Fact-checked by Grok 2 weeks ago

Prior probability

In , the prior probability, often referred to simply as the prior, is the assigned to an uncertain quantity or before any relevant is taken into account. It represents the initial degree of belief in a or , derived from existing knowledge, expert opinion, or assumptions of ignorance. This distribution serves as the starting point for inference, which is updated through by incorporating observed data to yield the . The concept of prior probability originated in the work of , an English mathematician and Presbyterian minister, whose essay "An Essay towards solving a Problem in the Doctrine of Chances" was published posthumously in 1763. Bayes' ideas were further developed by in the late , who applied them to problems using uniform priors and conjugate models, such as the beta-binomial for proportions. By the , the approach gained prominence among probabilists but faced criticism for the perceived subjectivity in selecting priors, leading to a decline in favor of frequentist methods in the early 20th century. The modern Bayesian revival began mid-century, driven by figures like , , Jimmy Savage, and Dennis Lindley, who emphasized subjective probability and ; computational advances in the late 1980s and , including methods, made prior-based inference practical for complex models. Priors can be classified as informative, drawing on historical or domain expertise to encode specific beliefs (e.g., a centered around an expected success rate), or non-informative, such as Jeffreys priors, which reflect minimal prior knowledge by being invariant to reparameterization and often yielding flat or weakly informative distributions. Conjugate priors, like the for likelihoods, are particularly useful because they result in posteriors from the same family, simplifying analytical computations. The choice and impact of priors remain central to Bayesian , balancing beliefs with to quantify , though debates persist over their subjectivity and sensitivity in analyses.

Fundamentals

Definition and Interpretation

In , the prior probability refers to the assigned to an unknown parameter θ or hypothesis before observing any data, denoted as p(θ), which encapsulates the initial state of knowledge or belief about θ. This distribution serves as the starting point for , representing or available prior to . The interpretation of prior probabilities can be subjective or . In the subjective view, priors reflect the personal beliefs or expert knowledge of the analyst, allowing for the incorporation of relevant prior information into the analysis. Conversely, the objective approach seeks priors that are minimally informative or derived from formal principles to ensure and lack of personal , as discussed in foundational works on Bayesian . Historically, introduced an early objective perspective through his 1812 principle of insufficient reason (also known as the principle of indifference), which posits that in the absence of information favoring one outcome over another, equal probabilities should be assigned to equiprobable possibilities. For continuous parameters, the prior is typically expressed as a probability density function p(θ), which integrates to 1 over the parameter space, while for discrete cases, it is a specifying probabilities for each possible value. A simple example illustrates this: consider estimating the bias θ (probability of heads) of a with no prior flips observed; a (1,1) prior, which is uniform over [0,1], represents complete ignorance about θ by assigning equal density to all values.

Role in Bayesian Inference

In Bayesian inference, the prior probability plays a central role by serving as the initial distribution over possible parameter values or hypotheses before observing the data, which is then updated through to form the posterior distribution. states that the posterior probability density function is proportional to the product of the likelihood and the prior: p(\theta \mid y) \propto p(y \mid \theta) \cdot p(\theta), where \theta represents the parameters, y the observed data, p(\theta) the , and p(y \mid \theta) the likelihood; this multiplicative ensures that the prior directly influences the weighting of the likelihood in yielding updated beliefs about \theta. The full posterior is obtained by normalizing this product: p(\theta \mid y) = \frac{p(y \mid \theta) \cdot p(\theta)}{p(y)}, with the p(y) = \int p(y \mid \theta) p(\theta) \, d\theta acting as the that ensures the posterior integrates to 1, thereby quantifying the total probability of the data under the model and facilitating model comparison. The posterior distribution encapsulates the updated beliefs, combining the prior's information with the data's evidential content via the likelihood; for predictive purposes, one can marginalize the posterior over \theta to obtain the predictive distribution p(\tilde{y} \mid y) = \int p(\tilde{y} \mid \theta) p(\theta \mid y) \, d\theta, where \tilde{y} denotes future observations. This integration reflects how the prior shapes not only parameter inference but also forecasts by propagating initial uncertainties through the model. Conceptually, Bayesian updating follows a sequential flow: begin with the p(\theta) encoding pre-data knowledge, incorporate the likelihood p(y \mid \theta) to reflect data compatibility, and arrive at the posterior p(\theta \mid y) as the synthesis, with the serving as a normalizing bridge. A simple discrete example illustrates this in disease testing: suppose the probability of having a is 0.01 (1% ), and a with 99% (true positive rate) and 99% specificity (true negative rate) yields a positive result. The likelihood of a positive given the is 0.99, and given no is 0.01 (). Applying , the of having the is approximately 0.50, calculated as p(\text{disease} \mid +) = \frac{0.99 \times 0.01}{0.99 \times 0.01 + 0.01 \times 0.99} \approx 0.50, demonstrating how the low tempers the 's evidential strength to avoid overconfidence.

Informative Priors

Strong Priors

Strong priors, also known as highly informative priors, are probability distributions characterized by high concentration around specific values, typically featuring low variance or precision parameters that allow them to dominate the posterior distribution, particularly in scenarios with sparse data. For instance, a normal prior distribution N(\mu_0, \sigma^2) with a small \sigma^2 places substantial weight near \mu_0, effectively constraining the posterior mean toward this value even with limited observations. This concentration reflects strong expert beliefs or accumulated evidence from prior studies, enabling the prior to act as a robust anchor in Bayesian updating. The primary advantage of strong priors lies in their utility for small-sample studies or when reliable is available, as they incorporate substantive to improve estimation precision and reduce . In clinical trials, for example, historical data from previous experiments can inform a strong on treatment effects, allowing efficient borrowing of to enhance without requiring large new samples. This approach is particularly beneficial in pharmaceutical research, where a strong derived from prior studies on —such as a normal centered on an expected response rate from Phase II trials—can shift the posterior toward the prior mean when new Phase III data is limited, leading to more stable inferences about . However, strong priors carry the disadvantage of introducing if misspecified, as their dominant can skew the posterior away from the true value, potentially leading to misleading conclusions. analyses are essential to assess how posterior inferences change under prior perturbations, highlighting the need for careful validation against expertise. As a milder , weakly informative priors can provide regularization with less risk of overriding .

Weakly Informative Priors

Weakly informative priors are probability distributions that incorporate a minimal amount of prior knowledge, featuring broad spreads and some structural constraints to ensure computational stability and reasonable posterior inferences without substantially overriding the data. These priors typically employ heavy-tailed distributions such as the Cauchy or Student's t-distribution with low degrees of freedom and large scale parameters, which center around plausible values like zero for regression coefficients while allowing for extreme outcomes if supported by the evidence. For instance, a Cauchy distribution with location 0 and scale 2.5 (or 10 for intercepts) bounds the tails to prevent implausibly large parameter values, yet remains diffuse enough to let the likelihood dominate in most scenarios. The primary purpose of weakly informative priors is to mitigate pathological issues in , such as improper posteriors or infinite variance that can arise from fully non-informative alternatives, while staying close to objectivity by exerting only light regularization. They stabilize estimates in challenging settings like small sample sizes, high-dimensional models, or cases of parameter non-identifiability (e.g., complete separation in ), where the data alone might yield unstable or extreme results. By introducing just enough structure—such as finite variance and tail decay—these priors promote robust modeling without assuming strong domain-specific beliefs, making them suitable as default choices for exploratory analyses or when prior is difficult. Weakly informative priors gained prominence in the through the advocacy of and collaborators, who emphasized their role in robust Bayesian via tools like and detailed methodological guidance. In their seminal work, Gelman et al. recommended these priors for hierarchical and regression models to balance flexibility and reliability, influencing their adoption in fields requiring reproducible inference. A representative example occurs in , where a normal prior on coefficients with mean 0 and a large standard deviation (e.g., 10) provides mild shrinkage toward zero, regularizing the model against multicollinear predictors while permitting data-driven deviations for truly important effects—thus avoiding the pitfalls of flat priors that can lead to erratic predictions.

Non-Informative Priors

Objective Priors

Objective priors in are probability distributions selected to exert minimal influence on the posterior distribution, thereby allowing the observed to predominantly determine the . These priors aim to represent a state of ignorance or objectivity regarding the parameter values, ensuring that the posterior closely approximates the normalized likelihood when sufficient are available. A prominent example is the , derived as the square root of the determinant of the matrix, \pi(\theta) \propto \sqrt{\det \mathbf{I}(\theta)}, where \mathbf{I}(\theta) quantifies the amount of information the data provide about \theta. This construction, originally proposed by , ensures invariance under smooth reparameterizations of the model, meaning the prior transforms appropriately to maintain the same inferential properties regardless of how the parameter is expressed. For instance, in a with known variance, the Jeffreys prior for the \mu is , \pi(\mu) \propto 1, while for the standard deviation \sigma with known mean, it is \pi(\sigma) \propto 1/\sigma. Another key type is the uniform prior, often used for bounded parameters to express uniformity over the possible range. For a proportion parameter \theta in a binomial model, a uniform prior on [0,1] corresponds to a Beta(1,1) distribution, which integrates to 1 and results in a posterior that is simply the likelihood normalized over the parameter space. If y successes are observed in n trials, the posterior becomes Beta(1 + y, 1 + n - y), directly reflecting the data's evidential content without additional prior weighting. Reference priors extend this objectivity asymptotically, maximizing the expected Kullback-Leibler divergence between the prior and the posterior to ensure the prior adds the least possible information relative to the data. Developed by José M. Bernardo, these priors coincide with Jeffreys priors in one-dimensional cases but provide a more robust approach for multiparameter models by prioritizing parameters of interest. This property makes reference priors particularly suitable for achieving consistent inference as sample sizes grow, preserving the data's dominance in the limit.

Improper Priors

Improper priors are probability distributions that do not integrate to a finite over their , meaning ∫ p(θ) dθ = ∞, rendering them non-normalizable as formal probability densities. Classic examples include the over the entire real line, (-∞, ∞), and the prior proportional to 1/θ for positive scale parameters θ > 0, both of which assign equal weight across unbounded spaces but fail to sum to unity. Despite their mathematical impropriety, these priors can serve as limiting cases of proper distributions with increasingly diffuse , facilitating non-informative Bayesian . For to be valid, an improper must yield a proper , which requires that the integral ∫ p(data|θ) p(θ) dθ remains finite and normalizable, ensuring the posterior integrates to 1. This condition holds when the likelihood p(data|θ) sufficiently bounds the parameter space, dominating the 's divergence. A prominent example is the Haldane , (0,0), which is proportional to p^{-1}(1-p)^{-1} for a success probability p ∈ (0,1) and is improper due to singularities at the boundaries. When combined with data showing at least one success and one failure, the resulting (n+0, m+0) posterior—where n and m are the counts—is proper and equivalent to the maximum likelihood estimate, highlighting the 's utility in objective settings. These priors offer computational simplicity, as they often lead to analytically tractable posteriors without imposing subjective beliefs, making them appealing for default analyses. However, risks arise if the posterior remains improper, which can occur with insufficient data or ill-posed models, leading to paradoxes such as undefined marginal likelihoods that invalidate model comparisons. Careful verification of posterior propriety is essential to avoid misleading inferences. In , a flat improper on the coefficients β, such as p(β) ∝ 1, paired with a similar on the 1/σ² ∝ 1/σ², exemplifies these dynamics. Without data, the posterior stays improper, reflecting the model's underidentification. In contrast, with sufficient observations, the likelihood regularizes the posterior into a proper multivariate normal for β (conditional on σ²) and inverse-gamma for σ², enabling standard Bayesian estimates akin to but with . This setup underscores how data can salvage inference from improper priors in well-posed problems.

Prior Selection

Elicitation Techniques

Elicitation techniques for prior probabilities involve systematic processes to incorporate expert knowledge or existing data into prior distributions within Bayesian analysis. One primary method is expert elicitation through structured questionnaires, which quantify subjective beliefs from domain specialists. The , for instance, facilitates this by conducting iterative rounds of anonymous surveys where experts provide probability assessments, followed by feedback on group responses to converge toward consensus and reduce individual biases. This approach is particularly useful when direct data is scarce, allowing experts to express uncertainties via quantiles or intervals that can be aggregated into a prior distribution. Another technique aggregates historical data to inform priors, drawing on past observations or similar studies to construct distributions that reflect accumulated evidence. This often employs hierarchical models to borrow strength across related datasets, ensuring the prior captures patterns without overfitting to any single source. Empirical Bayes methods further refine this by using preliminary data from past datasets to estimate hyperparameters of the prior, treating the marginal likelihood as a basis for selecting a data-driven yet regularized distribution. These data-informed approaches balance objectivity with the need for prior specification in new analyses. Formal approaches to often encode elicited beliefs into distributional moments, such as mean and variance, before fitting a family like the normal or to match those characteristics. For continuous parameters, experts might provide judgments on expected values and spreads, which are then used to parameterize the via maximum likelihood or moment-matching techniques. In discrete cases, such as success probabilities, distributions are commonly fitted to elicited quantiles or ratios expressed by experts. Challenges in these techniques include avoiding cognitive biases, such as overconfidence or anchoring, which can distort elicited probabilities and lead to overly narrow priors. Handling is also critical, as experts may struggle to quantify second-order uncertainties, necessitating robust aggregation methods to propagate variability into the final prior.00175-9/pdf) Protocols like the SHELF framework address this by incorporating feedback loops and sensitivity checks during elicitation. A representative example involves eliciting a for earthquake magnitude from seismologists, where experts provide judgments on expected magnitudes in a . These assessments are then fitted to a to capture the skewed nature of seismic events, yielding a that informs probabilistic models.

Conjugate Priors

In , a for a \theta is a p(\theta) such that, when combined with a likelihood p(x|\theta) via , the resulting posterior p(\theta|x) belongs to the same distributional family as the prior. This conjugacy ensures analytical tractability, as the posterior can be obtained by simply updating the prior's hyperparameters based on the observed data. The formalization of conjugate priors traces back to Raiffa and Schlaifer (1961), who emphasized their role in decision-theoretic contexts where sufficient statistics of fixed dimension enable closed-form solutions. Conjugate priors are most naturally defined for likelihoods from exponential families, where the prior is constructed to mimic the likelihood's . Key examples include the as conjugate to the or likelihood for modeling success probabilities, the conjugate to the likelihood for rates, and the normal distribution conjugate to the normal likelihood for mean estimation with known variance. For multivariate settings, the serves as the conjugate prior for the of a multivariate normal likelihood. These pairs are summarized in the following table of common conjugate relationships:
Likelihood ModelParameter(s)Conjugate Prior FamilyHyperparameters
Bernoulli/Binomialp\alpha > 0, \beta > 0
\lambdaGamma\alpha > 0, \beta > 0
(known variance)\muMean m, precision \rho
(known mean)\sigma^2Inverse-gammaShape \alpha, scale \beta
Multivariate normal\SigmaInverse-Wishart \nu, scale S
The primary benefit of conjugate priors is the availability of closed-form posterior distributions, which simplifies inference by avoiding numerical integration. For instance, consider a beta prior \text{Beta}(\alpha, \beta) for the success probability p of a binomial likelihood with n trials and s successes; the posterior is then \text{Beta}(\alpha + s, \beta + n - s), where the hyperparameters are updated additively to incorporate the data as pseudo-observations. This updating rule extends analogously to other pairs, such as incrementing the gamma shape parameter by the sum of Poisson counts. Such properties make conjugate priors computationally efficient for sequential updating and exact inference in simple models. Despite their advantages, conjugate priors offer limited flexibility in capturing nuanced prior beliefs, as they constrain the posterior form to match the prior family regardless of the data's implications. They are also unavailable for many complex or non-exponential family models, restricting their applicability. The development of (MCMC) methods has largely alleviated these limitations by enabling approximate posterior sampling without requiring conjugacy, thereby supporting more general and realistic priors in modern Bayesian analysis.

Applications

In Statistical Mechanics

In statistical mechanics, prior probabilities are conceptualized as initial distributions over the microstates of a , representing the state of incomplete before incorporating constraints such as or . These priors serve as starting points for Bayesian updating, where the likelihood arises from the system's or energy constraints, leading to posterior distributions that describe ensembles. This adaptation frames as a of refining uncertainty about microstate probabilities based on macroscopic observables, ensuring that the resulting distributions are consistent with thermodynamic principles. A key connection emerges through Edwin Jaynes' formulation of the maximum entropy principle, which posits that the prior probability distribution should maximize informational entropy—subject to known constraints—to remain as uninformative as possible beyond the given data. For instance, in systems with equal a priori probabilities over accessible microstates (a uniform prior), the maximum entropy distribution under a fixed average energy constraint yields the canonical ensemble form, where probabilities are proportional to e^{-\beta E_i}, with \beta = 1/(kT) and E_i the energy of microstate i. This approach justifies the Boltzmann distribution as the least biased inference from partial knowledge, such as the system's total energy, rather than assuming it a priori from physical laws. An illustrative example is the derivation of the Maxwell-Boltzmann velocity distribution for an . Starting with a uniform prior over velocity space (maximizing without constraints), Bayesian updating incorporates the constraint of fixed average per particle, \langle \frac{1}{2} m v^2 \rangle = \frac{3}{2} [kT](/page/KT). The resulting posterior distribution is the Maxwell-Boltzmann form: f(v) dv = 4\pi v^2 \left( \frac{m}{2\pi [kT](/page/KT)} \right)^{3/2} \exp\left( -\frac{m v^2}{2 [kT](/page/KT)} \right) dv, which predicts the observed spread of molecular speeds in . This Bayesian perspective highlights how the distribution emerges from logical inference rather than dynamical assumptions alone. Historically, J. Willard Gibbs' ensemble method, introduced in 1902, can be reinterpreted in Bayesian terms as assigning prior probabilities uniformly across an of imagined systems to represent epistemic about a single system's . Gibbs' , with probabilities weighted by energy, aligns with maximum entropy priors updated by energy constraints, providing a foundational bridge between classical and modern inferential methods.

In

In , prior probabilities play a crucial role in Bayesian models by encoding assumptions about function forms and parameter behaviors, enabling regularization and . In Gaussian processes (GPs), the prior is typically specified as a multivariate Gaussian over functions, where the function determines the smoothness and structure, promoting interpretable predictions with built-in uncertainty estimates. This approach contrasts with frequentist methods by integrating the prior directly into inference, allowing GPs to model non-linear relationships while avoiding through the posterior's contraction around observed data. Similarly, in Bayesian neural networks (BNNs), priors are placed on network weights, often Gaussian distributions that induce regularization akin to weight decay, which penalizes large weights to improve generalization on unseen data. Hierarchical priors extend this framework by structuring priors across multiple levels to share information among related data groups, facilitating scalable modeling of complex dependencies. A prominent example is the (LDA) model for topic discovery in text corpora, where are applied hierarchically: a top-level governs the topic proportions for each document, while another governs word distributions within topics, enabling the model to capture shared thematic structures across documents without assuming a fixed number of topics a priori. This multi-level setup allows posterior inference via techniques like , promoting coherent topic assignments and reducing sensitivity to initialization. Priors also mitigate by imposing sparsity or smoothness constraints, with practical impacts seen in high-dimensional settings. For instance, the Laplace on coefficients, as in the Bayesian Lasso, encourages many coefficients to shrink to zero, effectively performing automatic while providing credible intervals for uncertainty. In variational autoencoders (VAEs), on latent variables—commonly standard Gaussian—enforce structured representations during training; the (ELBO) objective balances reconstruction fidelity with a Kullback-Leibler (KL) term that regularizes the approximate posterior toward the , yielding disentangled latent spaces suitable for generative tasks. These applications highlight ' role in enhancing model robustness and interpretability in modern pipelines.

References

  1. [1]
    Prior probability - StatLect
    The prior probability is the probability assigned to an event before the arrival of some information that makes it necessary to revise the assigned probability.Bayes' rule · Example · Priors in Bayesian statistics · Prior distribution
  2. [2]
    [PDF] Bayes' Theorem • Prior probability; posterior prob - UBC Zoology
    Probability is a measure of a degree of belief associated with the occurrence of an event. A probability distribution is a list of all mutually exclusive events ...
  3. [3]
    What is Bayesian Analysis?
    What we now know as Bayesian statistics has not had a clear run since 1763. Although Bayes's method was enthusiastically taken up by Laplace and other leading ...
  4. [4]
    [PDF] The Development of Bayesian Statistics - Columbia University
    Jan 13, 2022 · Abstract. The incorporation of Bayesian inference into practical statistics has seen many changes over the past century, including ...
  5. [5]
    Bayes for Beginners 2: The Prior
    Sep 30, 2015 · A prior distribution assigns a probability to every possible value of each parameter to be estimated.
  6. [6]
    [PDF] Lecture 20 — Bayesian analysis 20.1 Prior and posterior distributions
    random variable Θ having a probability distribution fΘ(θ), called the prior distribution. This distribution represents our prior belief about the value of this ...
  7. [7]
    [PDF] Prior distribution - Columbia University
    The prior distribution is a key part of Bayesian infer- ence (see Bayesian methods and modeling) and rep- resents the information about an uncertain parameter ...
  8. [8]
    [PDF] The Case for Objective Bayesian Analysis - Stat@Duke
    One of the mysteries of modern Bayesianism is the lip service that is often paid to subjective Bayesian analysis as opposed to objective Bayesian analysis, but ...
  9. [9]
    [PDF] Objective Bayesian Statistical Inference - Blogs at Kent
    In subjective Bayesian analysis, prior distributions for θ, π(θ), represent beliefs, which change with data via Bayes theorem. In objective Bayesian ...
  10. [10]
    [PDF] Noninformative prior probability density for a probability - Statistics
    In his Théorie Analytique des Probabilités of 1812, which set the stage for modern statistical inference, Laplace invoked what is now known as the principle of ...Missing: insufficient | Show results with:insufficient
  11. [11]
    23.2 - Bayesian Estimation | STAT 415 - STAT ONLINE
    In the case where the parameter space for a parameter θ takes on an infinite number of possible values, a Bayesian must specify a prior probability density ...<|control11|><|separator|>
  12. [12]
    [PDF] 6 Bayesian Parameter Estimation
    If we were completely ignorant about our coin-like object, what distribution would capture our state of ignorance? At first glance, it might appear obvious that ...
  13. [13]
    Bayes' Theorem - Stanford Encyclopedia of Philosophy
    Jun 28, 2003 · Bayes' Theorem is a simple mathematical formula used for calculating conditional probabilities. It figures prominently in subjectivist or Bayesian approaches ...Conditional Probabilities and... · Special Forms of Bayes...Missing: primary source
  14. [14]
    [PDF] Bayesian statistics and modelling - Columbia University
    The prior distribution (blue) and the likelihood function. (yellow) are combined in Bayes' theorem to obtain the posterior distribution (green) in our ...
  15. [15]
    Bayesian epistemology - Stanford Encyclopedia of Philosophy
    Jun 13, 2022 · Bayesian epistemology has a long history. Some of its core ideas can be identified in Bayes' (1763) seminal paper in statistics (Earman 1992: ch ...
  16. [16]
    [PDF] Chapter 12 Bayesian Inference - Statistics & Data Science
    Bayesian inference is appealing when prior information is available since Bayes' theorem is a natural way to combine prior information with data. However ...<|separator|>
  17. [17]
    Medical Tests — Bite Size Bayes
    The mistake they are making is called the base rate fallacy because it ignores the “base rate” of the condition, which is the prior.
  18. [18]
    [PDF] Bayesian Data Analysis, Third Edition
    ... strong prior information, DIC gives nonsensical results when the posterior ... high concentration, and estimated to cause several thousand lung cancer ...
  19. [19]
    [PDF] Bayesian Approaches to Clinical Trials and Health-Care Evaluation
    ... strong prior opinion against H0 (prior odds of 1/999) does not lead to a ... low variance. This may be reasonable behaviour but should be acknowledged ...
  20. [20]
    A weakly informative default prior distribution for logistic and other ...
    Andrew Gelman. Aleks Jakulin. Maria Grazia Pittau. Yu-Sung Su. "A ... prior distribution , weakly informative prior distribution. Rights: Copyright ...
  21. [21]
    [PDF] Bayesian Data Analysis Third edition (with errors fixed as of 20 ...
    ... Bayesian Data Analysis. Third edition. (with errors fixed as of 20 ... Weakly informative priors for variance parameters. 128. 5.8 Bibliographic note.
  22. [22]
    [PDF] Lecture 7: Jeffreys Priors and Reference Priors - People @EECS
    Feb 17, 2010 · The idea behind reference priors is to formalize what exactly we mean by an “uninformative prior”: it is a function that maximizes some measure ...
  23. [23]
    [PDF] Objective priors - Applied Bayesian Analysis
    ▷ The Jeffreys prior for parameter θ is p(θ) = pI(θ). ▷ The Fisher ... ▷ Reference priors are harder to compute than Jeffreys. 16 / 20. Page 17 ...
  24. [24]
    [PDF] Harold Jeffreys's Theory of Probability Revisited - arXiv
    Jeffreys recalls Laplace's rule that, if a parameter is real-valued, its prior probability should be taken as uniformly distributed, while, if this parameter ...
  25. [25]
    [PDF] Prior distributions - MRC Biostatistics Unit |
    The appropriate specification of priors that contain minimal information is an old problem in Bayesian statistics: the terms “objective” and “reference” are ...<|control11|><|separator|>
  26. [26]
    [PDF] Reference Posterior Distributions for Bayesian Inference
    Thus, using (5), the reference prior for @ is uniform and therefore, using Bayes' theorem, the reference posterior distribution π(0|z) is a uniform distribution.
  27. [27]
    Improper Prior - an overview | ScienceDirect Topics
    An improper prior is defined as a probability density function for which the integral over its entire domain is infinite, i.e., ∫ Θ π(θ) dθ = ∞. Despite being ...
  28. [28]
    Improper priors and improper posteriors - Wiley Online Library
    Jul 14, 2021 · We consider a theoretical framework for statistics that includes both improper priors and improper posteriors.
  29. [29]
    J. B. S. Haldane's Contribution to the Bayes Factor Hypothesis Test
    In Haldane's example, this means to fo- cus on estimating the cross-over rate parameter, using relevant real-world knowledge of the problem to con- struct a ...
  30. [30]
    [PDF] The selection of prior distributions by formal rules
    pressing ignorance." Despite Jeffreys's belief in an "initial" stage at which an investigator is ignorant, and his application of insufficient reason ...
  31. [31]
    [PDF] Bayesian Inference in the Linear Regression Model
    A standard “default” procedure is to place a non-informative. (improper) prior on (β,σ2). The first step in this regard is to assume prior independence between.
  32. [32]
    Recommendations on the Use of Structured Expert Elicitation ...
    The Delphi method uses repeated cycles of feedback and individual elicitation. At each iteration, experts are provided with summary information about how their ...
  33. [33]
    Prior Knowledge Elicitation: The Past, Present, and Future
    Prior elicitation transforms domain knowledge into well-defined prior distributions for Bayesian models, rather than assuming the analyst specifies them ...<|control11|><|separator|>
  34. [34]
    [PDF] Methods for Eliciting Informative Prior Distributions - arXiv
    Mar 11, 2022 · Bayesian Method: This approach treats each expert's prior as new data. The analyst updates their prior with the “new data". The resulting ...
  35. [35]
    Statistical Methods for Eliciting Probability Distributions
    Prior Elicitation, Variable Selection and Bayesian Computation for Logistic Regression Models ... The Assessment of Prior Distributions in Bayesian Analysis.
  36. [36]
  37. [37]
    What Do We Know Without the Catalog? Eliciting Prior Beliefs from ...
    Oct 18, 2024 · An alternative way to quantify parameter values for a seismic region is by eliciting expert opinions on the seismological characteristics that ...
  38. [38]
    [PDF] A Compendium of Conjugate Priors - Applied Mathematics Consulting
    The definition and construction of conjugate prior distributions depends on the existence and identification of sufficient statistics of fixed dimension for the ...
  39. [39]
    [PDF] The Multivariate Distributions: Normal and inverse Wishart
    ▷ The multivariate normal (MVN) distribution. ▷ Conjugate for the MVN distribution. ▷ The inverse Wishart distribution. ▷ Conjugate for the MVN distribution ( ...
  40. [40]
    [PDF] Conjugate priors: Beta and normal Class 15, 18.05
    Conjugate priors are useful because they reduce Bayesian updating to modifying the parameters of the prior distribution (so-called hyperparameters) rather than ...Missing: statistics | Show results with:statistics
  41. [41]
    [PDF] Re-examining informative prior elicitation through the lens of MCMC
    Conjugate priors, long the workhorse of classic methods for eliciting informative priors, have their roots in a time when modern computational methods were ...Missing: limitations | Show results with:limitations<|control11|><|separator|>
  42. [42]
    [PDF] Prior Probabilities
    The principle of maximum entropy (i.e., the prior probability assignment should be the one with the maximum entropy consistent with the prior knowledge) ...
  43. [43]
    Gaussian Processes for Machine Learning - MIT Press Direct
    Christopher K. I. Williams is Professor of Machine Learning and ... Open the PDF Link PDF for Appendix B: Gaussian Markov Processes in another window.
  44. [44]
    [PDF] Bayesian Learning for Neural Networks - University of Toronto
    Jul 23, 1996 · Bayesian Learning for Neural Networks is no longer available on-line. A revised version with some new material has now been published by ...
  45. [45]
    [PDF] Latent Dirichlet Allocation - Journal of Machine Learning Research
    LDA is a three-level hierarchical Bayesian model, in which each item of a collection is modeled as a finite mixture over an underlying set of topics. Each ...
  46. [46]
    [PDF] The Bayesian Lasso - People @EECS
    Trevor Park is Assistant Professor (E-mail: tpark@stat.ufl.edu) and George. Casella is Distinguished Professor (E-mail: casella@stat.ufl.edu), Department of ...
  47. [47]
    [PDF] Auto-encoding variational bayes - arXiv
    Dec 10, 2022 · Auto-Encoding Variational Bayes. Diederik P. Kingma. Machine Learning Group. Universiteit van Amsterdam dpkingma@gmail.com. Max Welling.