Fact-checked by Grok 2 weeks ago

Density estimation

Density estimation is the statistical process of constructing an estimate of the of a from a of observed points, aiming to reconstruct the underlying that generated the . This approach is fundamental in , where minimal assumptions are made about the form of the density, contrasting with parametric methods that presuppose a specific family of s, such as distribution. The primary goal is to produce a smooth and accurate approximation of the true density p(x), often denoted as \hat{p}(x), which depends on a smoothing parameter like h to balance bias and variance in the estimate. Historically, early ideas of smoothing data trace back to histograms, introduced by in 1895, with modern nonparametric techniques emerging in the mid-20th century. Pioneering work includes the kernel-based methods proposed by Fix and Hodges in 1951 for nearest-neighbor estimation, followed by formalizations from Rosenblatt in 1956 and Parzen in 1962, which established as a cornerstone of the field. These developments built on earlier smoothing concepts, such as those by Einstein in 1914 for periodograms, evolving into a robust framework for by the 1980s through influential texts like Silverman's Density Estimation for Statistics and Data Analysis. Key methods in density estimation encompass both parametric and nonparametric techniques, with the latter being more flexible for complex or unknown distributions. Parametric approaches assume a predefined form and estimate parameters via maximum likelihood, while nonparametric methods include histograms, which partition data into bins for crude approximation; kernel density estimation (KDE), formulated as \hat{p}(x) = \frac{1}{nh} \sum_{i=1}^n K\left( \frac{x - X_i}{h} \right) using kernels like Gaussian or Epanechnikov; and other variants such as orthogonal series expansions or penalized likelihood. Bandwidth selection remains critical, often via cross-validation, to achieve optimal convergence rates, such as n^{-4/5} for KDE under smoothness assumptions. Density estimation plays a vital role in , enabling visualization of distribution shapes, detection of multimodality, and assessment of features like skewness or tails. It underpins applications in for tasks like and clustering, in for modeling income distributions, and in scientific fields such as astronomy or for inferring population densities from samples. Advances continue to integrate it with Bayesian frameworks and high-dimensional data challenges, enhancing its utility in modern .

Overview

Definition

Density estimation is the construction of an estimate of the probability density function (PDF) of a random variable from a finite set of observed data points drawn from the underlying distribution. Specifically, given independent and identically distributed (i.i.d.) samples \{X_1, \dots, X_n\} from an unknown PDF f, the task is to construct an estimator \hat{f}(x) that approximates f(x) as closely as possible. A f(x) for a continuous X is a non-negative satisfying f(x) \geq 0 for all x in the and integrating to : \int f(x) \, dx = 1. This ensures that the probability of X falling in any interval (a, b) is given by P(a < X < b) = \int_a^b f(x) \, dx, providing a complete description of the distribution's shape and probabilities. The estimator \hat{f}(x) aims to mimic these properties, often evaluated by criteria such as the expected integrated squared error, E \int (\hat{f}(x) - f(x))^2 \, dx, which balances closeness to the true density with smoothness. Density estimation can be univariate, where the random variable X is one-dimensional (d=1), or multivariate, involving a d-dimensional vector \mathbf{X} with joint PDF f(\mathbf{x}). In higher dimensions, the problem becomes significantly more challenging due to the curse of dimensionality: as d increases, the volume of the space grows exponentially, leading to data sparsity and requiring an impractically large sample size n (scaling as O(2^d) or worse) for reliable estimation, which slows convergence rates dramatically. This phenomenon arises because points in high-dimensional space concentrate near boundaries and become increasingly isolated, complicating the approximation of f. It finds applications in statistics and machine learning for modeling underlying data distributions without strong parametric assumptions.

Purpose and applications

Density estimation serves several primary purposes in statistical analysis and data science. It enables the visualization of data distributions to reveal underlying patterns such as multimodality or skewness, facilitating intuitive understanding of complex datasets. Additionally, it supports anomaly detection by identifying observations that deviate significantly from the estimated density, which is crucial for quality control and fraud detection systems. The technique also aids in smoothing noisy data, providing a cleaner representation of the true underlying distribution without assuming a specific parametric form. Furthermore, density estimates act as foundational components for advanced statistical tasks, including classification via , hypothesis testing for distribution goodness-of-fit, and generative modeling in machine learning. The historical development of density estimation traces back to the late 19th century, when Karl Pearson introduced the histogram in 1895 as a graphical method for representing frequency distributions of continuous variables. This early approach allowed for basic empirical analysis of data distributions in fields like biometrics and astronomy. With computational advancements in the 20th century, particularly from the 1950s onward, more sophisticated nonparametric techniques emerged, enabling broader application beyond simple visualization to inferential statistics and predictive modeling. Seminal contributions, such as those by Rosenblatt (1956) and Parzen (1962), marked the evolution toward kernel-based methods, which addressed limitations of histograms in handling continuous data smoothly. In practice, density estimation finds key applications across diverse domains. In statistics, it underpins empirical distribution analysis and quantile estimation, as seen in demographic studies like mortality rate modeling using data from the . In machine learning, it powers generative models for data synthesis and outlier detection in unsupervised settings, enhancing tasks like clustering and dimensionality reduction. Financial applications include risk modeling, where density forecasts of asset returns inform value-at-risk calculations and portfolio optimization, as demonstrated in analyses of S&P 500 returns. In bioinformatics, it facilitates gene expression analysis by estimating distributions of expression levels to identify differentially expressed genes, and models sequence spaces for protein design and evolutionary studies. Kernel density estimation is a commonly used tool for these applications due to its flexibility in capturing complex shapes. A major advantage of density estimation, particularly in nonparametric forms, is its flexibility when the true probability density function is unknown or does not conform to standard parametric assumptions like normality. This model-free approach avoids misspecification biases that plague parametric methods, allowing reliable inference from data alone in exploratory or high-dimensional settings. Such versatility has made it indispensable in modern data-driven fields where distributional forms are often irregular or evolving.

Parametric methods

Maximum likelihood estimation

Maximum likelihood estimation (MLE) is a fundamental parametric method for density estimation, where the goal is to select parameters \theta from a specified family of probability density functions f(x \mid \theta) that best explain an observed sample X_1, X_2, \dots, X_n. The approach maximizes the likelihood function L(\theta) = \prod_{i=1}^n f(X_i \mid \theta), which represents the probability of observing the data under the model parameterized by \theta. To facilitate computation, the log-likelihood \ell(\theta) = \sum_{i=1}^n \log f(X_i \mid \theta) is typically maximized instead, as it is a monotonic transformation of L(\theta). The estimation proceeds by finding the value \hat{\theta} that solves the score equation \frac{\partial \ell(\theta)}{\partial \theta} = 0, assuming the conditions for differentiability hold. This first-order condition sets the derivative of the log-likelihood to zero, yielding the maximum likelihood estimator \hat{\theta}. For complex models, numerical optimization techniques such as may be employed if a closed-form solution is unavailable. A classic example is the univariate (normal) distribution with density f(x \mid \mu, \sigma^2) = \frac{1}{\sqrt{2\pi \sigma^2}} \exp\left( -\frac{(x - \mu)^2}{2\sigma^2} \right), where \theta = (\mu, \sigma^2). The log-likelihood is \ell(\mu, \sigma^2) = -\frac{n}{2} \log(2\pi \sigma^2) - \frac{1}{2\sigma^2} \sum_{i=1}^n (X_i - \mu)^2. Differentiating with respect to \mu gives \frac{\partial \ell}{\partial \mu} = \frac{1}{\sigma^2} \sum_{i=1}^n (X_i - \mu) = 0, so \hat{\mu} = \bar{X} = \frac{1}{n} \sum_{i=1}^n X_i. Substituting this into the derivative with respect to \sigma^2, \frac{\partial \ell}{\partial \sigma^2} = -\frac{n}{2\sigma^2} + \frac{1}{2(\sigma^2)^2} \sum_{i=1}^n (X_i - \mu)^2 = 0, yields \hat{\sigma}^2 = \frac{1}{n} \sum_{i=1}^n (X_i - \bar{X})^2. These estimators are the and the biased , respectively. Under standard regularity conditions—such as the density being twice differentiable, the support independent of \theta, and the Fisher information being positive definite—the MLE \hat{\theta} is asymptotically consistent, meaning \hat{\theta} \xrightarrow{p} \theta_0 as n \to \infty, where \theta_0 is the true parameter. Moreover, it is asymptotically efficient, achieving the Cramér-Rao lower bound on the variance, with \sqrt{n} (\hat{\theta} - \theta_0) \xrightarrow{d} \mathcal{N}(0, I(\theta_0)^{-1}), where I(\theta_0) is the Fisher information matrix. This efficiency follows from Cramér's theorem, which establishes that the MLE attains the minimal asymptotic variance among unbiased estimators. These properties make MLE particularly appealing for large samples when the parametric form is correctly specified. The method was pioneered by Ronald A. Fisher in the early 1920s, with its formal introduction in his 1922 paper on the foundations of theoretical statistics. However, MLE can perform poorly if the assumed parametric family is misspecified; for instance, fitting a unimodal normal distribution to multimodal data leads \hat{\theta} to converge to a pseudo-true value that minimizes the Kullback-Leibler divergence to the true density, resulting in a biased and potentially misleading estimate of the underlying distribution. In such cases, the estimated density may fail to capture essential features like multiple modes, highlighting the importance of model validation. Compared to non-parametric methods, MLE offers computational efficiency but at the cost of reduced flexibility when the true density deviates from the assumed form.

Bayesian estimation

Bayesian estimation provides a probabilistic framework for parametric density estimation by treating the parameters \theta of the density f(x \mid \theta) as random variables. A prior distribution \pi(\theta) encodes beliefs about \theta before observing data, and the likelihood L(\theta) = \prod_{i=1}^n f(x_i \mid \theta) quantifies how well the model fits the observed sample \{x_1, \dots, x_n\}. The posterior distribution is then given by Bayes' theorem: \pi(\theta \mid \data) \propto L(\theta) \pi(\theta), which updates the prior with the data. The resulting density estimate is the posterior predictive distribution, f(x \mid \data) = \int f(x \mid \theta) \pi(\theta \mid \data) \, d\theta, representing the marginal predictive density for new observations. Computing the posterior and predictive distributions analytically is feasible when conjugate priors are used, where the prior and likelihood belong to the same family, yielding a posterior in that family. For instance, the normal-inverse-gamma distribution serves as a conjugate prior for the mean and precision (inverse variance) of a Gaussian likelihood, allowing closed-form updates to the hyperparameters based on the data. In more complex models without conjugacy, numerical methods are essential; Markov Chain Monte Carlo (MCMC) algorithms, such as Gibbs sampling or Metropolis-Hastings, generate samples from the posterior to approximate integrals like the predictive density via Monte Carlo estimation. A specific example arises in estimating a univariate normal density f(x \mid \mu, \sigma^2) = \mathcal{N}(x \mid \mu, \sigma^2) with both parameters unknown, using non-informative priors to minimize influence from unsubstantiated assumptions. The Jeffreys prior, proportional to $1/\sigma, is often applied, leading to a posterior that is normal-inverse-gamma; the resulting posterior predictive density is then a Student's t-distribution centered at the sample mean with scale incorporating the sample variance and degrees of freedom equal to the sample size minus one. This approach offers key advantages over point-estimate methods, as the posterior fully quantifies uncertainty in \theta, enabling credible intervals for the density, and it mitigates overfitting in small samples by incorporating prior regularization, often yielding more stable estimates than maximum likelihood. With flat priors, Bayesian estimation specializes to maximum likelihood as a limiting case. The foundations trace to Thomas Bayes' 1763 essay on inverse probability, extended to modern density estimation in the 20th century through conjugate analysis and computational advances like MCMC.

Non-parametric methods

Histogram estimation

Histogram estimation is a fundamental non-parametric technique for approximating the probability density function of a continuous random variable from a sample of data points. It involves partitioning the range of the data into a series of contiguous intervals, known as bins, and constructing a bar graph where the height of each bar represents the frequency of observations falling within that bin, scaled to form a density estimate. This method provides a simple visualization of the data's distribution and serves as a basis for more sophisticated density estimation approaches. The procedure for histogram-based density estimation begins by selecting the bin width h and dividing the data range into k equal-width bins, where k is determined by the extent of the data and h. For a univariate sample X_1, \dots, X_n from a density f, let n_i denote the number of observations in the i-th bin, say [a_{i-1}, a_i), with a_i = a_{i-1} + h. The histogram estimator \hat{f}(x) is then defined as \hat{f}(x) = \frac{n_i}{n h} for x in the i-th bin, and zero otherwise; this ensures the estimate integrates to 1 over the real line, mimicking a probability density. The term "histogram" was coined by in his 1895 paper on skew variation, where he applied it to univariate data such as measurements of crab foreheads to graphically represent frequency distributions and assess skewness, often resulting in jagged estimates due to the discrete binning. A critical aspect of histogram estimation is the choice of bin width h or equivalently the number of bins k, as these parameters directly influence the smoothness and accuracy of the estimate. Common rules include for the number of bins, k \approx 1 + \log_2 n, which aims to balance detail and readability for moderate sample sizes but can lead to oversmoothing for large n. Another widely used guideline is for bin width, h \approx 3.5 \sigma n^{-1/3}, where \sigma is the sample standard deviation, optimizing the integrated mean squared error under normality assumptions. These selections help mitigate arbitrary choices, though the optimal h depends on the underlying density's smoothness. Despite its intuitiveness, the histogram estimator exhibits key statistical properties that highlight its limitations. It is discontinuous at bin boundaries, leading to artifacts like jagged edges that can misrepresent smooth underlying densities, and it is highly sensitive to the choice of binning scheme, with shifts in origin or width altering the estimate substantially. Asymptotically, under suitable conditions, the bias of \hat{f}(x) is of order O(h^2), reflecting smoothing over the bin width, while the variance is O(1/(n h)), capturing the variability from finite samples in each bin; this bias-variance tradeoff necessitates careful parameter tuning, often improved by smoothing methods like .

Kernel density estimation

Kernel density estimation (KDE) is a non-parametric technique that produces a smooth approximation of an unknown probability density function from a finite set of observed data points. It achieves this by placing a kernel function centered at each data point and averaging these kernels, weighted by a parameter that controls the degree of smoothing. This approach, pioneered by Rosenblatt in 1956 and further developed by Parzen in 1962, offers flexibility in capturing the underlying density shape without assuming a specific parametric form. The estimator is given by \hat{f}(x) = \frac{1}{n h} \sum_{i=1}^n K\left( \frac{x - X_i}{h} \right), where n is the sample size, X_i are the observed data points, h > 0 is the bandwidth, and K is the kernel function. A common choice for K is the Gaussian kernel, K(u) = \frac{1}{\sqrt{2\pi}} \exp\left( -\frac{u^2}{2} \right). This formulation ensures that \hat{f}(x) integrates to 1 and provides a continuous estimate, contrasting with discrete approximations like histograms. The kernel K must satisfy certain properties to yield a valid density estimate: it integrates to 1 over the real line, is typically symmetric around zero for unbiased estimation at the , and has finite second moment \mu_2(K) = \int u^2 K(u) \, du < \infty to control in . Common kernels include the uniform kernel K(u) = \frac{1}{2} I(|u| \leq 1), which produces a piecewise constant estimate similar to a in the limit; the Epanechnikov kernel K(u) = \frac{3}{4} (1 - u^2) I(|u| \leq 1), known for optimality in minimizing asymptotic integrated squared error; and the Gaussian kernel, valued for its infinite and despite slightly higher variance. Among these, the Epanechnikov kernel is asymptotically most efficient, though practical differences in performance are often minor. Bandwidth selection is critical, as it balances bias and variance in the estimate. The asymptotic mean squared error (MSE) at a point x is approximately \text{MSE}(\hat{f}(x)) \approx \frac{R(K)}{n h} f(x) + \frac{h^4}{4} [\mu_2(K) f''(x)]^2, where R(K) = \int K(u)^2 \, du is the roughness of the kernel and f''(x) is the second derivative of the true density; this yields bias of order O(h^2) and variance of order O(1/(n h)). Methods for choosing h include least squares cross-validation (LSCV), which minimizes an estimate of the integrated squared error \int (\hat{f}(x) - f(x))^2 \, dx via \text{ISCV}(h) = \int \hat{f}_h(x)^2 \, dx - \frac{2}{n} \sum_{i=1}^n \hat{f}_{h,-i}(X_i), where \hat{f}_{h,-i} is the estimator omitting the i-th observation; and plug-in rules, which estimate the optimal h_{\text{MISE}} \approx \left( \frac{R(K)}{n \mu_2^2(K) R(f'')} \right)^{1/5} by iteratively estimating higher-order derivatives of f using a pilot bandwidth. LSCV is unbiased but can be unstable for small samples, while plug-in methods provide consistent estimates with lower variability. In the multivariate case, the estimator generalizes to \hat{f}(\mathbf{x}) = \frac{1}{n |\mathbf{H}|^{1/2}} \sum_{i=1}^n K_d \left( \mathbf{H}^{-1/2} (\mathbf{x} - \mathbf{X}_i) \right), where \mathbf{x} \in \mathbb{R}^d, \mathbf{H} is a positive definite bandwidth matrix controlling smoothing in each direction, and K_d is a d-dimensional kernel, often a product or radial form of univariate kernels. The optimal bandwidth scales with n^{-1/(d+4)}, leading to slower convergence rates as d increases—a phenomenon known as the curse of dimensionality, which exacerbates sparseness and requires exponentially more data for reliable estimation in high dimensions.

Other non-parametric techniques

Orthogonal series estimators represent the density function as an expansion in a complete orthogonal basis, such as Fourier, Legendre polynomials, or wavelets, given by \hat{f}(x) = \sum_{k=0}^m \hat{c}_k \phi_k(x), where \hat{c}_k are empirical coefficients computed from the data and the series is truncated at order m \approx n^{1/5} to balance bias and variance. This approach, first proposed by Čencov in 1962, excels in high-dimensional settings by leveraging sparse approximations when the true density aligns well with the basis functions, allowing efficient computation and adaptation to structured densities. Nearest neighbor estimators construct the density at a point x by considering the local data configuration, defined as \hat{f}(x) = \frac{k}{n V_d(r(x))}, where k is the number of neighbors within a data-adaptive radius r(x) such that the d-dimensional volume V_d(r(x)) encloses exactly k points, with k typically growing as k(n) \to \infty but k(n)/n \to 0. Introduced by Loftsgaarden and Quesenberry in 1965, this method provides inherent adaptivity to local variations in density without fixed bandwidths, making it suitable for unevenly distributed data, though it can suffer from boundary biases in finite samples. Wavelet-based orthogonal series extend this framework to capture localized features effectively, using multiresolution bases that allow thresholding of coefficients to denoise the estimate while preserving discontinuities or sharp changes in the density. Donoho and colleagues in the 1990s demonstrated that wavelet thresholding achieves near-optimal rates for densities in Besov spaces, outperforming global bases like for non-smooth functions by adaptively selecting resolution levels. Partition estimation techniques divide the support into adaptive regions and estimate constant densities within each, often applied to mixture models without presupposing the number of components by using data-driven splits like binary trees or sequential partitioning to maximize likelihood. These methods, as explored in sieve maximum likelihood frameworks, offer flexibility for multimodal densities by refining partitions based on empirical evidence, avoiding parametric assumptions on component shapes. Compared to , orthogonal series methods provide sparse, global representations ideal for smooth or structured densities, while nearest neighbor approaches emphasize local adaptivity for irregular or clustered data patterns.

Evaluation and properties

Bias-variance tradeoff

In density estimation, the performance of an \hat{f} is often evaluated using the mean integrated squared error (MISE), which decomposes into components reflecting systematic and random errors. Specifically, the MISE is defined as \text{MISE}(\hat{f}) = E\left[\int (\hat{f}(x) - f(x))^2 \, dx \right] = \int \text{Bias}^2(\hat{f}(x)) \, dx + \int \text{Var}(\hat{f}(x)) \, dx, where the bias term \text{Bias}(\hat{f}(x)) = E[\hat{f}(x)] - f(x) captures the expected deviation due to model assumptions or smoothing, and the variance term measures the estimator's sensitivity to fluctuations in the sample. This decomposition highlights the inherent tradeoff: reducing bias typically increases variance, and vice versa, necessitating careful balancing to minimize overall error. A key aspect of this tradeoff arises in smoothing parameter selection, such as the bandwidth h in kernel-based methods. Under-smoothing, achieved with a small h, minimizes by closely following local features but amplifies variance through heightened sensitivity to sampling noise. Conversely, over-smoothing with a large h reduces variance by averaging more points but introduces substantial by oversimplifying the underlying structure. The optimal h minimizes the MISE and, for twice-differentiable densities in one , scales as O(n^{-1/5}), where n is the sample size; this rate ensures the integrated squared and integrated variance contribute equally to the . In higher dimensions d, the curse of dimensionality exacerbates the tradeoff, with the optimal MISE converging at a slower rate of O(n^{-4/(d+4)}), making accurate estimation increasingly challenging as d grows due to sparse coverage. For specific methods like histograms and (), the tradeoff manifests in explicit error rates. In histograms, bin width choices yield proportional to the width squared and variance inversely proportional to n times the width, leading to an optimal bin width of O(n^{-1/3}) in one . For , the asymptotic MISE (AMISE) provides a refined , balancing kernel-specific (often O(h^2)) and variance (O(1/(n h^d))), with the one-dimensional optimal yielding an AMISE of O(n^{-4/5}). These rates underscore how sample size n drives convergence while dimensionality d intensifies variance, limiting practical utility in high dimensions without . Diagnostic tools like the bootstrap have been instrumental in quantifying this since the , when nonparametric density estimation gained prominence through early techniques. The bootstrap estimates MISE by resampling the data to simulate the distribution of \hat{f}, allowing empirical approximation of bias and variance without assuming the true density f. Pioneering applications in the 1980s and 1990s demonstrated its effectiveness for bandwidth selection, often outperforming analytical approximations in finite samples by directly minimizing bootstrap-estimated MISE. This approach remains a cornerstone for method selection, enabling robust performance assessment across varying data characteristics.

Selection of parameters

In density estimation, selecting appropriate parameters such as bandwidth in (KDE) or bin width in histograms is crucial for balancing smoothness and fidelity to the . Data-driven methods like cross-validation provide objective ways to optimize these parameters by minimizing estimates of prediction error. One prominent approach is unbiased cross-validation (UCV), which selects the bandwidth h by minimizing an unbiased estimate of the integrated squared error, given by \text{UCV}(h) = \iint \hat{f}_h(x)^2 \, dx - 2 \frac{1}{n} \sum_{i=1}^n \hat{f}_{h,-i}(X_i), where \hat{f}_{h,-i} is the density estimate omitting the i-th and the double integral term is estimated unbiasedly to avoid from the data points themselves. Another data-driven technique is likelihood cross-validation (LCV), which chooses h to maximize the leave-one-out log-likelihood \sum_{i=1}^n \log \hat{f}_{-i}(X_i), offering robustness in scenarios where squared error may overemphasize outliers. Rule-of-thumb selectors offer quick, heuristic approximations when computational resources are limited. Silverman's rule, for instance, recommends h = 0.9 \hat{\sigma} n^{-1/5} for Gaussian kernels, where \hat{\sigma} is the sample standard deviation and n is the sample size; this formula assumes near-normal data and provides a reasonable starting point for many univariate cases. Extensions to adaptive bandwidths address non-uniform data densities by allowing h(x) to vary locally, often scaling a global bandwidth by a factor like |\hat{f}(x)|^{-1/5} based on a pilot density estimate, improving performance in multimodal distributions. For other density estimators, parameter selection follows analogous principles. In histogram estimation, the Freedman-Diaconis rule sets the bin width as h = 2 \frac{\mathrm{IQR}}{n^{1/3}}, where IQR is the interquartile range, yielding a robust choice insensitive to outliers compared to standard deviation-based rules. For series-based estimators, such as those using orthogonal polynomials or wavelets, model complexity is selected via information criteria like Akaike's Information Criterion (AIC), which penalizes overfitting by $2k where k is the number of parameters, or the Bayesian Information Criterion (BIC), applying a stronger k \log n penalty for large samples. Parameter selection becomes challenging in high dimensions due to of dimensionality, where computational costs for cross-validation scale exponentially with dimension, often making exhaustive searches infeasible without approximations like . Practical implementations mitigate this through libraries such as R's stats::density() function, which defaults to Silverman's rule but supports cross-validation via bw="nrd0" or custom methods, and Python's scipy.stats.gaussian_kde, offering rule-of-thumb bandwidths with options for cross-validation extensions. Poor choices can exacerbate the bias-variance , leading to either oversmoothed estimates missing structure or undersmoothed ones amplifying noise.

References

  1. [1]
    [PDF] Lecture 7: Density Estimation
    Density estimation is the problem of reconstructing the probability density function using a set of given data points. Namely, we observe X1, ··· ,Xn and we ...
  2. [2]
    [PDF] Density Estimation 36-708 1 Introduction - Statistics & Data Science
    The goal of nonparametric density estimation is to estimate p with as few assumptions about p as possible. We denote the estimator by bp. The estimator will ...
  3. [3]
    [PDF] density estimation including examples - UC Davis Statistics
    (Comprehensive yet concise overview over kernel density estimation, including bandwidth choice and extensions to kernel regression and other kernel.
  4. [4]
    [PDF] DENSITY ESTIMATION FOR STATISTICS AND DATA ANALYSIS
    Mar 15, 2002 · The two main aims of the book are to explain how to estimate a density from a given data set and to explore how density estimates can be used, ...
  5. [5]
    [PDF] Chapter 1 - Density Estimation - Princeton University
    The estimation of probability density functions (PDFs) and cumulative distribution functions (CDFs) are cornerstones of applied data analysis.
  6. [6]
    Multivariate Density Estimation: Theory, Practice, and Visualization
    Aug 17, 1992 · Kernel Density Estimators (Pages: 125-193) · Summary · PDF · Request permissions. CHAPTER 7. no. The Curse of Dimensionality and Dimension ...
  7. [7]
    Lecture 9: Density Estimation — Applied ML - Kuleshov Group
    The task of density estimation is to learn a probabilistic model on an unsupervised dataset to approximate the true data distribution . We can naturally ...
  8. [8]
    Density estimation using deep generative neural networks - PNAS
    Given the observed data, density estimation aims to recover the underlying density while deep generative modeling aims to generate new data similar to the ...
  9. [9]
    Machine Learning: Algorithms, Real-World Applications and ... - NIH
    The most common unsupervised learning tasks are clustering, density estimation, feature learning, dimensionality reduction, finding association rules, anomaly ...
  10. [10]
    (PDF) On the origin of Karl Pearson's term "histogram" - ResearchGate
    A histogram is another widely used and common display chart in scientific study and coined by famous statistician Karl Pearson [80]. It is constructed using ...
  11. [11]
    On the origin of Karl Pearson's term “histogram” | Semantic Scholar
    espanolMany modern scholars think that the term “histogram” is related to the word “history”. Recent work in the field of the history of statistics has only ...
  12. [12]
    [PDF] Evaluating Density Forecasts with Applications to Financial Risk
    We study density forecasts of daily value-weighted S&P 500 returns, with divi- dends, from 02/03/62 through 12/29/95.
  13. [13]
    Field-theoretic density estimation for biological sequence space with ...
    Density estimation in sequence space is a fundamental problem in machine learning that is also of great importance in computational biology.
  14. [14]
    Nebulosa recovers single-cell gene expression signals by kernel ...
    Jan 18, 2021 · Here, we introduce Nebulosa, an R package that uses weighted kernel density estimation to recover signals lost through drop-out or low expression.
  15. [15]
    Nonparametric Density Estimation for High-Dimensional Data - arXiv
    Mar 30, 2019 · Density Estimation is one of the central areas of statistics whose purpose is to estimate the probability density function underlying the ...
  16. [16]
  17. [17]
    On the mathematical foundations of theoretical statistics - Journals
    A recent paper entitled "The Fundamental Problem of Practical Statistics," in which one of the most eminent of modern statisticians presents what purports to ...
  18. [18]
    Normal distribution - Maximum likelihood estimation - StatLect
    Maximum likelihood estimation (MLE) of the parameters of the normal distribution. Derivation and properties, with detailed proofs.
  19. [19]
    Maximum Likelihood Estimation of Misspecified Models - jstor
    This paper examines the consequences and detection of model misspecification when using maximum likelihood techniques for estimation and inference.
  20. [20]
    Bayesian inference | Introduction with explained examples - StatLect
    Introduction to Bayesian statistics with explained examples. Learn about the prior, the likelihood, the posterior, the predictive distributions.
  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.
  22. [22]
    [PDF] A Catalog of Noninformative Priors * Contents - Statistics
    The Uniform Prior: By this, we just mean the constant density, with the constant typically being chosen to be 1 (unless the constant can be chosen to yield a ...
  23. [23]
    Remarks on Some Nonparametric Estimates of a Density Function
    This note discusses some aspects of the estimation of the density function of a univariate probability distribution.
  24. [24]
    On Estimation of a Probability Density Function and Mode
    September, 1962 On Estimation of a Probability Density Function and Mode. Emanuel Parzen · DOWNLOAD PDF + SAVE TO MY LIBRARY. Ann. Math. Statist.
  25. [25]
    [PDF] A Review of Kernel Density Estimation with Applications to ... - arXiv
    Dec 12, 2012 · The most used approach is kernel smoothing, which dates back to. Rosenblatt (1956) and Parzen (1962). The aim of this paper is to review the ...
  26. [26]
    Recent Developments in Nonparametric Density Estimation - jstor
    The early density estimation methods, such as the histogram, kernel estimators, and orthogonal series estimators are still very popular, and recent research ...Missing: seminal | Show results with:seminal
  27. [27]
    Cencov, N.N. (1962) Estimation of an Unknown Distribution Density ...
    Cencov, N.N. (1962) Estimation of an Unknown Distribution Density from Observations. Soviet Mathematics, 3, 1559-1562. has been cited by the following article:.Missing: exact orthogonal
  28. [28]
    Density estimation by wavelet thresholding - Project Euclid
    This paper explores density estimation using thresholding of wavelet coefficients, studying minimax rates of convergence and a single wavelet threshold  ...
  29. [29]
    [PDF] Multivariate Density Estimation via Adaptive Partitioning (I): Sieve MLE
    In this paper we study the asymptotic properties of density estimators based on adaptive partitioning. The data- adaptive partition is obtained by maximizing ...
  30. [30]
    Density Estimation - Project Euclid
    This paper provides a practical description of density estimation based on kernel methods. An important aim is to encourage practicing statisticians to apply ...<|control11|><|separator|>