Fact-checked by Grok 2 weeks ago

Maximum a posteriori estimation

Maximum a posteriori () estimation is a Bayesian statistical method for obtaining a point estimate of an unknown by selecting the value that maximizes the , given observed and a . Formally, for a \theta and x, the estimate \hat{\theta}_{\text{[MAP](/page/Map)}} is defined as \hat{\theta}_{\text{[MAP](/page/Map)}} = \arg\max_{\theta} \pi(\theta \mid x), where \pi(\theta \mid x) is the posterior density derived from : \pi(\theta \mid x) = \frac{L(x \mid \theta) \pi(\theta)}{P(x)}, with L(x \mid \theta) as the likelihood, \pi(\theta) as the , and P(x) as the (which is constant for maximization purposes). This approach incorporates prior knowledge about \theta into the estimation process, distinguishing it from (), which solely maximizes the likelihood L(x \mid \theta) and equates to only under a . The roots of MAP estimation trace back to the development of in the , with foundational contributions from in 1763, who introduced , and in 1774, who advanced by minimizing posterior loss functions—a direct precursor to MAP. It gained prominence in the through the neo-Bayesian revival, influenced by figures like (1939) and Leonard J. Savage (1954), who formalized subjective priors and decision-theoretic frameworks that solidified MAP as a practical tool for parameter estimation. Unlike fully Bayesian methods that integrate over the posterior for , MAP provides a single mode-based estimate, making it computationally efficient while still leveraging priors to regularize estimates, especially in scenarios with limited data. MAP estimation is widely applied across fields requiring robust parameter inference under uncertainty. In machine learning, it underpins for incorporating prior beliefs on coefficients and enhances Naive Bayes classifiers by optimizing class probabilities with priors. In signal processing and , MAP addresses ill-posed inverse problems, such as reconstructing images from noisy measurements by maximizing the posterior under log-concave models. Other notable uses include sequence estimation for tasks like , pharmacokinetics for dosing optimization via population models, and for identifying system parameters from data. These applications highlight MAP's versatility in balancing data-driven likelihood with expert prior knowledge to yield reliable estimates.

Foundations

Bayesian Probability

Bayesian probability interprets probability as a measure of or degree of uncertainty in a , rather than as a long-run frequency of events. In this framework, initial beliefs about unknown parameters are quantified through a distribution, which is then updated with observed data to form a revised , known as the distribution. This updating process embodies the core principle of : learning from evidence by coherently combining knowledge with new information using the rules of probability. In contrast, the frequentist approach treats parameters as fixed but unknown constants and defines probability based on the limiting frequency of events in repeated trials under identical conditions, without incorporating subjective prior beliefs. Frequentist methods focus on testing and intervals derived from sampling distributions, whereas Bayesian methods emphasize the full posterior distribution to quantify and make probabilistic statements about parameters directly. This subjective allows Bayesian to flexibly integrate domain-specific , making it particularly useful in scenarios with limited or strong priors. The distribution represents the updated belief about the parameters after observing the , serving as the foundation for all subsequent inferences in . It balances the influence of the prior distribution—reflecting preconceived notions—and the , which encodes how well the support different parameter values, to produce a coherent synthesis of information. The origins of trace back to , a Presbyterian minister and mathematician, who formulated the key theorem in his posthumously published 1763 essay "An Essay towards Solving a Problem in the Doctrine of Chances," communicated by . This work addressed , laying the groundwork for updating beliefs based on evidence. The theorem was independently rediscovered and popularized in the early by mathematician in his 1774 memoir "Mémoire sur la probabilité des causes des événements," which applied it to problems in astronomy and . At the heart of this framework is the informal statement of , which posits that the is proportional to the product of the likelihood and the : p(\theta \mid x) \propto p(x \mid \theta) \cdot p(\theta) Here, the posterior distribution of the parameters \theta given data x is directly informed by the likelihood p(x \mid \theta) and the prior p(\theta), with the ensuring the result integrates to 1.

Prior and Likelihood Concepts

In , the prior distribution specifies the over the model parameters \theta prior to observing the data, encapsulating existing knowledge, beliefs, or uncertainty about \theta. This distribution serves as a foundational component for updating beliefs in light of new evidence. The , denoted L(\theta \mid x) = P(x \mid \theta), represents the probability of the observed data x given the parameters \theta, treating \theta as fixed while varying to assess model fit. Prior distributions are classified into types that facilitate different inferential goals, notably conjugate priors and non-informative priors. Conjugate priors are chosen such that the resulting posterior distribution belongs to the same family as the , enabling closed-form updates and computational tractability. For example, the is conjugate to the likelihood, where the prior on the success probability p remains Beta after incorporating binary data. Non-informative priors, such as the uniform prior or , minimize the influence of prior assumptions, allowing the data to primarily drive the posterior. The , specifically, is defined as proportional to the of the of the matrix, promoting invariance under parameter transformations. Priors enable the integration of domain-specific knowledge by assigning higher probabilities to parameter values consistent with expert insights or historical data, effectively acting as a regularization mechanism. This regularization constrains the parameter space, reducing the risk of overfitting by penalizing extreme or implausible values that might otherwise fit noise in limited datasets. An illustrative case of conjugate prior updating occurs in the Beta-Binomial framework. Suppose a Beta() prior is placed on the success probability p for a Binomial likelihood with n trials and k observed successes; the posterior then becomes Beta(\alpha + k, \beta + n - k), seamlessly incorporating the data into the prior parameters. p(\theta \mid x) = \text{Beta}(\alpha + k, \beta + n - k)

Formulation

Bayes' Theorem Application

Bayes' theorem provides the foundational framework for , including maximum a posteriori (MAP) estimation, by updating prior beliefs about model in light of observed . In the context of estimating an unknown \theta given x, the theorem expresses the posterior as p(\theta \mid x) = \frac{p(x \mid \theta) \, p(\theta)}{p(x)}, where p(\theta \mid x) represents the posterior probability density of \theta given x. The numerator consists of two key components: p(x \mid \theta), the that quantifies how well the x fits the parameter \theta, and p(\theta), the prior distribution encoding beliefs about \theta before observing the . The denominator, p(x), known as the or evidence, serves as a that ensures the posterior integrates to 1, thereby forming a valid over \theta. This evidence is computed as the p(x) = \int p(x \mid \theta) \, p(\theta) \, d\theta, which marginalizes over all possible values of \theta to yield the predictive probability of the data under the prior. While the evidence plays a crucial role in normalizing the posterior, its computation poses significant challenges, particularly in high-dimensional parameter spaces where the integral becomes intractable due to the exponential growth in integration complexity. In such cases, the high dimensionality exacerbates the difficulty of evaluating the marginal likelihood accurately, often requiring specialized techniques to handle the curse of dimensionality.

MAP Objective Function

The maximum (MAP) is defined as the value of the \theta that maximizes the posterior P(\theta | x), where x denotes the observed . This point estimate identifies the of the posterior, providing a single most probable value for \theta given both the and prior knowledge. Formally, the MAP is given by \theta_{\text{MAP}} = \arg\max_{\theta} P(\theta | x), which, by application of from , is proportional to the product of the likelihood and the prior density. For computational convenience, the maximization is typically performed on the log-posterior, which is additive and avoids underflow issues in high dimensions. Ignoring the constant normalizing factor from the , this yields \theta_{\text{MAP}} = \arg\max_{\theta} \left[ \log P(x | \theta) + \log P(\theta) \right] = \arg\max_{\theta} \left[ \log L(\theta | x) + \log \pi(\theta) \right], where L(\theta | x) is the and \pi(\theta) is the density. Equivalently, the MAP solution can be obtained by minimizing the negative log-posterior, -\log P(\theta | x), which reframes the problem as an optimization task. In this formulation, the prior term \log \pi(\theta) acts as a regularization penalty that constrains the space, penalizing values of \theta that deviate strongly from prior beliefs and thereby promoting solutions that are both data-driven and informed by expert knowledge. The intuition behind the MAP objective lies in its role as a compromise between fitting the observed and adhering to preconceived notions about the parameters. The likelihood term \log L(\theta | x) emphasizes fidelity to the , favoring parameters that make the observations most probable, while the term \log \pi(\theta) incorporates subjective or empirical beliefs to mitigate , especially in scenarios with limited . This balance is particularly evident in models where the is chosen as a Gaussian distribution, leading to regularization in the negative log-posterior, or a , resulting in L1 regularization akin to penalties. Overall, MAP estimation thus yields a robust point estimate that leverages the full Bayesian framework without requiring integration over the entire posterior.

Estimation Techniques

Closed-Form Solutions

Closed-form solutions for maximum a posteriori () estimation are feasible when the prior distribution is conjugate to the , ensuring that the posterior distribution belongs to the same parametric family as the prior and allowing the to be found analytically without numerical optimization. This conjugacy simplifies the computation of the estimate, which is the of the posterior, by leveraging closed-form expressions for the posterior parameters. A prominent example occurs in the estimation of the mean of a univariate Gaussian with known variance, where both the likelihood and are Gaussian, forming a conjugate pair. Consider x drawn from \mathcal{N}(\theta, \sigma^2) (likelihood) and a \theta \sim \mathcal{N}(\mu_0, \sigma_0^2). The posterior is also Gaussian, \mathcal{N}(\mu_n, \sigma_n^2), with \sigma_n^2 = \left( \frac{1}{\sigma^2} + \frac{1}{\sigma_0^2} \right)^{-1}, \quad \mu_n = \frac{ x / \sigma^2 + \mu_0 / \sigma_0^2 }{ 1 / \sigma^2 + 1 / \sigma_0^2 }. The MAP estimate is \theta_{\text{MAP}} = \mu_n, representing a precision-weighted average of the data x and prior mean \mu_0, where the weights are the inverse variances of the likelihood and prior, respectively. This form highlights how the prior regularizes the estimate toward \mu_0 with strength proportional to $1/\sigma_0^2. Analytical solutions become unavailable in non-conjugate settings, where the posterior does not retain a tractable form, or when the posterior is multimodal, complicating the identification of the global mode without approximation techniques. In such cases, the MAP estimate requires numerical methods to evaluate the objective function defined by .

Optimization Algorithms

When closed-form solutions for the maximum (MAP) estimate are intractable due to complex priors or likelihoods, numerical optimization algorithms are employed to maximize the log-posterior objective function. These methods iteratively update estimates to find a local of the posterior , often leveraging the differentiability of the log-posterior for efficient . An illustrative non-conjugate case is a Gaussian likelihood with a on the parameters, which induces L1 regularization in the MAP objective and forms the Bayesian foundation for the method in . For observations y = X\theta + \epsilon where \epsilon \sim \mathcal{N}(0, \sigma^2 I), and \theta following a p(\theta) \propto \exp(-\lambda \|\theta\|_1 / \sigma^2), the negative log-posterior yields the \theta_{\text{MAP}} = \arg\min_\theta \frac{1}{2\sigma^2} \|y - X\theta\|_2^2 + \frac{\lambda}{2\sigma^2} \|\theta\|_1. This closed-form objective, while not always solvable analytically for arbitrary X, directly connects the Laplace prior's sparsity-promoting properties to the L1 penalty in regression, enabling alongside estimation. Gradient-based methods form a cornerstone of MAP optimization, particularly for continuous spaces. ascent directly maximizes the log-posterior by iteratively updating parameters in the direction of the , with step sizes controlled by learning rates or techniques to ensure . For large datasets where computing the full is prohibitive, stochastic ascent variants approximate the gradient using mini-batches of data, enabling scalable optimization while introducing beneficial noise that aids escape from poor local modes. These approaches are widely adopted in Bayesian neural networks and generalized linear models, where the log-posterior's smoothness supports reliable . In latent variable models, the expectation-maximization (EM) algorithm can be adapted for MAP estimation by incorporating the prior into the maximization step. The standard EM alternates between an expectation step that computes a lower bound on the log-posterior using current parameter estimates and a maximization step that updates parameters to increase this bound, but for MAP, the maximization explicitly includes the log-prior term to penalize deviations from prior beliefs. This adaptation, as detailed by Neal and Hinton, justifies variants like incremental EM for online settings, where updates occur after each data point, enhancing efficiency for streaming Bayesian inference. Such modifications are particularly effective in mixture models and hidden Markov models, where latent structures complicate direct optimization. Variational inference provides an framework that indirectly targets the posterior through mean-field assumptions, where the joint posterior is factorized into independent distributions over parameters. By optimizing a variational lower bound on the log-evidence (, ELBO), the of the approximating distribution aligns closely with the true posterior under certain conjugacy conditions, offering a scalable alternative to direct maximization. This method, pioneered in graphical models, uses coordinate ascent or gradient-based updates on the variational parameters, balancing computational tractability with quality. It excels in high-dimensional settings, such as topic models, by avoiding exhaustive enumeration of the posterior. Multimodal posteriors pose challenges for local optimization methods, as they may converge to suboptimal modes; handling this requires strategies like careful initialization and search techniques. Multiple random initializations followed by local optimization, such as ascent from diverse starting points, increase the likelihood of discovering the mode by exploring different basins of . For more robust exploration, adapts a temperature schedule to the log-posterior, allowing probabilistic acceptance of worse solutions early on to escape local modes, gradually cooling to focus on the maximum—a technique rooted in principles for discrete and continuous MAP problems. This method has proven effective in image restoration and state estimation tasks with rugged posteriors.

Applications and Examples

Parameter Estimation in Models

In probabilistic models, maximum a posteriori () estimation is frequently applied to infer model parameters by maximizing the posterior , incorporating both the likelihood of observed and beliefs about the parameters. A prominent example is its use in , where assuming a Gaussian likelihood for the errors and a Gaussian on the regression coefficients leads to the MAP estimate being equivalent to the ridge solution, which includes an L2 penalty term to regularize the coefficients and mitigate in the presence of . Specifically, for a \mathbf{y} = \mathbf{X}\boldsymbol{\beta} + \boldsymbol{\epsilon} with \boldsymbol{\epsilon} \sim \mathcal{N}(\mathbf{0}, \sigma^2 \mathbf{I}) and \boldsymbol{\beta} \sim \mathcal{N}(\mathbf{0}, \tau^2 \mathbf{I}), the MAP objective is to minimize \|\mathbf{y} - \mathbf{X}\boldsymbol{\beta}\|^2 + \lambda \|\boldsymbol{\beta}\|^2, where \lambda = \sigma^2 / \tau^2 controls the strength of regularization, yielding the closed-form solution \hat{\boldsymbol{\beta}}_{\text{MAP}} = (\mathbf{X}^T \mathbf{X} + \lambda \mathbf{I})^{-1} \mathbf{X}^T \mathbf{y}. In Bayesian networks, MAP estimation serves dual roles in parameter inference and structure learning, particularly for discrete models where conditional probability tables must be estimated from . For parameter inference, given a fixed , MAP uses Dirichlet priors on the parameters of each ; the resulting estimates add pseudocounts to observed frequencies, smoothing the probabilities toward the prior means and improving robustness with limited . For structure learning, MAP approaches maximize the of (DAG) by scoring potential based on fit and prior preferences for simplicity, often employing optimization techniques like hill-climbing to navigate the combinatorial search space. MAP estimation extends to hyperparameter tuning in hierarchical models by treating hyperparameters—such as variance components or regularization strengths—as additional parameters to be inferred from the joint posterior, rather than fixed values. This approach, akin to , maximizes the posterior of the hyperparameters conditional on the data, enabling adaptive regularization that balances model complexity and fit without requiring full posterior sampling. For instance, in regression, MAP hyperparameter optimizes the kernel parameters and noise variance to maximize the marginal posterior, providing a point estimate that approximates the full Bayesian solution efficiently. Practical implementations of MAP estimation for parameter inference in probabilistic models are supported by libraries such as PyMC and , which facilitate optimization-based computation of MAP solutions within broader Bayesian workflows. PyMC, a Python-based framework, integrates MAP via its optimization routines for quick approximations in complex hierarchical models. Similarly, , implemented in C++ with interfaces for multiple languages, employs gradient-based optimizers like L-BFGS to compute MAP estimates, making it suitable for high-dimensional parameter spaces in models like Bayesian networks.

Signal Processing Example

In , a common application of maximum a posteriori (MAP) estimation is in denoising noisy signals, where the goal is to recover the underlying clean signal from corrupted observations. The setup typically involves a noisy observation y = x + n, where x is the true signal of interest, and n is with zero and variance \sigma_n^2. To incorporate knowledge about the signal, a sparse is placed on x, such as the Laplacian distribution p(x) = \frac{\lambda}{2} \exp(-\lambda |x|) for each component (or in vector form promoting \ell_1-sparsity), which is suitable for signals exhibiting sparsity in a transform domain, like piecewise constant or edge-dominated signals in or contexts. The MAP formulation for denoising maximizes the posterior \hat{x} = \arg\max_x p(x|y), which is proportional to the likelihood p(y|x) times the prior p(x). Under the Gaussian noise model, the likelihood is p(y|x) = \frac{1}{(2\pi \sigma_n^2)^{N/2}} \exp\left( -\frac{1}{2\sigma_n^2} \|y - x\|_2^2 \right), where N is the signal dimension. Combining with the Laplacian prior p(x) \propto \exp(-\lambda \|x\|_1), the MAP estimate solves the optimization problem: \hat{x} = \arg\max_x \left[ -\frac{1}{2\sigma_n^2} \|y - x\|_2^2 - \lambda \|x\|_1 \right]. This is equivalent to \ell_2-regularized \ell_1-minimization, known as the LASSO problem in this context, leading to MAP denoising that balances data fidelity with sparsity promotion. For practical computation, especially in the orthogonal wavelet transform domain where sparsity is enhanced, the solution is obtained via element-wise soft-thresholding, which admits a closed-form proximal operator for the \ell_1 term. The step-by-step process is as follows: (1) Apply the forward wavelet transform to the noisy signal y to obtain coefficients \tilde{y} = W y, where W is the orthogonal wavelet matrix, yielding \tilde{y}_k = \tilde{x}_k + \tilde{n}_k for each coefficient index k; (2) For Gaussian noise and Laplacian prior on coefficients, the scalar MAP estimate per coefficient is \hat{\tilde{x}}_k = \sign(\tilde{y}_k) (|\tilde{y}_k| - T)_+, where T = \lambda \sigma_n^2 is the threshold derived from the noise variance \sigma_n^2 and prior rate parameter \lambda (with (a)_+ = \max(a, 0)); this soft-thresholding shrinks coefficients toward zero by T if |\tilde{y}_k| > T, or sets them to zero otherwise; (3) Apply the inverse wavelet transform \hat{x} = W^T \hat{\tilde{x}} to reconstruct the denoised signal. While the per-coefficient operation is direct, the overall algorithm can be iterated across wavelet scales or in non-orthogonal bases using proximal gradient methods for refinement. The resulting MAP denoised signal demonstrates effective by suppressing small-amplitude fluctuations (corresponding to noise-dominated coefficients), while preserving sharp signal features such as edges through retention of large coefficients. Visually, compared to the raw noisy data—which exhibits high-frequency artifacts and elevated variance across the signal—the MAP output shows smoother homogeneous regions and maintained discontinuities at edges, achieving near-optimal performance over Besov classes without introducing spurious oscillations.

Comparisons and Limitations

Versus Maximum Likelihood Estimation

Maximum likelihood estimation (MLE) seeks to find the parameter values \theta that maximize the probability of the observed data, formulated as \hat{\theta}_{\text{MLE}} = \arg\max_{\theta} P(\mathbf{x} | \theta), relying exclusively on the likelihood without incorporating any prior beliefs about the parameters. The primary distinction between MAP and MLE lies in the incorporation of a prior distribution in MAP, which serves as a regularizing term to favor parameter values consistent with existing knowledge or assumptions, thereby reducing the variance of the estimate while potentially introducing a small away from the true value. Asymptotically, under regularity conditions, the MAP estimate converges to the MLE as the dataset size grows large or the prior becomes increasingly diffuse, since the likelihood dominates the posterior; this convergence is underpinned by the Bernstein-von Mises theorem, which establishes that the posterior distribution approximates a centered at the MLE. MAP is advantageous for small sample sizes or scenarios with informative priors that provide meaningful regularization, whereas MLE is more suitable for large datasets where the data itself suffices to inform the parameters reliably without prior influence.

Bias and Consistency Properties

The maximum a posteriori () estimator incorporates prior information, leading to a bias toward the prior or in finite samples, as the posterior mode is pulled away from the maximum likelihood estimate (MLE) by the prior . For instance, in a model with a prior, the MAP estimate lies between the sample proportion and the prior \alpha/(\alpha + \beta), with the bias most pronounced when the sample size is small. However, under standard regularity conditions—such as the true data-generating belonging to the model family—this bias vanishes asymptotically as the sample size n \to \infty, since the likelihood dominates the fixed prior, rendering the MAP estimator asymptotically unbiased. MAP estimators are consistent provided the prior density is positive and continuous in a neighborhood of the true parameter value, ensuring the posterior concentrates around the true parameter as n \to \infty. This requires mild conditions, including a compact parameter space and the true parameter not lying on the boundary, allowing the posterior mode to converge to the true value. In contrast to the MLE, which achieves consistency without priors, the MAP's consistency hinges on the prior's support but aligns asymptotically with the MLE under these assumptions. In finite samples, the shrinkage induced by the often reduces the variance of the MAP estimator compared to the MLE, yielding a lower (MSE) overall, particularly in settings with noisy data or weak signals. For example, in with Cauchy priors, MAP shrinks extreme coefficients (e.g., from 10.2 to 5.4) while narrowing confidence intervals (e.g., from \pm 6.4 to \pm 2.2), demonstrating without substantial in moderate samples. Asymptotically, the MAP achieves the same as the MLE, with MSE approaching the Cramér-Rao lower bound. In high-dimensional settings, the curse of dimensionality complicates specification for estimation, as noninformative priors become difficult to justify and may lead to improper posteriors or failure of consistency when the parameter space expands with n. Sparse data in high dimensions exacerbates this, requiring hierarchical or regularizing priors (e.g., peaked at zero) to mitigate overestimation, though such choices demand careful to avoid undue bias.

References

  1. [1]
    [PDF] 7.5: Maximum A Posteriori Estimation
    7.5: Maximum A Posteriori Estimation. (From “Probability & Statistics with ... Definition 7.5.1: Maximum A Posteriori (MAP) Estimation. Let x = (x1 ...
  2. [2]
    When Did Bayesian Inference Become “Bayesian”? - Project Euclid
    While Bayes' theorem has a 250-year history, and the method of in- verse probability that flowed from it dominated statistical thinking into the twen- tieth ...
  3. [3]
    Maximum A Posteriori (MAP) Estimation - GeeksforGeeks
    Jul 23, 2025 · Applications of MAP Estimation in Machine Learning · 1. Bayesian Regression- · 1. Bayesian Regression- · 2. Naive Bayes Classifier- · 2. Naive Bayes ...
  4. [4]
    Revisiting Maximum-A-Posteriori Estimation in Log-Concave Models
    Maximum-a-posteriori (MAP) estimation is the main Bayesian estimation methodology in imaging sciences, where high dimensionality is often addressed by using ...Missing: sources | Show results with:sources
  5. [5]
    [PDF] On Maximum a Posteriori Estimation of Hidden Markov Processes
    We present a theoretical analysis of Maxi- mum a Posteriori (MAP) sequence estima- tion for binary symmetric hidden Markov processes. We reduce the MAP ...
  6. [6]
    Easy and reliable maximum a posteriori Bayesian estimation of ...
    The mapbayr package was developed to perform maximum a posteriori Bayesian estimation (MAP‐BE) in R from any population PK model coded in mrgsolve.
  7. [7]
    Maximum a posteriori estimation for linear structural dynamics ...
    We apply MAP estimation in the context of structural dynamic models, where the system response can be described by the frequency response function.Missing: sources | Show results with:sources
  8. [8]
    A Gentle Introduction to Bayesian Analysis - PubMed Central - NIH
    A major difference between frequentist and Bayesian methods is that only the latter can incorporate background knowledge (or lack thereof) into the analyses by ...
  9. [9]
    LII. An essay towards solving a problem in the doctrine of chances ...
    An essay towards solving a problem in the doctrine of chances. By the late Rev. Mr. Bayes, FRS communicated by Mr. Price, in a letter to John Canton, AMFR S.
  10. [10]
    Memoir on the Probability of the Causes of Events - Project Euclid
    Memoir on the Probability of the Causes of Events. Pierre Simon Laplace. Download PDF + Save to My Library. Statist. Sci. 1(3): 364-378.
  11. [11]
    [PDF] Likelihood and Bayesian inference and computation
    The simplest form of Bayesian inference uses a uniform prior distribution, so that the posterior distribution is the same as the likelihood function (when ...
  12. [12]
    [PDF] Conjugate priors: Beta and normal Class 15, 18.05
    With a conjugate prior the posterior is of the same type, e.g. for binomial likelihood the beta prior becomes a beta posterior.
  13. [13]
    [PDF] Conjugate Priors, Uninformative Priors - UBC Computer Science
    If we don't have strong beliefs about what θ should be, it is common to use an uninformative or non-informative prior, and to let the data speak for itself.
  14. [14]
    Bayes, Jeffreys, Prior Distributions and the Philosophy of Statistics
    In this brief discussion I will argue the following: (1) in thinking about prior distributions, we should go beyond Jeffreys's principles and move toward weakly.
  15. [15]
    [PDF] 1 Bayesian approach 2 Regularization with priors: MAP inference
    Mar 19, 2024 · Today we discussed a Bayesian perspective on regularization. To summarize briefly, we can interpret any cost function as a log-prior.
  16. [16]
    Bayesian Interpretation of Regularization - SpringerLink
    May 14, 2022 · The use of priors is most useful whenever the data alone are not sufficient to provide reliable parameter estimates but there exists some a ...
  17. [17]
    Chapter 3 The Beta-Binomial Bayesian Model - Bayes Rules!
    Via Bayes' Rule, the conjugate Beta prior combined with the Binomial data model produce a Beta posterior model for π π . The updated Beta posterior ...
  18. [18]
    [PDF] Pattern Recognition and Machine Learning - Microsoft
    This new textbook reflects these recent developments while providing a compre- hensive introduction to the fields of pattern recognition and machine learning.
  19. [19]
    [PDF] Bayesian Data Analysis Third edition (with errors fixed as of 20 ...
    This book is intended to have three roles and to serve three associated audiences: an introductory text on Bayesian inference starting from first principles, a ...
  20. [20]
    [PDF] Challenges in Computing and Optimizing Upper Bounds of Marginal ...
    Estimating the marginal likelihood in probabilistic models is the holy grail of Bayesian inference. Marginal likelihoods allow us to compute the posterior ...
  21. [21]
    [PDF] Conjugate Bayesian analysis of the Gaussian distribution
    Oct 3, 2007 · The use of conjugate priors allows all the results to be derived in closed form. Unfortunately, different books use different conventions on how ...Missing: maximum posteriori
  22. [22]
    [PDF] A Geometric View of Conjugate Priors - IJCAI
    Another problem with considering a non-conjugate prior is that the ... This tells us why non-conjugate does not give us a closed form solution for ˆθMAP .Missing: fails | Show results with:fails<|control11|><|separator|>
  23. [23]
    [PDF] a view of the em algorithm that justifies incremental, sparse, and ...
    Abstract. The EM algorithm performs maximum likelihood estimation for data in which some variables are unobserved. We present a function that.
  24. [24]
    [1905.04365] Hyperparameter Estimation in Bayesian MAP Estimation
    May 10, 2019 · In this paper we study the effect of the choice of parameterization on MAP estimators when a conditionally Gaussian hierarchical prior ...Missing: tuning seminal
  25. [25]
    [PDF] A Derivation of the Soft-Thresholding Function
    The soft-thresholding function is a non-linear function used for signal denoising, applied to the transform-domain representation of a signal.
  26. [26]
    [PDF] MLE vs. MAP
    Jan 30, 2023 · ○ Maximum a posteriori (MAP) estimation. Choose value that is ... Gradient Ascent for M(C)LE. 15. Gradient ascent rule for w0: = Xj "yj.
  27. [27]
    High dimensional Bernstein-von Mises: simple examples - PMC - NIH
    The Bernstein-von Mises theorem says, informally, that this posterior distribution is, in large samples, approximately normal with mean approximately the MLE, θ ...