Fact-checked by Grok 2 weeks ago

Gibbs sampling

Gibbs sampling is a (MCMC) for generating a sequence of observations from a specified multivariate when direct sampling is difficult, achieved by iteratively drawing samples from the full conditional distributions of each variable given the current values of the others. The method constructs a whose is the target joint distribution, allowing the generated samples to approximate the desired distribution after a sufficient number of iterations, typically after to . Originally introduced by Stuart Geman and Donald Geman in 1984 for Bayesian restoration of images using Gibbs random fields, the algorithm drew from concepts in statistical mechanics and provided a stochastic relaxation technique for sampling from complex joint distributions. Its adoption in statistics surged following the 1990 paper by Alan E. Gelfand and Adrian F. M. Smith, who demonstrated its utility for computing marginal posterior densities in Bayesian inference through data augmentation and iterative conditional sampling. This development marked a pivotal advancement in MCMC methods, building on earlier work like the Metropolis-Hastings algorithm and the Hammersley-Clifford theorem, which established equivalence between local conditional specifications and global joint distributions. Gibbs sampling excels in high-dimensional settings where full joint densities are intractable but conditional distributions are more accessible, making it a cornerstone for applications in , , and spatial modeling. Key advantages include its simplicity in implementation when conditionals are known and its ability to handle dependencies among variables without explicit , though it requires careful assessment of and mixing rates to ensure reliable approximations. Variations such as blocked or systematic-scan Gibbs samplers have extended its efficiency for specific problems, while diagnostic tools monitor chain to validate inferences.

Introduction

Definition and Motivation

Gibbs sampling is a (MCMC) algorithm designed to generate a sequence of samples from the of multiple interdependent random variables. It achieves this by iteratively drawing samples from the full conditional distribution of each variable, conditioned on the most recent values of all other variables, thereby constructing a whose approximates the target joint distribution. The algorithm derives its name from the physicist , owing to its connection with Gibbs distributions in , which model systems in ; however, its adoption in statistics was driven by Geman and Geman (1984), who applied it to Bayesian image problems using Gibbs random fields. The motivation for Gibbs sampling stems from the practical difficulties in directly sampling from high-dimensional or irregularly shaped multivariate distributions, where analytical forms are often unavailable or computationally prohibitive, as is common in Bayesian models with intractable normalizing constants for posterior distributions. By breaking down the joint sampling into simpler conditional steps—each typically univariate or low-dimensional—the method enables approximation of complex distributions through iteration, avoiding the need for direct evaluation of the full joint density. To illustrate intuitively, consider two correlated variables, such as the coordinates X and Y in a bivariate . Starting from an initial guess for Y, one samples a value for X conditioned on that Y, then updates Y based on the new X, and repeats this alternation; after many cycles, the paired samples (X, Y) trace out the contours of the joint distribution, even if direct joint sampling is cumbersome due to dependence.

Historical Development

The Gibbs sampling algorithm draws its name from the foundational work of physicist , who in his 1902 treatise Elementary Principles in Statistical Mechanics developed the concept of Gibbs distributions to describe equilibrium states in physical systems, laying the groundwork for probabilistic modeling of interacting particles. Although Gibbs's ideas originated in , they remained largely within physics until the 1980s, when statisticians began adapting them for computational purposes. The statistical adoption of Gibbs sampling began with its introduction as a computational tool by Stuart Geman and Donald Geman in 1984, who applied it to Bayesian restoration of binary lattice images in , framing the problem as sampling from a Gibbs distribution over pixel configurations. This marked the first prominent computational implementation, particularly for discrete lattice models in spatial statistics, as further explored by Julian Besag in 1986, who used it to analyze ""—noisy spatial data on grids—and emphasized its utility for simulating conditional distributions in Markov random fields. In the early 1990s, the method evolved to handle continuous variables through extensions by Alan E. Gelfand and Adrian F. M. Smith, who in 1990 demonstrated its applicability to marginal in Bayesian models with continuous parameters, broadening its scope beyond cases. Popularization in followed with the 1993 review by Adrian F. M. Smith and Geoffrey O. Roberts, which highlighted Gibbs sampling as a cornerstone of (MCMC) methods for posterior inference. Concurrently, the BUGS project, initiated by in 1989 at the MRC Biostatistics Unit, integrated Gibbs sampling into user-friendly software for Bayesian modeling, with its first release in 1991 facilitating widespread adoption during the MCMC explosion of the 1990s. This period saw Gibbs sampling drive the surge in MCMC applications across statistics, shifting focus from physics-inspired theory to practical computational inference.

Prerequisites

Markov Chain Monte Carlo Methods

Markov chain Monte Carlo (MCMC) methods constitute a class of algorithms designed to generate samples from a target \pi when direct sampling is infeasible due to high dimensionality or the intractability of the . These methods construct a whose is \pi, allowing the chain's long-run behavior to approximate integrals or expectations with respect to \pi via estimation. The foundational MCMC algorithm, introduced by et al., relies on a proposal distribution to generate candidate states and accepts or rejects them to maintain the target as invariant. Central to MCMC are the transition kernel K(x, y), which governs the probability of moving from state x to y, and the property of ergodicity, which ensures that the chain's time averages converge almost surely to the expectation under \pi as the number of iterations grows. Ergodicity typically requires irreducibility (the chain can reach any state from any other) and aperiodicity (no cyclic behavior), guaranteeing that sample averages from sufficiently long chains provide unbiased estimates of \pi-expectations. Unlike direct sampling techniques, such as or , which may suffer from inefficiency or bias in complex spaces like Bayesian posteriors with unknown marginal likelihoods, MCMC excels by iteratively exploring the state space without requiring the full form of \pi. Gibbs sampling represents one specific instance of an MCMC method, tailored for multivariate distributions via conditional updates. A key condition for ensuring \pi is the stationary distribution is detailed balance, which states that for all states x and y, \pi(x) K(x, y) = \pi(y) K(y, x). This equation implies that the flux of probability from x to y equals the reverse flux, preserving \pi under repeated transitions; it can be derived by requiring the Chapman-Kolmogorov equations to hold with \pi as invariant, leading to the symmetry in the joint \pi(x) K(x, y). To mitigate in generated samples, practitioners employ , discarding initial iterations until the chain approaches stationarity, and , retaining every k-th sample to approximate independence and reduce variance in estimators. These techniques address the transient phase where the chain may not yet reflect \pi, though their optimal application depends on the specific and target.

Bayesian Inference Basics

Bayesian inference updates beliefs about unknown parameters θ given observed data y using , which states that the posterior distribution is proportional to the product of the likelihood and the : p(θ|y) ∝ p(y|θ) p(θ). This framework treats parameters as random variables, allowing probabilistic statements about their values after incorporating evidence from the . In high-dimensional settings, computing the in the full expression p(θ|y) = p(y|θ) p(θ) / p(y), where p(y) = ∫ p(y|θ) p(θ) dθ requires intractable multi-dimensional , posing significant computational challenges. The distribution p(θ) encodes initial knowledge or assumptions about the parameters before observing the , while the likelihood p(y|θ) quantifies how well the model explains the observed under a given θ. The posterior p(θ|y) then combines these to yield updated beliefs. For analytical tractability, conjugate priors are often employed, where the prior and likelihood belong to the same of distributions, ensuring the posterior retains the same form as the prior; for instance, a prior paired with a likelihood yields a posterior. The primary goals of include estimating parameters through summaries of the posterior, such as the mean or mode, and constructing credible intervals that contain the true parameter value with a specified probability based on posterior samples. This intractability often stems from the need to marginalize over nuisance parameters—integrating them out to focus on parameters of interest—which exacerbates integration difficulties in complex models and motivates the use of methods for posterior approximation. A simple example is inferring the bias θ of a from n flips yielding y heads, modeled as y ~ (n, θ). With a conjugate (α, β) prior on θ, the posterior is (α + y, β + n - y), which can be computed directly without sampling: for n=10 flips with y=7 heads and a (1,1) prior, the posterior is (8,4), with mean 8/12 ≈ 0.667. This illustrates how Bayesian updating shifts beliefs toward the data while incorporating information, though such closed-form solutions become rare in more intricate scenarios.

Mathematical Foundations

Multivariate Probability Distributions

In multivariate probability distributions, the joint distribution p(\mathbf{x}) = p(x_1, \dots, x_d) specifies the probability (or ) over a d-dimensional random \mathbf{x}, capturing the simultaneous of all components. The marginal distribution for a single component x_i is obtained by integrating (or summing) the joint over the remaining variables: p(x_i) = \int p(\mathbf{x}) \, d\mathbf{x}_{-i}, where \mathbf{x}_{-i} denotes all components except x_i. This marginalization process reduces the dimensionality while preserving the relevant probabilities for x_i alone, and it extends naturally to subsets of variables. Conditional distributions play a central role in multivariate settings, with the full conditional for x_i given all other variables defined as p(x_i \mid \mathbf{x}_{-i}) = \frac{p(\mathbf{x})}{p(\mathbf{x}_{-i})}, assuming p(\mathbf{x}_{-i}) > 0. Partial conditionals, in contrast, condition x_i on only a of the other variables, such as p(x_i \mid \mathbf{x}_S) where S \subset \{1, \dots, d\} \setminus \{i\}, which can simplify computations when full conditioning is intractable. These conditionals encode dependencies among variables and are foundational for deriving properties of the joint. Under positivity assumptions (where all densities are strictly positive), the full conditionals uniquely determine the joint , a result formalized by the Hammersley-Clifford theorem for discrete cases, which equates Markov random fields to Gibbs distributions via conditional specifications. This theorem ensures compatibility, meaning a set of consistent conditionals corresponds to a unique joint, facilitating the specification of complex multivariate structures from local interactions. Multivariate distributions often leverage graph representations to model conditional independencies, particularly directed acyclic graphs (DAGs) in Bayesian networks, where nodes represent variables and directed edges indicate direct influences. In a DAG, conditional independencies are read off the structure using d-separation criteria, allowing the joint to factorize according to parental relationships. Specifically, the distribution factorizes as p(\mathbf{x}) = \prod_{i=1}^d p(x_i \mid \mathbf{x}_{\mathrm{pa}(i)}), where \mathrm{pa}(i) denotes the set of parents of node i in the DAG, enabling compact representation and efficient inference for high-dimensional problems.

Conditional and Joint Distributions

In Gibbs sampling, iteratively drawing samples from the full conditional distributions of each variable given the current values of the others induces the target distribution as the invariant measure of the resulting . This property arises because the satisfies with respect to the , ensuring that the is precisely the desired . A fundamental in this framework is that the for any variable x_i aligns with the through the : p(x_i) = \int p(x_i \mid x_{-i}) \, p(x_{-i}) \, dx_{-i}, where x_{-i} denotes all variables except x_i. The Gibbs sampling updates preserve this marginal consistency iteratively, as each conditional sample is drawn from the exact conditional under the joint, maintaining the overall joint as invariant. The set of full conditional distributions uniquely determines the joint distribution when the joint density is positive everywhere. This uniqueness follows from the Hammersley-Clifford theorem, which equates Markov random fields (defined via local conditionals) to Gibbs distributions (specified globally via an energy function). For positive densities p(x) > 0 for all x, the full conditionals p(x_i \mid x_{-i}) fix the ratios p(x)/p(x') for any configurations x, x' differing in one component, allowing recursive computation of the entire joint up to a ; the positivity ensures no zero probabilities disrupt this and that the measure is finite, yielding a unique joint. For variables, the Gibbs sampler applied to these full conditionals corresponds directly to , a heat-bath algorithm from for equilibrating systems under Gibbs measures, such as Ising models on lattices. To illustrate how conditionals reconstruct the , consider a trivariate distribution over binary variables X, Y, Z \in \{0,1\} with specified full conditionals, for example: P(X=1 \mid Y=y, Z=z) = \frac{1}{1 + e^{-(0.5y + 0.5z)}}, P(Y=1 \mid X=x, Z=z) = \frac{1}{1 + e^{-(0.5x + 0.5z)}}, and P(Z=1 \mid X=x, Y=y) = \frac{1}{1 + e^{-(0.5x + 0.5y)}}, corresponding to a symmetric interaction model. These conditionals imply a P(X=x, Y=y, Z=z) \propto \exp(0.5 (xy + xz + yz)) with \beta = 0.5, normalized over the 8 configurations; sampling from these conditionals will converge to this unique , demonstrating reconstruction via iterative marginalization and ratio consistency.

Algorithm Description

Core Procedure

The Gibbs sampler is a algorithm that generates a sequence of samples from a multivariate target distribution \pi(\mathbf{x}) by iteratively drawing from the full conditional distributions of each component. The process starts with an initial configuration \mathbf{x}^{(0)}, which can be chosen arbitrarily or from a simple prior distribution. For each iteration t = 1, 2, \dots, T, the algorithm updates the components of \mathbf{x}^{(t)} sequentially or randomly by sampling x_i^{(t)} \sim \pi(x_i \mid \mathbf{x}_{-i}^{(t-1)}) for i = 1, \dots, d, where \mathbf{x}_{-i} denotes all components except the i-th, and d is the of \mathbf{x}. This iterative conditional sampling ensures that the generated sequence \{\mathbf{x}^{(t)}\} forms a whose is the target \pi. Two primary variants differ in the order of component updates: systematic scan and random scan. In systematic scan, components are updated in a fixed sequential order (e.g., i = 1 to d) during each full sweep, which is computationally straightforward and often used in practice for its locality benefits in implementation. In contrast, random scan selects each component i uniformly at random for updating at every step, potentially improving mixing properties by avoiding order-dependent biases. Both approaches preserve the target as the invariant distribution, though their convergence rates can vary depending on the model structure. The transition kernel K(\mathbf{x}, \mathbf{y}) for a systematic scan Gibbs sampler, representing the probability of moving from state \mathbf{x} to \mathbf{y} in one full sweep, is given by the product of the conditional densities: K(\mathbf{x}, \mathbf{y}) = \prod_{i=1}^d \pi(y_i \mid y_{1:i-1}, x_{i+1:d}) if \mathbf{y} results from sequentially updating all components of \mathbf{x} via the conditionals; otherwise, K(\mathbf{x}, \mathbf{y}) = 0. This kernel ensures the chain transitions only through valid conditional samples. A key property of the Gibbs sampler is its reversibility with respect to the target distribution \pi, meaning the chain satisfies : \pi(\mathbf{x}) K(\mathbf{x}, \mathbf{y}) = \pi(\mathbf{y}) K(\mathbf{y}, \mathbf{x}) for all \mathbf{x}, \mathbf{y}. This holds for both scan variants under standard assumptions of positive densities, facilitating theoretical analysis and ensuring the chain's equilibrium aligns with \pi. In practice, the algorithm runs for a fixed number of iterations T or until empirical diagnostics indicate the chain has mixed sufficiently, such as monitoring or potential scale reduction factors across multiple chains.

Practical Implementation

Implementing Gibbs sampling in practice requires careful attention to initialization, iterative sampling from conditional distributions, and storage of the generated for posterior analysis. The basic proceeds by selecting starting values for the variables and iteratively updating each one conditionally on the current values of the others, typically for a large number of iterations to approximate the target distribution. The full pseudocode for a standard Gibbs sampler targeting a joint distribution p(\theta_1, \dots, \theta_d) is as follows:
Initialize θ^(0) = (θ_1^(0), ..., θ_d^(0)) arbitrarily
For i = 1 to N:
    For j = 1 to d:
        Sample θ_j^(i) ~ p(θ_j | θ_1^(i), ..., θ_{j-1}^(i), θ_{j+1}^(i-1), ..., θ_d^(i-1))
    Set θ^(i) = (θ_1^(i), ..., θ_d^(i))
Discard initial burn-in samples if needed
Output the chain {θ^(i)} for i = 1 to N
This procedure generates a Markov chain whose stationary distribution is the target joint distribution, provided the conditionals are correctly specified. Numerical considerations are crucial for stable implementation, particularly in models where densities may span many orders of magnitude. To avoid underflow or overflow when evaluating conditional densities, computations should be performed in log-space, where log-densities are summed rather than products of probabilities, ensuring especially for high-dimensional problems. In high dimensions, vectorized operations—such as multiplications for updating multiple components simultaneously—can significantly improve efficiency by leveraging optimized routines, reducing the per-iteration computational cost from O(d^2) to O(d) in suitable cases. A concrete example is the Gibbs sampler for a bivariate N(0, \Sigma) with \rho, where the conditionals are also normal: \theta_1 | \theta_2 \sim N(\rho \theta_2, 1 - \rho^2) and \theta_2 | \theta_1 \sim N(\rho \theta_1, 1 - \rho^2). The following Python-like snippet illustrates the implementation, storing samples in arrays and computing posterior summaries after a period:
python
import [numpy](/page/NumPy) as np

rho = 0.6  # Correlation parameter
burn_in = 1000
keep_draws = 10000
total_draws = burn_in + keep_draws

theta1 = np.zeros(total_draws)
theta2 = np.zeros(total_draws)
np.random.seed(123)  # For reproducibility

# Initialize (arbitrary starting point)
theta1[0] = 0
theta2[0] = 0

for i in range(1, total_draws):
    theta2[i] = rho * theta1[i-1] + np.sqrt(1 - rho**2) * np.random.randn()
    theta1[i] = rho * theta2[i] + np.sqrt(1 - rho**2) * np.random.randn()

# Discard [burn-in](/page/Burn-in) and store kept samples as [matrix](/page/Matrix)
samples = np.column_stack((theta1[burn_in:], theta2[burn_in:]))

# Compute [means](/page/Mean) and [covariance](/page/Covariance) from samples
means = np.mean(samples, axis=0)
cov_matrix = np.cov(samples, rowvar=False)
This generates samples approximating the target distribution, with the sample means close to zero and the approximating \begin{pmatrix} 1 & \rho \\ \rho & 1 \end{pmatrix}. During , trace plots— of the sampled values—is essential to assess chain mixing, as good mixing appears as rapid, random fluctuations without trends or stuck regions, indicating the sampler is exploring the posterior effectively. Samples are typically stored as matrices, with rows as iterations and columns as variables, facilitating efficient of posterior summaries like means, variances, and quantiles via column-wise operations.

Theoretical Properties

Ergodicity and Stationarity

In the context of Gibbs sampling, stationarity refers to the property that the target joint \pi(\mathbf{x}) serves as an invariant measure for the induced by successive conditional transitions. That is, if the chain starts from a distribution equal to \pi, the distribution remains \pi after any number of transitions. This invariance is ensured because the Gibbs sampler's transition kernel satisfies the condition: for any states \mathbf{x} and \mathbf{y}, \pi(\mathbf{x}) P(\mathbf{x}, \mathbf{y}) = \pi(\mathbf{y}) P(\mathbf{y}, \mathbf{x}), where P denotes the transition probability. A proof sketch of detailed balance for the Gibbs sampler proceeds by considering a single full sweep through the variables. Suppose the chain is at \mathbf{x} = (x_1, \dots, x_d); the transition to \mathbf{y} = (y_1, \dots, y_d) involves sequentially sampling y_i \sim \pi(x_i | \mathbf{x}_{-i}) for i = 1, \dots, d, where \mathbf{x}_{-i} denotes all components except the i-th. The probability of this path is \prod_{i=1}^d \pi(y_i | y_1, \dots, y_{i-1}, x_i, \dots, x_d). Reversing the path from \mathbf{y} to \mathbf{x} yields the same product due to the symmetry in conditional densities under the joint \pi, confirming \pi(\mathbf{x}) P(\mathbf{x}, \mathbf{y}) = \pi(\mathbf{y}) P(\mathbf{y}, \mathbf{x}). implies global balance, \sum_{\mathbf{x}} \pi(\mathbf{x}) P(\mathbf{x}, \mathbf{y}) = \pi(\mathbf{y}), establishing stationarity. Ergodicity of the Gibbs chain requires irreducibility, aperiodicity, and positive Harris recurrence to guarantee of empirical averages to expectations under \pi. Irreducibility means that from any state \mathbf{x}, the chain can reach any other state \mathbf{y} in the support of \pi with positive probability in finite steps, allowing exploration of the full state space. Aperiodicity ensures no deterministic cycles, so the chain does not oscillate between subsets periodically. Harris recurrence strengthens ordinary recurrence by stipulating that, for any starting state and measurable set A with \pi(A) > 0, the probability of returning to A is 1, and the expected return time is finite; this holds under mild regularity conditions on the conditionals, such as . Under , the for Markov chains applies: for a \pi-integrable function f, the ergodic average \frac{1}{T} \sum_{t=1}^T f(\mathbf{x}^{(t)}) \to \mathbb{E}_\pi[f(\mathbf{X})] as T \to \infty, where \mathbf{x}^{(t)} are the chain samples. This convergence relies on the chain being \psi-irreducible (with reference measure \psi), aperiodic, and positive Harris recurrent, ensuring a stationary distribution \pi and time-average equivalence to the ensemble average.

Convergence Behavior

The convergence of Gibbs sampling is fundamentally characterized by its mixing time, defined as the minimum number of iterations required for the distance between the chain's distribution and the target to fall below a small ε > 0, specifically τ_mix(ε) = min {k ≥ 1 : sup_x ||P^k(x, ·) - π||_{TV} ≤ ε}, where P^k denotes the k-step transition kernel and π is the measure. This metric quantifies how quickly the explores the state space and approaches stationarity, with slower mixing leading to inefficient sampling and biased estimates. Under appropriate conditions, such as those ensuring , Gibbs samplers can achieve geometric , meaning the chain converges exponentially fast to the at a rate ρ < 1, where ||P^k(x, ·) - π||{TV} ≤ M(x) ρ^k for some function M and initial state x. This property is typically established via drift conditions, which require the existence of a Lyapunov function V such that the expected change in V under the chain's dynamics satisfies E[V(X{t+1}) | X_t = x] ≤ λ V(x) + b 1_C(x) for λ < 1, bounded b, compact set C, and indicator 1_C, paired with a minorization condition to ensure positive probability of jumping to a reference measure. A canonical measure of serial dependence in Gibbs chains, which directly impacts effective sample size, is the autocorrelation function ρ(k) = Cov(f(X^{(t)}), f(X^{(t+k)})) / Var(f(X^{(t)})) for a suitable test function f, with the integrated autocorrelation time given by τ = 1 + 2 ∑_{k=1}^∞ ρ(k), where lower τ indicates faster decorrelation and better convergence. In practice, this time scales with the inverse of the spectral gap of the transition operator, providing a practical proxy for mixing efficiency. Strong correlations between variables in the target distribution can significantly slow mixing in Gibbs sampling, as updates to one component propagate sluggishly through the chain, often exacerbated by the curse of dimensionality in high-dimensional spaces where mixing times grow polynomially or worse with the number of variables. For uniform ergodicity, a stronger form implying geometric rates independent of starting point, Doeblin minorization conditions suffice in compact state spaces, requiring that there exists ε > 0 and a ν such that P(x, ·) ≥ ε ν(·) for all x in a substantial set, ensuring rapid regardless of initialization.

Applications

Bayesian Posterior Sampling

Gibbs sampling serves as a cornerstone method for approximating the posterior distribution in , where the goal is to sample from p(\theta \mid y) \propto p(y \mid \theta) p(\theta), with \theta denoting parameters and y the observed . By iteratively drawing from the full conditional distributions p(\theta_i \mid \theta_{-i}, y) for each component \theta_i, the algorithm generates a whose converges to the joint posterior, enabling estimation of posterior expectations, marginals, and quantiles. This approach is particularly valuable when direct sampling from the intractable posterior is infeasible to high dimensionality or non-conjugacy. In hierarchical Bayesian models, Gibbs sampling facilitates by alternately sampling latent variables and hyperparameters, conditioning on all other variables at each step. For instance, in models with multiple levels—such as random effects in multilevel —samples are drawn from the conditionals of latent variables given observed and parameters, followed by updates to the parameters given the latents and . This block-wise leverages the structure inherent in hierarchical priors, often simplifying the full conditionals even when the joint posterior is complex. The method's in such settings was demonstrated in early Bayesian applications, where it outperformed direct for multi-parameter problems. A canonical example is for the of a with unknown , employing a for the mean \beta and \tau = 1/\sigma^2. Given y_i \sim \mathcal{N}(\beta, \sigma^2) for i=1,\dots,n (equivalent to X as a column of ones) and prior \beta \mid \tau \sim \mathcal{N}(\mu_0, (\kappa_0 \tau)^{-1}), \tau \sim \Gamma(\alpha_0, \beta_0), the full conditionals are: \beta \mid \tau, y \sim \mathcal{N}\left( \mu_n, (\kappa_n \tau)^{-1} \right), where \mu_n = \frac{\kappa_0 \mu_0 + n \bar{y}}{\kappa_n} and \kappa_n = \kappa_0 + n, with n the sample size and \bar{y} the sample mean; and \tau \mid y, \beta \sim \Gamma\left( \alpha_n, \beta_n \right), with \alpha_n = \alpha_0 + n/2 and \beta_n = \beta_0 + \frac{1}{2} \left[ \sum_{i=1}^n (y_i - \beta)^2 + \kappa_0 (\beta - \mu_0)^2 \right]. These closed-form updates allow efficient iteration, yielding samples that approximate the posterior for \beta and \sigma^2. This conjugate setup, while idealized, illustrates Gibbs sampling's ability to handle uncertainty in both the mean and variance. From the generated samples, Bayesian practitioners can construct credible intervals by taking empirical quantiles of the \theta draws, providing probabilistic that integrates beliefs with data. Additionally, posterior predictive checks involve simulating new data y^{\text{rep}} \sim p(y \mid \theta^{(s)}) from each sample \theta^{(s)}, then comparing these replicates to observed y to assess model fit and validate assumptions. These diagnostics, derived directly from the Gibbs chain, enhance inferential reliability in applied Bayesian analysis.

Machine Learning and Beyond

In , Gibbs sampling has been instrumental in probabilistic topic modeling, particularly through the collapsed Gibbs sampler applied to (LDA). Introduced as a scalable method for LDA, this approach marginalizes out the topic distribution parameters, enabling efficient sampling from the posterior over topic assignments for words in a corpus, which facilitates the discovery of coherent latent topics in large text collections. For instance, in LDA, each is represented as a of topics, and each topic as a of words; Gibbs sampling iteratively updates the topic assignment for each word based on its given the current assignments and hyperparameters, converging to samples that approximate the posterior distribution over topics. This method has become a cornerstone for applications in , such as clustering and recommendation systems, due to its ability to handle high-dimensional data without variational approximations. Beyond topic modeling, Gibbs sampling supports image denoising tasks within frameworks, where it samples values to reconstruct clean images from noisy observations by enforcing local constraints. In these models, the Gibbs sampler alternates updates across sites, drawing from conditional distributions that balance data fidelity and prior spatial dependencies, often modeled via Ising or Gaussian potentials, leading to effective while preserving edges. Recent enhancements, such as moment-matching techniques, further accelerate for high-resolution images by incorporating denoising steps within the sampling process. In spatial statistics, Gibbs sampling underpins Bayesian mapping models, such as the Besag-York-Mollié (BYM) framework, which incorporates spatially structured random effects to estimate across regions while accounting for unobserved heterogeneity and . By sampling from the full conditional distributions of latent spatial effects, it enables robust inference on relative risks, as demonstrated in analyses of cancer incidence where it smooths extreme rates in small areas. Similarly, in , conjugate Gibbs sampling facilitates of evolutionary trees by alternately augmenting and updating tree topologies and branch lengths from their conditionals, improving efficiency over standard Metropolis-Hastings for complex substitution models. More recently, quantum extensions of Gibbs sampling have emerged for preparing thermal states of sparse Hamiltonians, where quantum circuits implement conditional updates to sample from Gibbs distributions efficiently on near-term devices, offering potential advantages over classical methods for simulating quantum many-body systems at constant temperatures. Advancing into interdisciplinary applications, 2025 developments include efficient Gibbs samplers for generating longitudinal synthetic populations, which iteratively condition on universal and disaggregate variables to produce time-consistent for transport and simulations, ensuring privacy-preserving representations of dynamic behaviors. In time series , fast Gibbs sampling has been adapted for Bayesian models with local-seasonal-global trends, enabling rapid posterior inference on trend and components to capture in economic and demand data. approaches, such as the Gibbs-MH , integrate Metropolis-Hastings steps within Gibbs iterations to update marginal posteriors in model tasks, addressing rate issues in high-dimensional Bayesian updating for applications.

Variations and Extensions

Blocked and Collapsed Variants

Blocked Gibbs sampling extends the standard Gibbs sampler by partitioning the variables into disjoint blocks and jointly sampling the variables within each block from their conditional distribution given the values in the other blocks, denoted as p(\mathbf{x}_B \mid \mathbf{x}_{-B}). This approach captures dependencies among highly correlated variables within a block, which can lead to faster mixing of the Markov chain compared to sampling variables individually. For instance, in models with tree-structured dependencies, such as hidden Markov models, the joint conditional for a block can be computed efficiently using algorithms like forward-backward dynamic programming. Collapsed Gibbs sampling improves efficiency by marginalizing out certain variables, typically nuisance parameters like model hyperparameters, to sample directly from the predictive distributions over the remaining variables. In a Bayesian model with latent variables \mathbf{z} and parameters \theta, the conditional for a component z_i becomes p(z_i \mid \mathbf{z}_{-i}, \text{data}) \propto \int p(z_i, \mathbf{z}_{-i}, \theta \mid \text{data}) \, d\theta = \frac{p(\mathbf{z} \mid \text{data})}{p(\mathbf{z}_{-i} \mid \text{data})}. This integration is often deterministic and analytically tractable when conjugate priors are used, avoiding the need to sample the marginalized variables explicitly. A prominent application of collapsed Gibbs sampling appears in (LDA) models for topic modeling, where parameters for topic-word and document-topic distributions are marginalized out, yielding a sampling rule based on word and topic counts. Collapsing in such latent variable models reduces among samples, enhancing rates relative to non-collapsed samplers.

Recent Advances

Recent advances in Gibbs sampling have focused on hybrid approaches that integrate Metropolis-Hastings (MH) steps to address challenges in sampling from complex marginals, particularly those exhibiting low acceptance rates. In 2024, a novel Gibbs-MH algorithm was proposed for Bayesian model updating, embedding MH proposals within the Gibbs framework to sample from joint distributions of parameters, improving efficiency when exact conditionals are intractable. This method enhances mixing by allowing adaptive proposals for marginals that would otherwise lead to poor acceptance in standard MH, as demonstrated in applications to structural reliability analysis. Similarly, the Metropolis-within-Gibbs (MwG) algorithm, analyzed in 2025 for log-concave distributions, replaces exact conditional updates with MH steps, achieving a spectral gap that scales favorably under certain conditions, thus mitigating slow mixing in high dimensions. Ordered overrelaxation has emerged as a complementary to accelerate mixing in Gibbs samplers by suppressing random walk behavior. In this variant, for updating component x_i given x_{-i}, a proposal y is first drawn from the conditional p(x_i \mid x_{-i}), and then accepted with probability \min\left(1, \frac{p(x_{-i} \mid y)}{p(x_{-i} \mid x)}\right), reflecting the sampler across the to explore farther from the current state while preserving . Recent implementations, such as in Langevin-based sampling from 2017 onward, have integrated ordered overrelaxation to draw momentum variables, reducing in neural network training and other high-dimensional tasks. Quantum extensions of Gibbs sampling, particularly via dissipative , have gained traction for preparing Gibbs s of sparse Hamiltonians. Between 2023 and 2025, algorithms based on Lindblad master equations have enabled efficient of open , where dissipative channels drive the system toward the target Gibbs \rho = e^{-\beta H}/Z with mixing times bounded by polylogarithmic factors in system size n. For random sparse models, these quantum Gibbs samplers achieve polynomial speedups in mixing time compared to classical counterparts, establishing quantum advantage for constant-temperature sampling on O(1)-local Hamiltonians. Key works include dissipative variational quantum algorithms that approximate free energy minimization, applicable to non-commuting Hamiltonians in . Adaptive block updates have further refined Gibbs sampling for scalable , dynamically grouping variables to optimize . In 2023, approximate blocked Gibbs methods were developed for Bayesian neural networks, updating parameter blocks conditionally to handle non-conjugacy while maintaining . By 2025, efficient samplers incorporating adaptive scanning of blocks achieved rapid mixing in probabilistic graphical models, as seen in motif discovery tasks combining advances. One-step Gibbs sampling has advanced synthetic data generation, enabling consistent hierarchical populations in a single iteration. In 2025, an adaptive one-step Gibbs sampler was introduced for longitudinal synthetic populations, using universal variables to derive disaggregated data while preserving privacy and statistical fidelity over time. This approach outperforms multi-step methods in efficiency for transportation modeling, generating households and activity patterns with minimal computational overhead.

Challenges

Failure Modes

Gibbs sampling, as a method, relies on the underlying chain being irreducible and aperiodic to ensure to the target distribution. Violations of these properties represent primary failure modes, where the chain either fails to explore the entire state space or cycles without converging. Specifically, the chain may lack irreducibility if high-probability regions are disconnected, or it may exhibit periodicity greater than one, leading to oscillatory behavior that prevents stationary . In distributions, Gibbs sampling often fails due to the chain becoming trapped in local s, violating irreducibility as it cannot cross low-density barriers between disconnected high-probability regions. This occurs because the method proposes updates based on conditional distributions that remain confined within a single , resulting in samples that inadequately represent the global posterior. For instance, in mixtures of Gaussians, standard Gibbs samplers initialized near one exhibit limited exploration, with trajectories rarely bridging to other s despite extended iterations. Strong dependencies between variables exacerbate mixing issues in Gibbs sampling, particularly in funnel-shaped posteriors where high correlations create regions of varying curvature. In such cases, the chain mixes slowly due to the need for tiny steps in narrow, high-curvature areas (like the "neck" of the ) to maintain , while larger steps are feasible but inefficient in broader regions, leading to poor overall exploration of the tails. A classic example is Neal's distribution, defined for variables \beta \in \mathbb{R} and \alpha \in \mathbb{R}^+ with density proportional to \alpha^{-1} \exp\left(-\frac{1}{2} \beta^2 / \alpha^2 - \frac{1}{2} (\log \alpha)^2\right), where Gibbs sampling fails to efficiently sample beyond \beta \approx -5 in reasonable time, as the conditional updates struggle with the multiscale geometry. Outlier problems with strong masking also cause Gibbs sampling to fail, as multiple high-leverage outliers can bias the conditional posteriors, creating a feedback loop that keeps the chain stuck in a masked regime without detecting the anomalies. This issue, prominent in robust regression models during the 1990s, arises because masking effects—where outliers mutually obscure each other's influence—make convergence unlikely, with the sampler rarely escaping the initial high-probability area around the masked data fit.

Diagnostics and Mitigation

Diagnosing in Gibbs sampling involves visual and quantitative methods to assess whether the has reached its and mixes adequately. Trace plots, which graph sampled values against iteration number, reveal patterns such as slow mixing or trends indicating non-stationarity; for instance, a well-mixing should exhibit rapid across the parameter space without long-term drifts. The Gelman-Rubin statistic, also known as the potential scale reduction factor (PSRF), monitors between-chain variance by comparing the pooled within-chain variance to the between-chain variance across multiple parallel chains; values below 1.1 typically indicate adequate . Another key diagnostic is the effective sample size (ESS), which quantifies the number of independent draws equivalent to the correlated MCMC samples, accounting for in the chain. The ESS is computed as \text{ESS} = \frac{n}{1 + 2 \sum_{k=1}^{\infty} \rho(k)}, where n is the number of samples and \rho(k) is the autocorrelation at lag k; low ESS values signal poor mixing and require longer runs. These diagnostics help detect common failure modes like poor mixing in high-dimensional or correlated posteriors, though they do not guarantee theoretical . To mitigate convergence issues in Gibbs sampling, running multiple chains from dispersed starting points enhances reliability, as it allows PSRF computation and reduces the impact of initial transients. , which discards intermediate samples to reduce , can improve but is often less effective than increasing chain length; reparameterization, such as centering variables to decorrelate components, accelerates mixing by altering the sampling geometry. For particularly challenging cases with funnel-shaped or multimodal posteriors, switching to (HMC) exploits gradient information for better exploration, often yielding higher per iteration than standard Gibbs. An illustrative example involves applying diagnostics from the package in to a Gibbs sampler for a multimodal Bayesian mixture model, where trace plots may show stuck chains in local modes and elevated R-hat values exceeding 1.2, prompting reparameterization or longer burns to achieve ESS > 100 per parameter. A recent advancement in mitigation is adaptive scanning in Gibbs samplers, which dynamically adjusts the order and probability of component updates based on empirical mixing rates, improving efficiency in large-scale models as demonstrated in 2024 benchmarks.

Software Tools

Key Libraries and Packages

JAGS (Just Another Gibbs Sampler) is a cross-platform program designed for the statistical analysis of Bayesian hierarchical models using (MCMC) simulation, particularly Gibbs sampling. It allows model specification in a BUGS-like declarative , enabling users to define complex probabilistic models intuitively without writing low-level sampling code. JAGS supports variants such as collapsed Gibbs sampling, which integrates out certain parameters to improve efficiency in models like . As of 2025, the latest stable release remains version 4.3.2, released in 2023, with ongoing maintenance for compatibility across Windows, macOS, and platforms. The family of software, including Classic BUGS, WinBUGS, and OpenBUGS, represents the foundational tools for using Gibbs sampling, originally developed in the late 1980s at the MRC Unit. These programs pioneered automated full conditional distribution generation for Gibbs samplers in graphical models, supporting a wide range of hierarchical structures through a simple . While highly influential, they are now considered legacy software, with limited updates since the early ; WinBUGS and OpenBUGS remain available for Windows and cross-platform use, respectively, but users are encouraged to migrate to modern alternatives for better performance and extensibility. Other prominent libraries integrate Gibbs sampling within broader MCMC frameworks. PyMC, a Python-based probabilistic programming library, supports Gibbs sampling through specific step methods like BinaryGibbsMetropolis for variables or custom implementations for conjugate models, while seamlessly combining it with advanced samplers like No-U-Turn Sampler (NUTS) for inference in scalable models. As of 2025, PyMC version 5.x emphasizes GPU acceleration and integration for faster MCMC, making it suitable for applications beyond pure Gibbs workflows. Stan, implemented in its domain-specific Stan language with interfaces for , Python, and others, primarily relies on (HMC) via NUTS but accommodates Gibbs sampling implementations for discrete or conditionally conjugate parameters through user-defined blocks or incremental strategies. Nimble, an for flexible MCMC, extends BUGS-like modeling with customizable samplers including Gibbs for hierarchical models, supporting build-process interventions for efficiency. TensorFlow Probability, a Python library for probabilistic reasoning, includes MCMC tools that facilitate Gibbs sampling components within broader pipelines, particularly for deep generative models. BLUPF90, a suite tailored for and , incorporates dedicated Gibbs samplers like GIBBS1F90 and GIBBS2F90 for estimating variance components in large-scale mixed models, with optimizations for multi-trait and ; its 2025 standard edition includes enhanced data handling for genomic integrations.
LibraryPrimary Language/InterfaceEase of Use (for Bayesian Modelers)Supported Gibbs Variants
JAGSBUGS-like scripting (C++ core)High: Intuitive model syntax, no coding for samplersStandard, collapsed, blocked
BUGS Family scripting (various)Medium: Legacy interface, graphical options in WinBUGSStandard, basic blocked
PyMCHigh: Integrates with / ecosystemsSpecific (e.g., BinaryGibbsMetropolis), custom/conjugate, integrated with NUTS
Stan language (// interfaces)Medium: Requires compilation, strong for HMC hybridsUser-implemented for discrete/conditional
NimbleHigh: Extends with custom samplersStandard, customizable blocked
TensorFlow ProbabilityMedium: Integrates with TF ecosystem for MLComponents within MCMC, user-defined
BLUPF90Parameter files ( core)Medium: Specialized for genetics, steep for generalistsStandard, multi-trait blocked, threshold

Implementation Examples

JAGS provides a flexible framework for implementing Gibbs sampling through its BUGS-like , interfaced with via the rjags package. Consider a bivariate normal model where observations are drawn from a with unknown mean vector and precision matrix. The model specification in JAGS is as follows:
model {
  for (i in 1:n) {
    y[i, 1:2] ~ dmnorm(mu[], Omega[,])
  }
  mu[1] ~ dnorm(0, 0.001)
  mu[2] ~ dnorm(0, 0.001)
  tau ~ dgamma(0.001, 0.001)
  rho ~ dnorm(0, 0.001) T(-1, 1)
  Omega[1,1] <- tau * (1 - rho * rho)
  Omega[1,2] <- -tau * rho
  Omega[2,1] <- Omega[1,2]
  Omega[2,2] <- tau
}
To run this via the interface, first prepare , such as n <- 100; y <- mvtnorm::rmvnorm(n, mean = c(1, 2), sigma = matrix(c(1, 0.5, 0.5, 1), 2)) (requiring the mvtnorm package). Load the rjags library, compile the model with jagmodel <- jags.model(textConnection(model_string), data = list(y = y, n = n), n.chains = 2, n.adapt = 1000), initialize with random starting values using init <- list(mu = c(0,0), tau = 1, rho = 0), and monitor parameters with params <- c("mu", "tau", "rho"). Sample from the posterior using samples <- coda.samples(jagmodel, params, n.iter = 10000). This setup leverages JAGS's automatic derivation of full conditional distributions for conjugate models like the multivariate normal, enabling efficient Gibbs sampling. For non-conjugate models, such as those involving binary outcomes with logistic links, JAGS handles custom conditionals by allowing users to specify deterministic relations or auxiliary variables that facilitate sampling, often through techniques like the Polya-Gamma method integrated in extensions. In PyMC, Gibbs sampling can be applied to hierarchical models using specific step methods for suitable variables or custom conjugate samplers. For example, consider a simple with group-level effects on , where we use steps to approximate conditional sampling (note: full Gibbs requires custom implementation for non-binary/conjugate parts). Define the model using:
import pymc as pm
import numpy as np
import arviz as az

np.random.seed(42)
n_groups, n_per_group = 5, 20
group_idx = np.repeat(np.arange(n_groups), n_per_group)
true_betas = np.array([0, 1, 2])
X = np.column_stack([np.ones(n_groups * n_per_group), np.random.randn(n_groups * n_per_group), group_idx])
y = X @ true_betas + np.random.randn(n_groups * n_per_group) * 0.5

with pm.Model() as hierarchical_model:
    mu_alpha = pm.Normal("mu_alpha", 0, 10)
    sigma_alpha = pm.HalfNormal("sigma_alpha", 5)
    alpha = pm.Normal("alpha", mu=mu_alpha, sigma=sigma_alpha, shape=n_groups)
    [beta](/page/Beta) = pm.Normal("beta", 0, 10, shape=2)
    sigma = pm.HalfNormal("sigma", 5)
    mu = alpha[group_idx] + [beta](/page/Beta)[0] + [beta](/page/Beta)[1] * X[:, 1]
    y_obs = pm.Normal("y_obs", mu=mu, sigma=sigma, observed=y)
    step = pm.Metropolis()  # Approximation; for true Gibbs, use custom conjugate steps
    trace = pm.sample(2000, step=step, tune=1000, chains=2, return_inferencedata=True)
This uses a Metropolis step method as an approximation; for exact Gibbs sampling in conjugate settings, custom step methods can be defined to sample from full conditionals, such as for the normal priors and likelihood. To visualize convergence, plot the trace for key parameters like alpha using az.plot_trace(trace). From model specification to posterior summaries, proceed step-by-step: (1) Define priors and likelihood as above; (2) compile and sample to obtain the trace; (3) compute summaries such as means and 95% quantiles with az.summary(trace, var_names=["alpha", "beta", "sigma"]), yielding results like mean alpha ≈ [0.02, 1.01, 1.98, ...] and credible intervals confirming the true values. Available packages like rjags and PyMC facilitate these implementations across languages. A practical tip for integrating recent advances involves implementing a one-step Gibbs sampler in custom code for generating synthetic hierarchical data, as in tasks; for 2025 synthetic datasets mimicking structures, initialize with marginal distributions and iteratively condition once across variables to produce correlated samples efficiently, avoiding multi-iteration overhead. For instance, in a bivariate case with marginals N(μ1, σ1) and N(μ2, σ2), draw x1 from its marginal, then x2 | x1 ~ N(μ2 + ρ σ2 / σ1 (x1 - μ1), σ2 (1 - ρ²)), completing the joint in one step for quick posterior approximation on .

References

  1. [1]
    [PDF] BU-1098-MA Explaining the Gibbs Sampler - Cornell eCommons
    The Gibbs sampler is a technique for generating random variables from a (marginal) distribution indirectly, without having to calculate the density. Although ...
  2. [2]
    [PDF] A Short History of Markov Chain Monte Carlo - arXiv
    Jan 9, 2012 · This paper is also responsible for the name Gibbs sampling, because it implemented this method for the Bayesian study of Gibbs random fields ...
  3. [3]
    Stochastic Relaxation, Gibbs Distributions, and the Bayesian ...
    Nov 30, 1984 · Stochastic Relaxation, Gibbs Distributions, and the Bayesian Restoration of Images. Publisher: IEEE. Cite This. PDF. Stuart Geman; Donald Geman.
  4. [4]
    [PDF] Sampling-Based Approaches to Calculating Marginal Densities
    Jul 14, 2002 · June 1990, Vol. 85, No. 410, Theory and Methods. 398. Page 3. Gelfand and Smith: Sampling-Based Approaches to Calculating Marginal Densities. 2 ...
  5. [5]
    [PDF] Explaining the Gibbs Sampler George Casella - Stat@Duke
    Feb 5, 2008 · The Gibbs sampler enjoyed an initial surge of pop- ularity starting with the paper of Geman and Geman (1984), who studied image-processing ...Missing: seminal | Show results with:seminal
  6. [6]
    A Short History of Markov Chain Monte Carlo - Project Euclid
    Gibbs random fields which, in turn, derive their name from the physicist Josiah Willard Gibbs (1839–1903). This original implementation of the Gibbs sampler was.
  7. [7]
    [PDF] Elementary Principles in Statistical Mechanics
    This book, Elementary Principles in Statistical. Mechanics, first published in 1902, gives his mature vision of these subjects. Mathematicians, physicists and ...
  8. [8]
    [PDF] Stochastic Relaxation, Gibbs Distributions, and the Bayesian ...
    GEMAN AND GEMAN: STOCHASTIC RELAXATION, GIBBS DISTRIBUTIONS, AND BAYESIAN RESTORATION tional burden. We shall return to these points in Sections IV and XI ...
  9. [9]
    [PDF] On the Statistical Analysis of Dirty Pictures Julian Besag Journal of ...
    Feb 6, 2008 · As Geman and Geman (1984) point out, any property of the (posterior) distribution P(x I y) can be simulated by running the. Gibbs sampler at " ...
  10. [10]
    [PDF] Sampling-Based Approaches to Calculating Marginal Densities Alan ...
    Nov 17, 2007 · Gelfand and Smith: Sampling-Based Approaches to Calculating Marginal Densities ... generalization to k variables, we see that Gibbs sampling.
  11. [11]
    Bayesian Computation Via the Gibbs Sampler and Related Markov ...
    SUMMARY. The use of the Gibbs sampler for Bayesian computation is reviewed and illustrated in the context of some canonical examples.
  12. [12]
  13. [13]
    A Short Review of Ergodicity and Convergence of Markov chain ...
    Oct 13, 2021 · This short note reviews the basic theory for quantifying both the asymptotic and preasymptotic convergence of Markov chain Monte Carlo ...
  14. [14]
    [PDF] Principle of Detailed Balance and Convergence Assessment ... - arXiv
    Jul 20, 2008 · Definition 1 (Principle of detailed balance) Transition probability matrix P and probability distribution π are said to be in detailed balance, ...
  15. [15]
    [PDF] Fast compression of MCMC output arXiv:2107.04552v2 [stat.CO] 19 ...
    Jul 19, 2021 · Historically, the two common recipes for compressing MCMC output are: • burn-in, which amounts to discarding the b first states; and. • thinning ...
  16. [16]
    [PDF] Bayesian Data Analysis Third edition (with errors fixed as of 20 ...
    ... Bayesian data analysis. This book is concerned with practical methods for making inferences from data using prob- ability models for quantities we observe ...
  17. [17]
    [PDF] CPSC 540: Machine Learning - Conjugate Priors
    Consider again a coin-flipping example with a Bernoulli variable, x ∼ Ber(θ). Previously we considered that either θ = 1 or θ = 0.5. Today: θ is a ...
  18. [18]
    Bayesian Inference of a Binomial Proportion - The Analytical Approach
    Our example is that of a sequence of coin flips. We are interested in the probability of the coin coming up heads. In particular, we are interested in the ...
  19. [19]
    [PDF] Chapter 3. Multivariate Distributions. All of the most interesting ...
    it is possible to construct a bivariate distribution from two components: either marginal distribution and the conditional distribution of the other variable ...
  20. [20]
    [PDF] Chapters 5. Multivariate Probability Distributions - Brown University
    General description: The marginal cdf for X is. FX(x) = F(x,∞). Page 11. Joint distribution determines the marginal distributions. Not vice versa. x1 x2 x3.
  21. [21]
    [PDF] Spatial Interaction and the Statistical Analysis of Lattice Systems ...
    Feb 6, 2008 · This paper examines conditional probability models for spatial interaction, using the Hammersley-Clifford theorem, and applies these models to ...
  22. [22]
    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 ...
  23. [23]
    Sampling-Based Approaches to Calculating Marginal Densities - jstor
    June 1990, Vol. 85, No. 410, Theory and Methods. 398. Page 2. Gelfand and Smith: Sampling-Based Approaches to Calculating Marginal Densities 399. 2. SAMPLING ...
  24. [24]
    Gibbs Sampler - an overview | ScienceDirect Topics
    The Gibbs sampler is an MCMC algorithm for obtaining a sequence of samples x i from a multivariate joint probability distribution when the joint distribution is ...
  25. [25]
    Metropolis-Hastings using log of the density - Cross Validated
    Feb 14, 2015 · The reason for my question is that you often store the log of the density when designing a Gibbs sampler, especially when using an object- ...
  26. [26]
    None
    ### Extracted Pseudocode/Algorithm Description for Gibbs Sampling
  27. [27]
    Gibbs Sampling from a Bivariate Normal Distribution - Aptech
    Use the Gibbs sampler to generate bivariate normal draws. Create side-by-side plots of the parameter paths. Calculate the drawn distribution's mean and variance ...
  28. [28]
    [PDF] An Improved !R for Assessing Convergence of MCMC
    Adapted from Gelman et al. (2013). visually inspect the sample paths of the chains via trace plots as well as study summary statistics such as the empirical ...
  29. [29]
    [PDF] Markov Chains for Exploring Posterior Distributions Luke Tierney ...
    Mar 12, 2008 · Ergodicity of degree 2 is typically very difficult to verify in practice. Two stronger forms of ergodicity are calledgeometric and uniform ...
  30. [30]
    Minorization Conditions and Convergence Rates for Markov Chain ...
    This paper provides methods using minorization conditions to analyze convergence of MCMC, establishing bounds on how long simulations should run.
  31. [31]
    Geometric ergodicity of Gibbs samplers in Bayesian penalized ...
    We show that the MCMC samplers used in the three models converge to their respective stationary distribution at a geometric rate. That is, we show that the ...
  32. [32]
    [PDF] Moment Matching Denoising Gibbs Sampling
    We illustrate how to scale our method for high-dimensional data and demonstrate the generation of high-quality images using only a single level of fixed noise.
  33. [33]
    Bayesian spatial analysis and disease mapping - PubMed Central
    The aim of our study was to predict the spatial distributions of Schistosoma haematobium and Schistosoma mansoni infections to assist planning the ...
  34. [34]
    Conjugate Gibbs sampling for Bayesian phylogenetic models
    This method, which we call conjugate Gibbs, relies on analytical conjugacy properties, and is based on an alternation between data augmentation and Gibbs ...
  35. [35]
    Sparse Random Hamiltonians Are Quantumly Easy | Phys. Rev. X
    Feb 9, 2024 · In this work, we identify a large class of Hamiltonians, random sparse matrices, which are quantumly easy to solve but are nontrivial for classical machines.
  36. [36]
    [PDF] Gibbs Sampler for Generating Longitudinal Synthetic Populations
    The Gibbs sampler generates longitudinal synthetic populations using universal variables, ensuring consistency, efficient data derivation, and disaggregated ...
  37. [37]
    Fast Gibbs sampling for the local-seasonal-global trend Bayesian ...
    Apr 9, 2025 · A generalised exponential smoothing model was proposed that is able to capture strong trends and volatility in time series.
  38. [38]
    A novel Gibbs-MH sampling algorithm for Bayesian model updating
    Gibbs sampling, a special form of the Metropolis-Hastings (MH) sampling method, samples the posterior probability function with multidimensional parameters. It ...Missing: motivation | Show results with:motivation
  39. [39]
    [PDF] The Collapsed Gibbs Sampler in Bayesian Computations with ...
    Feb 27, 2012 · This article describes a method of “grouping” and “collapsing” in using the Gibbs sampler and proves from an operator theory.
  40. [40]
    [PDF] Dynamic Blocking and Collapsing for Gibbs Sampling
    Joint sampling is more ... A much better solution in this case is to collapse F, create two blocks {A, B, C, D} and {E} and perform blocked Gibbs sampling over ...
  41. [41]
    [PDF] Collapsed & Blocked Gibbs Samplers - Brown Computer Science
    Nov 6, 2014 · A blocked Gibbs sampler iteratively resamples subsets of the original variables that are tractable (form a sub-graph without cycles):.
  42. [42]
    The Collapsed Gibbs Sampler in Bayesian Computations with ...
    Feb 27, 2012 · This article describes a method of “grouping” and “collapsing” in using the Gibbs sampler and proves from an operator theory viewpoint that the method is in ...
  43. [43]
    Finding scientific topics - PNAS
    We applied our Gibbs sampling algorithm to this dataset, together with the two algorithms that have previously been used for inference in Latent Dirichlet ...
  44. [44]
    Suppressing Random Walks in Markov Chain Monte Carlo Using ...
    Jun 22, 1995 · Such random walks can sometimes be suppressed using ``overrelaxed'' variants of Gibbs sampling (a.k.a. the heatbath algorithm), but such methods ...
  45. [45]
    Ordered over-relaxation based Langevin Monte Carlo sampling for ...
    Jan 12, 2017 · We first use the ordered over-relaxation method to draw the momentum variable which could suppress the random walk behavior in Gibbs sampling ...
  46. [46]
    Lindblad engineering for quantum Gibbs state preparation under the ...
    Aug 29, 2025 · ``Gibbs Sampling gives Quantum Advantage at Constant ... ``Mixing time of quantum Gibbs sampling for random sparse Hamiltonians'' (2024).
  47. [47]
    Mixing Time of Quantum Gibbs Sampling for Random Sparse ...
    Sep 9, 2025 · In this work, we establish a polylog(n) upper bound on its mixing time for various families of random n × n sparse Hamiltonians at any constant ...
  48. [48]
    Approximate blocked Gibbs sampling for Bayesian neural networks
    Aug 10, 2023 · A blocked Gibbs sampling algorithm samples groups (blocks) of two or more parameters conditioned on all other other parameters, rather than ...<|separator|>
  49. [49]
    Efficient Gibbs Samplers - Emergent Mind
    Aug 5, 2025 · Efficient Gibbs samplers are randomized algorithms designed to sample from probability distributions, particularly complex or high-dimensional ...
  50. [50]
    [PDF] FiGS-MoD: Feature-informed Gibbs Sampling Motif Discovery ...
    Oct 23, 2025 · Combined with recent advances in machine learning, this creates new opportunities for large-scale, high- resolution motif discovery. Results ...
  51. [51]
    Adaptive synthetic generation using one-step Gibbs Sampler
    This paper introduces an adaptive approach to synthetic population generation using a one-step Gibbs Sampler that allows for the maintenance of synthetic data ...
  52. [52]
    [PDF] MCMC - Art Owen
    We saw in Figure 12.3 that the Gibbs sampler can fail to be irreducible when the space is disconnected. A more common worry with the Gibbs sampler is that ...<|separator|>
  53. [53]
    [PDF] Markov Chain Monte Carlo and Gibbs Sampling
    Apr 26, 2004 · Gelfand and Smith (1990) illustrated the power of the Gibbs sampler to ad- dress a wide variety of statistical issues, while Smith and Roberts ( ...<|control11|><|separator|>
  54. [54]
  55. [55]
  56. [56]
  57. [57]
    Gibbs Sampling Will Fail in Outlier Problems with Strong Masking
    Section 3 analyzes the reasons of the slow convergence of the algorithm in data set with masked high leverage outliers and justifies that this problem does not ...
  58. [58]
    Markov Chain Monte Carlo Convergence Diagnostics
    We provide an expository review of 13 convergence diagnostics, describing the theoretical basis and practical implementation of each.
  59. [59]
    [PDF] Inference from Iterative Simulation Using Multiple Sequences
    GELMAN, A. and RUBIN, D. B. (1992). A single series from the. Gibbs sampler provides a false sense of security. In Bayesian. Statistics 4 ...
  60. [60]
    [PDF] General Methods for Monitoring Convergence of Iterative Simulations
    The aim of this article has been to suggest a number of alternative implementations of the original convergence diagnostic proposed by Gelman and Rubin (1992a).
  61. [61]
    [PDF] On Reparametrization and the Gibbs Sampler - Statistics
    In this article, we specify conditions under which reparametrization leaves the convergence rate of a Gibbs chain unchanged. An example illus- trates how ...Missing: mitigation multiple
  62. [62]
    [PDF] CODA: Convergence Diagnosis and Output Analysis for MCMC
    Trace plots are useful for diagnosing very poor mixing, a phenomenon in which the MCMC sampler covers the support of the posterior distribution very slowly.
  63. [63]
    Accelerated Markov Chain Monte Carlo Using Adaptive Weighting ...
    Aug 23, 2024 · In this paper, we focus on a random scan Gibbs sampling method that selects each latent variable non-uniformly.
  64. [64]
    JAGS - Just Another Gibbs Sampler
    See the JAGS NEWS blog for news about the project. If you want to be kept informed of updates to JAGS, then subscribe to the RSS news feed. Latest version.Missing: 2025 | Show results with:2025
  65. [65]
    JAGS: Just Another Gibbs Sampler - SourceForge.net
    Rating 5.0 (31) · Free · Utilities/ToolsJAGS is Just Another Gibbs Sampler. It is a program for the statistical analysis of Bayesian hierarchical models by Markov Chain Monte Carlo.Windows · Mac OS X · Files · Browse /JAGS/4.x at...
  66. [66]
    Just Another Gibbs Sampler (JAGS): Flexible Software for MCMC ...
    However, JAGS continues to be updated regularly, offering new samplers, modules, and other modeling features. Some of these developments are designed to ...
  67. [67]
    The BUGS Project | MRC Biostatistics Unit - University of Cambridge
    BUGS is a language and various software packages for Bayesian inference Using Gibbs Sampling, conceived and initially developed at the BSU.
  68. [68]
    WinBUGS: a tutorial - Lykou - 2011 - Wiley Interdisciplinary Reviews
    Jun 9, 2011 · HISTORICAL BACKGROUND. The BUGS (Bayesian inference using Gibbs sampling) project was introduced in 1989 by the research group of David ...
  69. [69]
    [PDF] The BUGS project: Evolution, critique and future directions
    BUGS is a software package for Bayesian inference using Gibbs sampling, which began in 1989 and has become a popular statistical modeling package.
  70. [70]
  71. [71]
    Using a custom step method for sampling from locally conjugate ...
    Nov 17, 2020 · In this notebook, we show how to implement a conjugate sampling scheme in PyMC and compare it against a full-HMC (or, in this case, NUTS) approach.
  72. [72]
    Current version of HMC/NUTS in PyMC - Development
    Mar 20, 2025 · We introduce a novel and flexible framework for constructing locally adaptive Hamiltonian Monte Carlo (HMC) samplers by Gibbs sampling the ...Missing: support | Show results with:support
  73. [73]
    Gibbs Sampling - Stan User's Guide
    19.7 Sampling Difficulties with Problematic Priors. With an improper posterior, it is theoretically impossible to properly explore the posterior.Missing: modes | Show results with:modes
  74. [74]
    Gibbs Sampling in Stan - Algorithms
    Mar 7, 2019 · The paper outlines how Gibbs sampling can be used to estimate the prevalence (\pi) of a disease, the sensitivity (S) , and the specificity (C) from the results ...
  75. [75]
    [PDF] Introduction to BLUPF90 suite programs Standard Edition (Ver 1.1.0)
    Apr 30, 2025 · This software is currently recommended for all users. How- ever, several Gibbs sampling programs were previously available in the BLUPF90 family ...
  76. [76]
    review_of_programs [BLUPF90]
    GIBBS2F90 adds joint sampling of correlated effects. This results in faster mixing with random regression and maternal models. GIBBS2F90 is usually the program ...Missing: features | Show results with:features
  77. [77]
    [PDF] JAGS Version 4.3.0 user manual
    Jun 28, 2017 · JAGS is “Just Another Gibbs Sampler”. It is a program for the analysis of Bayesian models using Markov Chain Monte Carlo (MCMC) which is not ...
  78. [78]
    [PDF] Adaptive synthetic generation using one-step Gibbs Sampler
    Apr 24, 2025 · This paper introduces an adaptive approach to synthetic population generation using a one-step Gibbs Sampler that allows for the main-.Missing: scan | Show results with:scan
  79. [79]
    One-step Gibbs sampling for the generation of synthetic households
    We propose an alternative practical extension of Gibbs sampling for generating hierarchical datasets such as synthetic households.