cite{Friedman96}.}, we see that the BIC score decomposes just like the log-likelihood (Equation~\ref{eqn:loglik}). Hence all of the techniques invented for maximizing MI also work for the BIC case, and indeed, any decomposable scoring function.

Algorithms for finding the best model

Given that the score is decomposable, we can learn the parent set for each node independently. There are
\sum_{k=0}^n \choice{n}{k} = 2^n
such sets, which can be arranged in a lattice as shown in Figure~\ref{fig:lattice}. The problem is to find the highest scoring point in this lattice.

The approach taken by REVEAL \cite{Liang98} is to start at the bottom of the lattice, and evaluate the score at all points in each successive level, until a point is found with a score of 1.0. Since the scoring function they use is I(X,Pa(X))/\max\{H(X),H(Pa(X))\}, where I(X,Y) is the mutual information between X and Y, and H(X) is the entropy of X (see Appendix~\ref{app:MI} for definitions), 1.0 is the highest possible score, and corresponds to Pa(X) being a perfect predictor of X, i.e., H(X|Pa(X))=0.

A normalized MI of 1 is only achievable with a deterministic relationship. For stochastic relationships, we must decide whether the gains in MI produced by a larger parent set is ``worth it''. The standard approach in the RA community uses the fact \cite{Miller55} that \chi^2(X,Y) \approx I(X,Y) N \ln(4), where N is the number of samples. Hence we can use a \chi^2 test to decide whether an increase in MI is statistically significant. (This also gives us some kind of confidence measure on the connections that we learn.) Alternatively, we can use a complexity penalty term, as in the BIC score.

Of course, if we do not know if we have achieved the maximum possible score, we do not know when to stop searching, and hence we must evaluate all points in the lattice (although we can obviously use branch-and-bound). For large n, this is computationally infeasible, so a common approach is to only search up until level K (i.e., assume a bound on the maximum number of parents of each node), which takes O(n ^ K) time. %For example, in \cite{Liang98}, they use K=3 and n=50; %\choice{50}{3} = 19,600, a modest number. Unfortunately, in real genetic networks, it is known that some genes can have very high fan-in (and fan-out), so restricting the bound to K=3 would make it impossible to discover these important ``master'' genes.

The obvious way to avoid the exponential cost (and the need for a bound, K) is to use heuristics to avoid examining all possible subsets. (In fact, we must use heuristics of some kind, since the problem of learning optimal structure is NP-hard \cite{Chickering95}.) One approach in the RA framework, called Extended Dependency Analysis (EDA) \cite{Conant88}, is as follows. Start by evaluating all subsets of size up to two, keep all the ones with significant (in the \chi^2 sense) MI with the target node, and take the union of the resulting set as the set of parents.

The disadvantage of this greedy technique is that it will fail to find a set of parents unless some subset of size two has significant MI with the target variable. However, a Monte Carlo simulation in \cite{Conant88} shows that most random relations have this property. In addition, highly interdependent sets of parents (which might fail the pairwise MI test) violate the causal independence assumption, which is necessary to justify the use of noisy-OR and similar CPDs.

An alternative technique, popular in the UAI community, is to start with an initial guess of the model structure (i.e., at a specific point in the lattice), and then perform local search, i.e., evaluate the score of neighboring points in the lattice, and move to the best such point, until we reach a local optimum. We can use multiple restarts to try to find the global optimum, and to learn an ensemble of models. Note that, in the partially observable case, we need to have an initial guess of the model structure in order to estimate the values of the hidden nodes, and hence the (expected) score of each model (see Section~\ref{sec:known_struct_partial_obs}); starting with the fully disconnected model (i.e., at the bottom of the lattice) would be a bad idea, since it would lead to a poor estimate.

Sample complexity

In addition to the computational cost, another important consideration is the amount of data that is required to reliably learn structure. For deterministic boolean networks, this issue has been addressed from a statistical physics perspective \cite{Hertz98} and a computational learning theory (combinatorial) perspective \cite{Akutsu99}. In particular, \cite{Akutsu99} prove that, if the fan-in is bounded by a constant K, the number of samples needed to identify a boolean network of n nodes is lower bounded by \Omega(2^K + K \log n) and upper bounded by O(2^{2K} 2K \log n). Unfortunately, the constant factor is exponential in K, which can be quite large for certain genes, as we have already remarked. These analyses all assume a ``tabula rasa'' approach; however, in practice, we will often have strong prior knowledge about the structure (or at least parts of it), which can reduce the data requirements considerably.

Unknown structure, partial observability

Finally, we come to the hardest case of all, where the structure is unknown and there are hidden variables and/or missing data. One difficulty is that the score does not decompose. However, as observed in \cite{Friedman97,Friedman98b,Friedman98,Thiesson98}, the {\em expected} score {\em does} decompose, and so we can use an iterative method, which alternates between evaluating the expected score of a model (using an inference engine), and changing the model structure, until a local maximum is reached. This is called the Structural EM (SEM) algorithm. SEM was succesfully used to learn the structure of discrete DBNs within missing data in \cite{Friedman98}.

Inventing new hidden nodes

So far, structure learning has meant finding the right connectivity between pre-existing nodes. A more interesting problem is inventing hidden nodes on demand. Hidden nodes can make a model much more compact, as we see below.
(a) A BN with a hidden variable H. (b) The simplest network that can capture the same distribution without using a hidden variable (created using arc reversal and node elimination). If H is binary and the other nodes are trinary, and we assume full CPTs, the first network has 45 independent parameters, and the second has 708.

The standard approach is to keep adding hidden nodes one at a time, to some part of the network (see below), performing structure learning at each step, until the score drops. One problem is choosing the cardinality (number of possible values) for the hidden node, and its type of CPD. Another problem is choosing where to add the new hidden node. There is no point making it a child, since hidden children can always be marginalized away, so we need to find an existing node which needs a new parent, when the current set of possible parents is not adequate.

\cite{Ramachandran98} use the following heuristic for finding nodes which need new parents: they consider a noisy-OR node which is nearly always on, even if its non-leak parents are off, as an indicator that there is a missing parent. Generalizing this technique beyond noisy-ORs is an interesting open problem. One approach might be to examine H(X|Pa(X)): if this is very high, it means the current set of parents are inadequate to ``explain'' the residual entropy; if Pa(X) is the best (in the BIC or \chi^2 sense) set of parents we have been able to find in the current model, it suggests we need to create a new node and add it to Pa(X).

A simple heuristic for inventing hidden nodes in the case of DBNs is to check if the Markov property is being violated for any particular node. If so, it suggests that we need connections to slices further back in time. Equivalently, we can add new lag variables (see Section~\ref{sec:higher_order}) and connect to them. (Note that reasoning across multiple time slices forms the basis of the (``model free'') correlational analysis approach of \cite{Arkin95,Arkin97}.)

Another heuristic for DBNs might be to first perform a clustering of the timeseries of the observed variables (see e.g., \cite{Eisen98}), and then to associate hidden nodes with each cluster. The result would be a Markov model with a tree-structured hidden ``backbone'' c.f., \cite{Jordan96}. This is one possible approach to the problem of learning hierarchically structured DBNs. (Building hierarchical (object oriented) BNs and DBNs {\em by hand} is straightforward, and there are algorithms which can exploit this modular structure to a certain extent to speed up inference \cite{Koller97,Friedman98_doobn}.)

Of course, interpreting the ``meaning'' of hidden nodes is always tricky, especially since they are often unidentifiable, e.g., we can often switch the interpretation of the true and false states (assuming for simplicity that the hidden node is binary) provided we also permute the parameters appropriately. (Symmetries such as this are one cause of the multiple maxima in the likelihood surface.) Our opinion is that fully automated structure discovery techniques can be useful as hypothesis generators, which can then be tested by experiment.