Fact-checked by Grok 2 weeks ago

Conditional probability table

A conditional probability table (CPT) is a compact tabular structure in Bayesian networks that encodes the of a given the states of its parent variables in the network's . This representation allows for the factorization of the over multiple variables into a product of local conditional probabilities, leveraging conditional independencies to reduce . CPTs are essential for , enabling the calculation of probabilities for unobserved variables based on evidence from observed ones. The structure of a CPT for a node X with n parent nodes consists of rows enumerating all possible combinations of parent states and columns listing the probabilities of each possible state of X, ensuring that probabilities in each row sum to 1. For binary variables, a CPT might have $2^n rows and 3 columns: one for each parent configuration, one for X's state, and one for the probability value. In practice, CPTs can grow exponentially with the number of parents—a challenge known as the "parameter explosion" problem—prompting techniques like parameter learning from data or expert elicitation to populate them accurately. CPTs play a central role in applications of Bayesian networks across fields such as , decision support systems, and , where they facilitate belief updating and scenario analysis under uncertainty. For instance, in a network modeling alarm triggers from and events, the CPT for the alarm node would specify probabilities like P(\text{Alarm} \mid \text{Burglary}, \text{Earthquake}). Their discrete nature suits categorical variables, though extensions exist for continuous distributions via other mechanisms in probabilistic graphical models.

Fundamentals

Definition and Basic Concepts

A conditional probability table (CPT) is a table that enumerates the conditional probability mass function (or density) of a given the states of its variables in a . This representation captures the local dependencies between variables, allowing the joint probability distribution over all variables to be factored efficiently according to the model's structure. In standard notation, a CPT for a child variable X with parents Y specifies the probabilities P(X \mid Y = y) for each possible value of X and each combination y of states. The key components of a CPT include rows corresponding to the exhaustive combinations of states from the variables, and columns for the possible values (or discretized intervals) of the X. Each entry in the table contains the associated , ensuring that the values in each row sum to 1 to reflect a valid over X for that fixed configuration. This tabular format facilitates both specification and in models where variables are or where continuous densities are approximated tabularly. For nodes without parents ( nodes), the CPT reduces to the marginal of the . CPTs were formalized in the context of Bayesian networks by Judea Pearl in the 1980s, providing a structured way to quantify dependencies in directed acyclic graphs for probabilistic reasoning under uncertainty.

Relation to Conditional Probability

A conditional probability table (CPT) serves as a direct tabular representation of conditional probabilities, which quantify the likelihood of one event occurring given the occurrence of another. In probability theory, events are classified as independent if the occurrence of one does not affect the probability of the other, such that P(A \cap B) = P(A) P(B), or dependent otherwise, necessitating conditioning to capture interdependencies. Conditioning becomes essential for dependent events, as it adjusts probabilities based on observed evidence, enabling more precise modeling of real-world uncertainties. The foundational formula for conditional probability is P(X = x \mid Y = y) = \frac{P(X = x, Y = y)}{P(Y = y)}, where P(X = x, Y = y) is the joint probability and P(Y = y) is the marginal probability of Y, assuming P(Y = y) > 0. A CPT tabulates these values for each possible state of the conditioning variables, storing P(X = x \mid Y = y) entries that can be referenced directly without recomputing the ratio from joint probabilities each time. This avoids redundant calculations of joint distributions, which grow exponentially with the number of variables, making CPTs efficient for structured probabilistic reasoning. Unlike a probability , which enumerates all combinations of P(X, Y) across the full , a CPT emphasizes conditional dependencies by fixing the parents and listing outcomes for the child , facilitating the chain rule P(X, Y) = P(X \mid Y) P(Y). This decomposition exploits conditional independencies to represent high-dimensional compactly, a core principle in graphical models where the distribution factors as a product of local conditionals. Bayes' theorem, derived from the conditional probability definition, provides a mechanism for inverting dependencies: starting from P(A \mid B) = \frac{P(A \cap B)}{P(B)} and P(B \mid A) = \frac{P(A \cap B)}{P(A)}, equate the numerators to yield P(A \mid B) = \frac{P(B \mid A) P(A)}{P(B)}, where P(B) = P(B \mid A) P(A) + P(B \mid A^c) P(A^c) normalizes over all possibilities. In the context of a CPT, this theorem allows updating beliefs by combining prior probabilities from one table with likelihoods from another, enabling probabilistic without full joint enumeration.

Representation and Construction

For Discrete Random Variables

In Bayesian networks, (CPTs) for discrete random variables are constructed by enumerating all possible combinations of the variables' states and specifying the conditional probabilities for each state of the variable given those combinations. For a variable X with |X| possible states and k variables (each with 2 states), there are $2^k combinations, resulting in a with |X| rows (one for each state of X) and $2^k columns (one for each combination). Probabilities are assigned to each entry such that, for each fixed combination (column), the values across the states (rows) sum to 1, ensuring the table represents a valid . For example, consider a child with 2 states and two parents; the CPT would feature 2 rows and 4 columns, with entries like P(X = x_1 | Pa_1 = 0, Pa_2 = 0), P(X = x_1 | Pa_1 = 0, Pa_2 = 1), and so on, where each column's entries sum to 1. These probabilities can be estimated from data using (MLE), which involves counting the frequency of each child-parent configuration in the observed data and normalizing those counts within each column to obtain the conditional probabilities. Normalization is a core step in CPT construction, guaranteeing that the probabilities for all states of the child variable sum to 1 for every parent configuration; this is achieved by dividing the count of each child state by the total count for that parent combination in the MLE process. When data is incomplete (e.g., some variables unobserved), MLE for CPT parameters is performed via the expectation-maximization (EM) algorithm, where the E-step infers expected counts using current parameters and the M-step updates the table by normalizing those expected counts, iteratively maximizing the likelihood. The of representing a CPT arises from its size, which grows with the number of parents: for a child with |X| states and parents Y_1, \dots, Y_k having cardinalities |Y_i|, the table requires O(|X| \prod_{i=1}^k |Y_i|) entries to store. This exponential growth limits the practicality of exact CPTs for nodes with many parents, motivating techniques like parameter tying or approximations in large models.

For Continuous Random Variables

In the case of continuous random variables, conditional probability tables are not feasible due to the infinite support of the variables, and are instead represented by conditional probability density functions (PDFs) that depend on the values of the parent variables. The conditional PDF of a continuous random variable X given its parents Y = y is formally defined as f_{X|Y}(x|y) = \frac{f_{X,Y}(x,y)}{f_Y(y)}, where f_{X,Y} is the joint PDF and f_Y is the marginal PDF of Y, assuming f_Y(y) > 0. This parameterization allows the distribution to be specified functionally rather than through enumeration, enabling the modeling of dependencies in probabilistic graphical models like Bayesian networks. A common parameterized form is the conditional linear Gaussian distribution, where X given Y = y follows a X \mid Y = y \sim \mathcal{N}(\mu(y), \sigma^2), with the \mu(y) expressed as a of the parents, such as \mu(y) = \beta_0 + \sum \beta_i y_i, and the variance \sigma^2 often held constant or also dependent on y. This approach, known as a linear Gaussian , facilitates exact in hybrid Bayesian networks under certain structural constraints, such as no discrete nodes with continuous parents. More flexible representations include mixtures of truncated exponentials (MTE), which approximate arbitrary conditional densities using weighted sums of truncated exponential functions to capture or non-Gaussian behaviors without . Construction of these conditional PDFs typically involves specifying a functional form based on or data, such as to estimate the mean parameters \mu(y) from observed parent-child relationships, while kernel methods like Gaussian can be used to nonparametrically approximate the density shape. For instance, in a Gaussian CPT, the parameters \mu(y) and \sigma(y) are fitted via maximum likelihood or Bayesian methods, ensuring the conditional density integrates to unity over the support of X. Unlike discrete cases, this avoids exhaustive tabulation but requires careful selection of the parametric family to match the underlying data-generating process. One key challenge in using these parameterized conditional PDFs for exact is the computational complexity of integrating over continuous domains, often leading to the of variables as a practical to enable tabular approximations or numerical methods like . This , while simplifying computations, can introduce approximation errors, particularly for variables with wide ranges or nonlinear dependencies.

Applications in Probabilistic Models

In Bayesian Networks

In Bayesian networks, which are directed acyclic graphs (DAGs) that compactly represent s over s, each corresponds to a and is quantified by a conditional probability table (CPT) specifying its conditional distribution given the values of its parent s. The parents of a , along with its children and children's other parents, form its , but the CPT is defined solely based on the direct parents to capture local conditional dependencies. This setup enables the factorization of the full as P(X_1, \dots, X_n) = \prod_{i=1}^n P(X_i \mid \text{Parents}(X_i)), where each factor is directly obtained from the node's CPT, allowing efficient storage and computation relative to the full joint table. The use of CPTs in Bayesian networks facilitates a separation between qualitative modeling—defined by the DAG's structure, which encodes conditional independencies via d-separation—and quantitative modeling, provided by the CPTs that assign precise probabilistic strengths to the dependencies. This dichotomy supports modular construction, where the graph can be elicited from domain expertise and CPTs populated via statistical estimation or expert judgment. Learning CPTs in Bayesian s typically involves parameter estimation from observational data, assuming a known network structure. For complete data, computes the CPT entries as normalized frequencies of observed configurations of each node and its parents. With incomplete datasets, where some variable values are , the expectation-maximization () algorithm iteratively refines the parameters: the E-step computes expected sufficient statistics using current CPTs and current marginals over , while the M-step updates the CPTs via on those expectations. Inference in Bayesian networks, which computes posterior probabilities or marginals given evidence, leverages CPTs through algorithms that exploit the . Variable elimination performs inference by selecting an elimination order for non-query variables, successively summing out each one to form intermediate factors derived from multiplying and marginalizing relevant CPTs. , applicable to (singly connected) networks, propagates messages between nodes by combining evidence from subtrees using CPT lookups and local computations. These methods ensure that CPTs serve as the core quantitative primitives for propagating probabilities across the network.

In Other Graphical Models

In Markov random fields (MRFs), conditional probability tables (CPTs) are not directly used as in directed graphical models; instead, the joint distribution is parameterized by potential functions associated with maximal cliques in the undirected graph. These potential functions, often denoted as \psi_c for clique c, are nonnegative real-valued functions that encode interactions between variables without requiring normalization to probabilities, and the joint probability is given by P(\mathbf{X}) = \frac{1}{Z} \prod_c \psi_c(\mathbf{X}_c), where Z is the partition function. This approach contrasts with CPTs by focusing on unnormalized factors rather than explicit conditional probabilities, enabling representation of symmetric dependencies. However, CPTs can be derived from MRF potentials as normalized conditionals over the relevant cliques, providing a bridge to conditional inference. Conditional random fields (CRFs), an extension of MRFs for discriminative modeling, similarly employ potential functions over cliques to define the P(\mathbf{Y} | \mathbf{X}), where \mathbf{Y} are output variables and \mathbf{X} are observed inputs. The potentials in CRFs capture feature-based interactions, and the factorizes as P(\mathbf{Y} | \mathbf{X}) = \frac{1}{Z(\mathbf{X})} \prod_c \psi_c(\mathbf{Y}_c, \mathbf{X}_c), allowing CPT-like conditionals to be computed post- for tasks such as sequence labeling. The Hammersley-Clifford theorem underpins this framework by proving that, for strictly positive distributions satisfying the global Markov properties, the (or conditional) distribution factorizes into potentials over the cliques of the undirected , from which equivalent CPTs can be obtained via marginalization and . In hybrid models like dynamic Bayesian networks (DBNs), which combine directed acyclic structures with temporal extensions, CPTs are adapted for time-sliced representations to model sequences of variables. Each time slice contains nodes with CPTs defining intra-slice conditionals, while inter-slice CPTs capture transitions between consecutive slices, enabling efficient parameterization of evolving states without full unrolling for long sequences. This time-sliced use of CPTs facilitates applications in time-series modeling, such as or system monitoring. CPTs also play a key role in naive Bayes classifiers, a simplified variant where a central class node directly parents a set of conditionally independent feature nodes, with CPTs specifying P(\text{feature} | \text{class}) for each feature. This structure assumes feature independence given the class, allowing the joint posterior to factorize simply as P(\text{class} | \text{features}) \propto P(\text{class}) \prod P(\text{feature}_i | \text{class}), making it computationally efficient for tasks. As a directed counterpart to undirected graphical models, this highlights CPTs' versatility across dependency types.

Examples and Illustrations

Binary Variable Example

A simple yet illustrative example of a conditional probability table (CPT) involves two random variables: (R), indicating whether it is raining (states: yes or no), and Sprinkler (S), indicating whether the sprinkler is on (states: yes or no). In this scenario, the activation of the sprinkler depends on the rain condition, as rain typically suppresses the need for manual watering. The CPT for P(S \mid R) captures this dependency by specifying the over S for each state of R. The CPT is represented as follows:
RP(S = \text{yes} \mid R)P(S = \text{no} \mid R)
yes0.10.9
no0.80.2
Each row sums to 1, ensuring a valid probability distribution for S given the corresponding state of R. These entries quantify the influence of rain on sprinkler use: when it is raining (R = yes), the probability of the sprinkler being on drops to 0.1, reflecting suppression due to natural watering, whereas when it is not raining (R = no), the probability rises to 0.8, indicating higher likelihood of activation for irrigation. This table allows for probabilistic reasoning, such as computing the joint probability P(S = \text{yes}, R = \text{yes}) = P(S = \text{yes} \mid R = \text{yes}) \cdot P(R = \text{yes}). The probabilities in the CPT can be derived from frequencies in hypothetical observational data. For instance, suppose a dataset records weather and sprinkler usage over 1,000 days, with 300 rainy days (R = yes) during which the sprinkler is on 30 times, yielding P(S = \text{yes} \mid R = \text{yes}) = 30/300 = 0.1; on the remaining 700 non-rainy days, the sprinkler is on 560 times, yielding P(S = \text{yes} \mid R = \text{no}) = 560/700 = 0.8. The complementary probabilities follow as 1 minus these values. Such frequency-based estimation via maximum likelihood is a standard method for constructing CPTs from empirical data in Bayesian networks.

Multi-Parent Variable Example

In Bayesian networks, a multi-parent conditional probability table (CPT) arises when a child variable depends on multiple parent variables, requiring probabilities to be specified for all combinations of parent states. Consider an illustrative scenario modeling cancer risk, where the binary variable C (cancer: yes or no) has two binary parents: S (: yes or no) and P (: high or low). This results in 4 parent combinations, and the CPT provides the conditional probabilities P(C | S, P) for each, capturing joint dependencies. The full CPT is presented below, with probabilities summing to 1 across the rows for each parent combination: | S | P | P(C = yes | S, P) | P(C = no | S, P) | |-------|---------|---------------------|----------------------| | no | low | 0.001 | 0.999 | | yes | low | 0.030 | 0.970 | | no | high | 0.020 | 0.980 | | yes | high | 0.050 | 0.950 | These illustrative values can be obtained via applied to counts in hypothetical datasets, such as tallying occurrences in a of 1000 individuals for each parent configuration. This CPT illustrates interaction effects, where the cancer risk is highest (0.050) only when both parents are present ( and high ), compared to lower risks from either factor alone (0.030 or 0.020) or neither (0.001), demonstrating how multi-parent tables model synergistic influences beyond additive effects. With two binary parents, the table requires 8 entries (4 combinations × 2 child states, minus 4 for ), but scalability issues emerge as the number of parents increases, leading to an in table size (2^n combinations for n binary parents) that complicates specification and storage in larger networks.

Extensions and Limitations

Handling Large Tables

The curse of dimensionality in tables (CPTs) manifests as in the number of required parameters as the number of parent variables increases, making specification, , and increasingly burdensome for complex models. For a child variable with k -valued parents, the full CPT contains $2^{k+1} entries if the child is also , as each combination of parent states must map to probabilities over the child's states. For instance, a with 10 parents demands 2048 entries, illustrating how even moderate leads to rapid escalation in table size. This issue is particularly acute in Bayesian networks where nodes may have multiple parents representing interrelated factors. To mitigate the parameter explosion, parametric forms that impose structural assumptions on the CPT are commonly used, such as the Noisy-OR gate, which models the as the result of independent causal influences from parents, often in scenarios of inhibition or activation. Under the Noisy-OR assumption, the probability that the is false given active parents is the product of individual inhibition probabilities, reducing the parameterization from exponential in k to linear (k+1 parameters for a ). This approach leverages independence assumptions to drastically cut elicitation effort while maintaining expressiveness for many causal domains. For storage of large or sparse CPTs, where many entries are zero or near-zero due to low-probability configurations, dense representations are inefficient; instead, sparse data structures like hash tables or dictionaries store only non-zero probabilities, enabling compact usage and faster access in implementations. In practice, such optimizations are vital for . In real-world Bayesian networks for , the aggregate size of CPTs across nodes can exceed millions of parameters without approximations, as networks with thousands of variables require handling vast conditional dependencies to model patient symptoms and outcomes effectively.

Approximations and Inference Methods

Exact inference in Bayesian networks often relies on the junction tree algorithm, which transforms the network into a junction tree structure and performs to compute marginal probabilities from conditional probability tables (CPTs) through repeated marginalization operations. Marginalization in this context involves summing over the entries of CPTs corresponding to irrelevant variables to obtain the desired ; for variables, this is given by the formula P(x) = \sum_{y} P(x \mid y) P(y), where the summation integrates out the conditioning variable y to yield the marginal for x. This process exploits the conditional independencies encoded in the network to avoid full joint table enumeration, enabling efficient computation for tree-structured or low-treewidth networks. However, exact inference via methods like the junction tree algorithm is NP-hard in general Bayesian networks, as the problem of computing exact marginals reduces to known NP-complete problems such as 3-SAT. This complexity arises even for singly connected networks beyond polytrees, motivating the development of approximate inference techniques that trade accuracy for scalability. Approximate inference methods include approaches, such as , which generates samples from the joint distribution defined by the CPTs and uses these to estimate marginals empirically. In , variables are iteratively resampled from their conditional distributions given the current values of others, leveraging the CPTs directly to approximate the posterior without explicit marginalization. For networks with cycles, loopy extends the exact message-passing paradigm by iterating messages around loops until convergence, providing approximate marginals from CPTs in a computationally efficient manner despite lacking guarantees of exactness. Variational inference offers another approximation framework, where mean-field methods parameterize approximate distributions over latent variables and optimize a lower bound on the log-likelihood, effectively parameterizing and adjusting CPT entries to minimize the divergence to the true posterior. These techniques are particularly useful for large-scale networks, as they convert into an solvable via gradient-based methods, though they may underestimate posterior variance compared to sampling approaches.

References

  1. [1]
    6.3 Bayesian Network Representation
    We formally define a Bayes Net as consisting of: 1.A directed acyclic ... conditional probability table or CPT. Each CPT has n+2 n + 2 columns: one ...
  2. [2]
    Bayes Nets
    Answer Query from Conditional Probability Table (CPT) · Bayes nets are a subset of probabilistic graphical models · We describe how variables locally interact.
  3. [3]
    13.5: Bayesian Network Theory - Engineering LibreTexts
    Mar 11, 2023 · Node C from the sample DAG above would have a conditional probability table specifying the conditional distribution P(C|A,B). Since A and B have ...Introduction · Advantages and Limitations of... · Dynamic Bayesian Networks
  4. [4]
    Quantifying conditional probability tables in Bayesian networks
    This paper proposes a 'structured' elicitation approach [51] to quantify – with uncertainty – the conditional probability table (CPT) entries that define a ...
  5. [5]
    PMML 4.4 - Bayesian Network Models - Data Mining Group
    When a node is conditioned upon its parent nodes in the network, its value is given by a conditional probability table that specifies the value probabilities ...
  6. [6]
    [PDF] 15-780: Probabilistic Graphical Models
    A directed acyclic graph (DAG) G = (V = {X1,..., Xn }, E). 2. A set of conditional probability tables p(Xi |Parents(Xi )). Defines the joint probability ...
  7. [7]
    [PDF] Bayesian Networks - UBC Computer Science
    There are 3 conditional probability tables (CPDs) to be determined: P(J | A), P(M | A), P(A | E, B). ○. Requiring 2 + 2 + 4 = 8 probabilities. And 2 ...
  8. [8]
    Conditional Probability | Formulas | Calculation | Chain Rule
    If A and B are two events in a sample space S, then the conditional probability of A given B is defined as P(A|B)= P(A∩B) P(B) , when P(B)>0.
  9. [9]
    [PDF] Ba esian Networ s Judea Pearl Cognitive Systems - UCLA
    The strength of an influence is represented by conditional probabilities that are attached to each cluster of parents-child nodes in the network. Figure 1 ...
  10. [10]
    Probabilistic Reasoning in Intelligent Systems - ScienceDirect.com
    Probabilistic Reasoning in Intelligent Systems. Networks of Plausible Inference. Book • 1988. Author: Judea Pearl ... PDF version. Ways of reading. No ...
  11. [11]
  12. [12]
    Bayesian Networks – BayesFusion
    Bayesian networks are acyclic directed graphs that represent factorizations of joint probability distributions.
  13. [13]
    [PDF] 6 : Learning Fully Observed Bayesian Networks 1 Introduction
    We know that there are two different estimation approaches: (i) maximum likelihood estimation (MLE), and (ii) Bayesian estimates. 3.1 Parameter estimation in ...
  14. [14]
    Conditional probability density function - StatLect
    The following is a formal definition. Definition Let X and Y be two continuous random variables. The conditional probability density function of X ...
  15. [15]
    [PDF] Inference in Hybrid Bayesian Networks Using Mixtures of Gaussians
    All continuous chance variables must have conditional linear Gaussian distributions, and discrete chance nodes cannot have continuous parents. The methods ...<|control11|><|separator|>
  16. [16]
    Modeling Conditional Distributions of Continuous Variables in ...
    The MTE (mixture of truncated exponentials) model was introduced as a general solution to the problem of specifying conditional distributions for continuous ...
  17. [17]
    [PDF] Implementation of Continuous Bayesian Networks Using ... - arXiv
    In this paper, we ex tend Bayesian network structures to compute probability density functions for continuous random variables. We make this extension by ...
  18. [18]
  19. [19]
    Learning Bayesian networks: approaches and issues
    May 12, 2011 · Learning structured Bayesian networks: combining abstraction hierarchies and tree-structured conditional probability tables. Computational ...
  20. [20]
    [PDF] Local Computations with Probabilities on Graphical Structures and ...
    Local Computations with Probabilities on Graphical Structures and Their Application to. Expert Systems. Author(s): S. L. Lauritzen and D. J. Spiegelhalter.
  21. [21]
    [PDF] A simple approach to Bayesian network computations
    We call it the ordering determination component. Roughly speaking, an elimination ordering is good if the arithmetic calculations needed to sum out each.
  22. [22]
    [PDF] An Introduction to Conditional Random Fields - arXiv
    Nov 17, 2010 · For example, undirected models are commonly also referred to Markov random fields, Markov networks, and Gibbs distributions. As mentioned, we ...
  23. [23]
    [PDF] CS 6347 Lecture 4 Markov Random Fields
    • Each potential function can be specified as a table ... Conditional Random Fields (CRFs). • Undirected graphical models that represent conditional probability ...
  24. [24]
    [PDF] Conditional Random Fields: An Introduction
    Feb 24, 2004 · The potential functions must therefore ensure that it is possible to factorize the joint probability such that conditionally independent random ...<|separator|>
  25. [25]
    [PDF] Learning Factor Graphs in Polynomial Time & Sample Complexity
    the Hammersley-Clifford Theorem for Markov net- works (Hammersley & Clifford, 1971; Besag, 1974), we provide a parameterization of factor graph distributions.
  26. [26]
    [PDF] Dynamic Bayesian Networks∗ - UBC Computer Science
    Nov 12, 2002 · A dynamic Bayesian network (DBN) represents hidden states with a set of random variables, and uses directed arcs within and between time slices.
  27. [27]
    [PDF] Learning Bayesian Nets that Perform Well sl = VI, s4 = V4, s1 = V7 )?",
    also includes a conditional-probability-table that spec ifies how the node's ... Figure 1: Bayesian network structure for Example 4.2. ("Naive Bayes").
  28. [28]
    [PDF] Bayesian Networks and Markov Random Fields 1 Bayesian ... - TTIC
    The conditional probabilities of the form P(xi | parents of xi) are called conditional probability tables (CPTs). Suppose that each of the variables xi has d ...
  29. [29]
    [PDF] Bayesian Networks 2
    Mar 29, 2007 · E.g., estimate P(Rain|Sprinkler =true) using 100 samples. 27 samples have Sprinkler =true. Of these, 8 have Rain=true and 19 have Rain=false.
  30. [30]
    [PDF] Bayesian Networks Matthew Pettigrew Department of Mathematical ...
    May 5, 2016 · Cancer, however, has two parent nodes, and its local distribution is represented by a three dimensional conditional probability table (Table 2).
  31. [31]
    [PDF] Identifiability in Causal Bayesian Networks: A Gentle Introduction
    May 21, 2007 · In the sec- ond model, the joint probability of genotype, smoking, and lung cancer can be represented by the product of the prior probability of ...
  32. [32]
    Local Computations with Probabilities on Graphical Structures and ...
    S. L. Lauritzen. Aalborg University, Denmark. Search for more papers by this author · D. J. Spiegelhalter,. D. J. Spiegelhalter ... First published: 1988. https ...Missing: junction | Show results with:junction
  33. [33]
    Bayesian networks - Jensen - 2009 - WIREs Computational Statistics
    Dec 1, 2009 · equation image (4). This calculation is called marginalization, and we say that the variable B is marginalized out of P(A, B) [resulting in P(A)] ...Missing: formula | Show results with:formula
  34. [34]
    The computational complexity of probabilistic inference using ...
    We show that probabilistic inference using belief networks is NP-hard. Therefore, it seems unlikely that an exact algorithm can be developed to perform ...
  35. [35]
    Gibbs sampling in Bayesian networks - ScienceDirect.com
    References · 1. D.H. Ackley, G.E. Hinton, T.J. Sejnowski. A learning algorithm for Boltzmann machines · 2. J. Besag. Spatial interaction and the statistical ...Missing: reference | Show results with:reference
  36. [36]
    Generalized Belief Propagation - NIPS
    We show that BP can only converge to a stationary point of an approximate free energy, known as the Bethe free energy in statis(cid:173) tical physics.
  37. [37]
    [PDF] An Introduction to Variational Methods for Graphical Models
    The problem of probabilistic inference in graphical models is the problem of computing a conditional probability distribution over the values of some of the ...