Fact-checked by Grok 2 weeks ago

Statistical parameter

In statistics, a statistical parameter is a fixed numerical value that summarizes a characteristic of an entire , such as the (μ) or proportion (p), and is typically unknown because it requires data from every member of the . Unlike a , which is computed from a of the and serves as an estimate of the , a remains constant for a given and forms the basis for probabilistic models and . Statistical parameters play a central role in , where they represent true but often unobservable features of a that analysts aim to estimate or test using sample data. Common examples include the population mean (μ), which measures the ; the population variance (σ²) or standard deviation (σ), which quantify dispersion; and the (p), which indicates the fraction of the population with a specific trait. In , parameters also define the shape and properties of probability distributions, such as μ and σ for the normal distribution or λ for the , enabling the modeling of random phenomena. Estimating parameters involves methods like the method of moments or , which use sample statistics to approximate the true values and assess uncertainty through confidence intervals or hypothesis tests. These techniques are foundational in fields like , , and , where parameters inform decisions under uncertainty by bridging observed samples to broader populations. Parameters are distinguished from hyperparameters in advanced modeling, but in core statistics, they remain essential descriptors fixed by the population's underlying distribution.

Fundamentals

Definition and Scope

A statistical parameter is a numerical quantity that characterizes a specific feature of a population distribution, serving as a fixed value that summarizes an inherent property of the entire population. Examples include the population mean, denoted as \mu, which represents the average value across all members of the population, or the population variance, denoted as \sigma^2, which measures the spread of values around that mean. These parameters are typically unknown in practice, as they pertain to the complete population rather than observable data, forming the foundation for inferential statistics. The scope of statistical parameters extends to both finite and populations, where a finite population consists of a concrete, of units (such as all registered voters in a at a given time), allowing parameters to be theoretically computable if full data are available, though often impractical. In contrast, or conceptual populations treat the data-generating process as ongoing, making parameters abstract theoretical constructs that describe long-run behavior, such as the in repeated trials. This distinction highlights true parameters as fixed, population-level truths versus empirical approximations derived from observed data, which serve as estimates rather than the parameters themselves. The concept of statistical parameters was formally introduced by in the early as a core element of the parametric inference framework, emphasizing models where distributions are specified up to a of adjustable values. In his seminal 1922 paper, Fisher outlined the mathematical foundations of theoretical statistics, using parameters to bridge and , a development that revolutionized statistical methodology. This framework assumed that the form of the distribution is known, with parameters tuning its specifics, laying groundwork for modern estimation techniques. Understanding statistical parameters requires familiarity with foundational probability concepts, such as random variables—quantities that assume numerical values based on chance—and probability distributions, which assign probabilities to possible outcomes of those variables. These prerequisites enable the conceptualization of parameters as deterministic features embedded within probabilistic structures, distinct from the variability observed in samples.

Distinction from Sample Statistics

In statistics, a population parameter is a fixed numerical value that summarizes a characteristic of the entire , such as the true \mu of a , whereas a sample is a value computed from a of the , such as the sample \bar{x}, which serves as an of the . are inherently constant because they describe the complete without variability, while sample exhibit sampling variability due to the random selection process involved in drawing samples from the . This variability means that different samples from the same will yield different , reflecting the inherent in sampling, whereas the remains unchanged across all possible samples. The primary goal of inferential statistics is to use sample statistics as estimators to infer the unknown parameters, treating the parameters as fixed but typically inaccessible for . An estimator \hat{\theta} of a parameter \theta is generally expressed as a of the sample , \hat{\theta} = g(X_1, \dots, X_n), where X_1, \dots, X_n are and identically distributed random variables drawn from the . For unbiasedness, the of the estimator must equal the true parameter, \mathbb{E}[\hat{\theta}] = \theta, ensuring that, on average over repeated samples, the estimator does not systematically deviate from the parameter due to bias. Beyond unbiasedness, addresses the long-run behavior of estimators as the sample size n increases, requiring that \hat{\theta} converges in probability to \theta as n \to \infty, meaning the probability of the estimator being arbitrarily close to the true approaches 1. This property ensures that larger samples provide estimators that approximate the fixed population more reliably, without systematic error, thereby justifying parameters as the ultimate targets of .

Role in Probability and Distributions

Parameters in Parametric Distributions

distributions form a class of probability distributions where the entire form is defined by a finite-dimensional vector, distinguishing them from non- distributions that do not impose such a restrictive structure and may require an infinite number of parameters to fully describe the data-generating process. This finite parameterization enables concise modeling of complex phenomena under specific assumptions about the underlying . In these distributions, the parameters directly influence key characteristics such as , , and , fully determining the probability density or mass function. For instance, the normal distribution is parameterized by the \mu and variance \sigma^2, which control its and , respectively, as denoted by N(\mu, \sigma^2). Similarly, the , a foundational case, is governed by a single p \in [0,1], representing the probability of success in a binary trial. The general form of a parametric probability density function (for continuous cases) or mass function (for discrete cases) is expressed as f(x \mid \theta), where x is the random variable and \theta is the parameter vector that indexes the family of distributions. A core assumption of parametric distributions is that the parameter vector \theta fully specifies the distribution, meaning the family encompasses all necessary flexibility without redundancy or insufficiency. Under-parameterization occurs when the chosen family is too restrictive, failing to capture the true data distribution and often leading to biased inferences. Conversely, over-parameterization introduces redundant parameters, resulting in non-identifiability where multiple \theta values yield the same distribution, complicating unique recovery of parameters. This requirement ensures that parameters like shape descriptors remain meaningful within the model's structure.

Functional Forms and Identifiability

In statistical models, parameters often take functional forms that relate them directly to key characteristics of the , such as moments or quantiles, facilitating both theoretical analysis and estimation. For instance, the parameter \mu in a is expressed as the first moment, \mu = \mathbb{E}[X], while higher-order parameters like variance capture second moments. This moment-based parameterization underpins the method of moments estimation, where population parameters are solved as functions of theoretical moments equated to their sample counterparts. Similarly, quantile-parameterized distributions express parameters in terms of specific percentiles, such as the or , offering robustness to outliers in or modeling tasks. Reparameterization involves transforming the parameter space, such as substituting \mu with \log(\mu) for positive constraints, without altering the underlying probability model or its implications for . This change re-expresses the likelihood or posterior in terms of new parameters, potentially simplifying optimization or sampling in Bayesian or maximum likelihood frameworks, as the Jacobian adjustment ensures equivalence in the parameter space. For example, in multilevel models, reparameterizing variance components from standard deviations to correlations can enhance MCMC rates while preserving the model's probabilistic structure. Such transformations are particularly useful in hierarchical models, where they mitigate issues arising from overparameterization without affecting the validity of downstream inferences. A critical property of these functional forms is identifiability, which ensures that distinct parameter values \theta produce distinct distributions, enabling unique recovery from observed data. Formally, a parameter \theta is identifiable if the mapping from \theta to the induced probability measure P_\theta is injective, meaning \theta_1 \neq \theta_2 implies P_{\theta_1} \neq P_{\theta_2}. This condition guarantees that the likelihood function separates different parameter values, supporting consistent estimation. Non-identifiability arises in models like finite mixtures, where label switching among component parameters yields equivalent distributions, complicating inference unless constraints like ordering are imposed. In such cases, the mapping becomes non-injective, leading to multiple \theta values corresponding to the same P_\theta, as analyzed in early mixture model theory. The injectivity requirement can be expressed mathematically as the parameter-to-distribution map being : \theta \mapsto P_\theta \quad \text{is injective.} This structural must hold for the model to permit precise recovery, distinguishing identifiable formulations from those requiring auxiliary constraints.

Types and Classification

Location Parameters

A is a scalar in a that specifies its position along the real line, effectively translating the entire horizontally without affecting its shape or variability. This determines the central tendency of the , such as its or , and is fundamental in location families of where varying the shifts the (PDF). For instance, in a pure location , the PDF is given by f(x \mid \mu) = g(x - \mu), where \mu is the and g is the PDF of a fixed standard . In broader location-scale families, the location parameter \mu combines with a scale parameter \eta > 0 to form the PDF f(x \mid \mu, \eta) = \frac{1}{\eta} g\left( \frac{x - \mu}{\eta} \right), where the term (x - \mu)/\eta standardizes the variable relative to the location shift \mu. Properties of location parameters include their role as measures of central tendency; for symmetric distributions, \mu often coincides with both the mean and median. These parameters exhibit equivariance under affine transformations of the random variable—for example, if Y = aX + b with a > 0, the location of Y becomes a\mu + b, preserving the family's structure. Prominent theoretical examples illustrate this . In the normal N([\mu](/page/MU), \sigma^2), [\mu](/page/MU) serves as the , shifting the symmetric bell while \sigma^2 controls spread. Similarly, the Cauchy distribution's \mu translates its heavy-tailed, symmetric PDF, which lacks a defined but has \mu as its . The on interval (a, b) can be parameterized with location \mu = (a + b)/2 and (b - a)/2, where \mu centers the flat . In the , \mu acts as the , coinciding with the and , and shifting the S-shaped .

Scale and Shape Parameters

Scale parameters characterize the spread or dispersion of a probability distribution by multiplicatively stretching or compressing its scale, thereby affecting the variability of the random variable without altering its fundamental shape. In many parametric families, the scale parameter appears in the density function as a factor that normalizes the distribution after transformation, ensuring it integrates to unity. For instance, in the normal distribution, the standard deviation σ serves as the scale parameter, controlling the width of the bell-shaped curve. Similarly, in the , the parameter β (the mean, equivalent to the inverse of the rate λ) acts as a , where larger values of β expand the distribution's tail, representing longer expected waiting times. A canonical form for scale families, often in conjunction with a location parameter μ, is given by the probability density function: f(x \mid \mu, \sigma) = \frac{1}{\sigma} g\left( \frac{x - \mu}{\sigma} \right), where g is the base density (e.g., standard normal), σ > 0 is the scale parameter, and the transformation ensures scale invariance: if X has scale σ, then cX (c > 0) has scale cσ. This multiplicative property implies that scale parameters transform proportionally under linear scaling of the variable, preserving the relative spread. Shape parameters, in contrast, modify the underlying form of the distribution, influencing aspects such as , peakedness, or the heaviness of tails, which in turn affect higher-order moments like and . For example, in the defined on [0, 1], the parameters α > 0 and β > 0 are shape parameters that determine the distribution's : when α > β, the skews right; when β > α, it skews left; and occurs when α = β. In the kappa distribution, the shape parameter κ controls the tail behavior and boundedness, with negative κ yielding unbounded support and heavy tails suitable for modeling extreme events in or . An illustrative case is the , with density f(x \mid \alpha, \beta) = \frac{1}{\beta^\alpha \Gamma(\alpha)} x^{\alpha-1} e^{-x/\beta}, \quad x > 0, where α > 0 is the influencing (decreasing as α increases toward ) and β > 0 is the . parameters thus enable flexible modeling of non-standard forms within parametric families, distinct from mere rescaling.

Estimation Methods

Point Estimation Techniques

Point estimation techniques seek to produce a single numerical value, \hat{\theta}, as an approximation to an unknown population parameter \theta based on observed sample data from a random sample. This approach contrasts with methods that incorporate uncertainty through ranges, focusing instead on a direct, data-driven guess for \theta. The goal is to select \hat{\theta} such that it closely mirrors the true parameter in expectation or through optimization criteria, balancing factors like bias and variance in the estimation process. One classical method is the method of moments, pioneered by Karl Pearson in the late 19th century. It involves solving a system of equations where the first k sample moments are set equal to the first k theoretical population moments, with k matching the number of parameters to estimate. For instance, in estimating the mean \mu and variance \sigma^2 of a distribution, the first sample moment (the sample mean \bar{x}) equals \mu, and the second central sample moment equals \sigma^2, yielding the method of moments estimator for variance as \hat{\sigma}^2 = \frac{1}{n} \sum_{i=1}^n (x_i - \bar{x})^2, where n is the sample size. This technique is computationally simple and does not require assuming a specific distributional form beyond the moments, though it may yield less efficient estimators compared to other methods in finite samples. A more widely adopted technique is (MLE), formalized by A. Fisher in 1922. Given independent and identically distributed observations x_1, \dots, x_n from a probability density (or mass) function f(x_i \mid \theta), MLE defines the as L(\theta) = \prod_{i=1}^n f(x_i \mid \theta) and selects the estimator \hat{\theta} that maximizes L(\theta), often by maximizing the log-likelihood \ell(\theta) = \log L(\theta) for computational convenience. Equivalently, \hat{\theta} = \arg\max_{\theta} \ell(\theta). Under standard regularity conditions—such as the existence of the support not depending on \theta and differentiability of the log-likelihood—MLE estimators are consistent, meaning \hat{\theta} \to_p \theta as n \to \infty, and asymptotically normal, with \sqrt{n} (\hat{\theta} - \theta) \xrightarrow{d} \mathcal{N}\left(0, \mathcal{I}(\theta)^{-1}\right), where \mathcal{I}(\theta) is the . These properties establish MLE as efficient in large samples, though individual estimators may exhibit bias, prompting a between bias reduction and variance minimization in practical applications.

Interval Estimation and Confidence

Interval estimation extends point estimation by providing a range of plausible values for an unknown statistical parameter θ, rather than a single value, to account for sampling variability. A (CI) is constructed as an interval [L, U] such that the probability P(L ≤ θ ≤ U) = 1 - α holds asymptotically, where α is the significance level, meaning that in repeated sampling, the interval will contain the true parameter with frequency 1 - α as the sample size increases. This approach, formalized by , emphasizes the interval's behavior over hypothetical repetitions of the experiment rather than a direct probabilistic statement about θ given a fixed sample. Confidence intervals are typically constructed using the of a point estimator or a related . For instance, when estimating the μ of a with known standard deviation σ based on a sample of size n, the CI takes the form \bar{x} \pm z_{\alpha/2} \frac{\sigma}{\sqrt{n}}, where \bar{x} is the sample mean and z_{\alpha/2} is the (1 - α/2) of the standard ; this derives from the fact that \sqrt{n} (\bar{x} - μ)/σ follows a standard . More generally, a Q(X, θ) with a known independent of θ allows construction of a (1 - α) as the set of θ values for which α/2 ≤ F(Q(X, θ)) ≤ 1 - α/2, where F is the of Q. The of a is interpreted as the long-run proportion of that contain the true θ across repeated samples from the , not as the probability that a specific observed contains θ given the . This frequentist avoids assigning probability to fixed parameters and focuses on the procedure's reliability. For non- cases where the is unknown or complex, the bootstrap method resamples the with to approximate the of the , enabling or bias-corrected CIs; this technique, introduced by Bradley Efron, provides robust estimates without assuming a form for the underlying .

Applications and Examples

Univariate Case Studies

In the univariate case, the normal distribution provides a foundational example of statistical parameters, featuring a location parameter \mu, which determines the center of the distribution, and a scale parameter \sigma^2, which governs its spread or variance. These parameters fully characterize the distribution, with the probability density function given by f(x; \mu, \sigma^2) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(x - \mu)^2}{2\sigma^2}\right). Estimation of \mu is typically achieved using the sample mean \bar{x} = \frac{1}{n} \sum_{i=1}^n x_i, which serves as the maximum likelihood estimator (MLE), while \sigma^2 is estimated by the sample variance s^2 = \frac{1}{n-1} \sum_{i=1}^n (x_i - \bar{x})^2. For discrete univariate distributions, the illustrates a shape parameter p, representing the probability of success in a single trial, where the distribution takes values 0 or 1 with probabilities $1-p and p, respectively. This extends to the for n independent trials, where p remains the key parameter governing the proportion of successes. The MLE for p in a sample of n Bernoulli trials with k successes is the sample proportion \hat{p} = \frac{k}{n}, which unbiasedly estimates the true parameter and achieves the Cramér-Rao lower bound for efficiency. The offers another univariate example, parameterized by a \lambda > 0, often interpreted as the rate of events in a Poisson process, with the f(x; \lambda) = \lambda e^{-\lambda x} for x \geq 0. The of the is $1/\lambda, directly linking the parameter to the expected waiting time between events, while the variance equals $1/\lambda^2. This parameterization highlights \lambda's role in scaling the distribution's tail behavior. To illustrate estimation in practice, consider a hypothetical dataset from a normal distribution: observations x = \{2.1, 1.9, 2.3, 2.0, 2.2\} with assumed true \mu = 2. The sample mean is \bar{x} = \frac{2.1 + 1.9 + 2.3 + 2.0 + 2.2}{5} = 2.1, providing an estimate \hat{\mu} = 2.1 that is close to the true value, demonstrating the consistency of the estimator for moderate sample sizes. The sample variance is s^2 = \frac{1}{4} \sum (x_i - 2.1)^2 = 0.025, yielding \hat{\sigma} \approx 0.158, which captures the data's low variability.

Multivariate Extensions

In multivariate statistical models, parameters generalize from scalar values to vectors and matrices to capture joint behaviors across multiple dimensions. A prominent example is the , where the primary parameters are the mean vector \mu \in \mathbb{R}^k, which acts as the indicating of the distribution in k-dimensional , and the \Sigma \in \mathbb{R}^{k \times k}, a symmetric positive definite matrix that describes both the and interdependencies among variables. These parameters fully characterize the , enabling analysis of correlated data in fields such as , , and . The mean vector \mu extends the univariate location parameter, such as the , to specify the in each dimension, with marginal means aligning with its components. The \Sigma combines scale and shape aspects: its diagonal elements represent variances (scale in individual dimensions), off-diagonal elements capture covariances (linear dependencies), eigenvalues quantify scale along principal axes, and eigenvectors determine the orientation or shape of the elliptical contours of constant density. The |\Sigma| provides a measure of overall multivariate scale, reflecting the generalized volume of the , while the \operatorname{tr}(\Sigma) sums the variances for total . This allows \Sigma to model both isotropic (spherical) and anisotropic (elongated or rotated) spreads, distinguishing it from univariate scale parameters like standard deviation. Estimation of these parameters typically employs maximum likelihood methods for a sample \mathbf{x}_1, \dots, \mathbf{x}_n drawn from a k-variate . The maximum likelihood (MLE) for the vector is the sample vector \hat{\mu} = \bar{\mathbf{x}} = \frac{1}{n} \sum_{i=1}^n \mathbf{x}_i, which is unbiased and minimum-variance under . For the , the MLE is \hat{\Sigma} = \frac{1}{n} \sum_{i=1}^n (\mathbf{x}_i - \bar{\mathbf{x}})(\mathbf{x}_i - \bar{\mathbf{x}})^T, a biased but that converges to the true \Sigma as n increases; an unbiased alternative divides by n-1. These arise from maximizing the log-likelihood function derived from the multivariate normal . The probability density function for a single observation \mathbf{x} from the k-variate normal distribution is given by f(\mathbf{x} \mid \mu, \Sigma) = (2\pi)^{-k/2} |\Sigma|^{-1/2} \exp\left( -\frac{1}{2} (\mathbf{x} - \mu)^T \Sigma^{-1} (\mathbf{x} - \mu) \right), where |\Sigma| is the determinant and \Sigma^{-1} is the inverse covariance matrix (precision matrix). For n independent observations, the likelihood is the product of these densities, and taking the logarithm yields a function quadratic in the parameters, leading directly to the MLEs above upon differentiation and setting to zero. This framework underpins inference in multivariate settings, such as hypothesis testing for \mu or \Sigma, and extends to more complex models like factor analysis or Gaussian processes.

References

  1. [1]
    S.1 Basic Terminology | STAT ONLINE - Penn State
    A parameter is any summary number, like an average or percentage, that describes the entire population. The population mean μ (the greek letter ...
  2. [2]
    Population Parameters and Sample Statistics
    The parameter is the true but often unknown value that we would ideally like to know. Since populations are generally fixed, a parameter is generally also a ...
  3. [3]
    Parameters vs. Statistics - CUNY Pressbooks Network
    A parameter is a number that describes a population. A statistic is a number that we calculate from a sample. Let's use this new vocabulary to rephrase what we ...
  4. [4]
    [PDF] Purposes of Data Analysis Parameters and Statistics Variables and ...
    ❑ Parameters: Numbers that describe a population. For example, the population mean (µ)and standard deviation (σ). Statistics: Numbers that are calculated from ...
  5. [5]
    [PDF] Common Probability Distributions
    Definition. In statistics, a parameter θ = t(F) refers to a some function of a probability distribution that is used to characterize the distribution.
  6. [6]
    3 Probability Distributions – STAT 500 | Applied Statistics
    A normal curve has two parameters: mean μ (center of the curve); standard deviation ...
  7. [7]
    1.3.6.5. Estimating the Parameters of a Distribution
    There are various methods, both numerical and graphical, for estimating the parameters of a probability distribution. Method of moments · Maximum likelihood ...
  8. [8]
    Estimation of Parameters on Probability Density Function Using ...
    Parameter estimation is a field of statistics that involves estimating the parameters of a distribution utilizing data samples. For precise predicted results as ...<|control11|><|separator|>
  9. [9]
    [PDF] Parameter Estimation
    An estimator refers to the function g(·) that is applied to the sample to obtain the estimate ˆθ. The above definition uses standard notation in statistic, ...
  10. [10]
    Populations, Parameters, and Samples in Inferential Statistics
    In this blog post, learn the differences between population vs. sample, parameter vs. statistic, and how to obtain representative samples using random sampling.
  11. [11]
    [PDF] A primer on statistical inferences for finite populations
    Sep 2, 2020 · The traditional approach to statistical inference based on simple random sampling with replacement from infinite normal population ...
  12. [12]
    From evidence to understanding: a commentary on Fisher (1922 ...
    Ronald Fisher's seminal 1922 paper 'On the mathematical foundations of theoretical statistics' [1] was submitted to the Royal Society on 25 June 1921.
  13. [13]
    1.2 - Samples & Populations | STAT 200 - STAT ONLINE
    Values concerning a sample are referred to as sample statistics while values concerning a population are referred to as population parameters.
  14. [14]
    [PDF] Unbiased Estimation - Arizona Math
    In creating a parameter estimator, a fundamental question is whether or not the estimator differs from the parameter in a systematic manner.
  15. [15]
    1.3 - Unbiased Estimation | STAT 415 - STAT ONLINE
    In summary, we have shown that, if X i is a normally distributed random variable with mean μ and variance σ 2 , then S 2 is an unbiased estimator of σ 2 . It ...
  16. [16]
    [PDF] Properties of Estimators II 7.7.1 Consistency
    Definition 7.7.1: Consistency. An estimator ˆθn (depending on n iid samples) of θ is said to be consistent if it converges (in. probability) to θ. That is, for ...
  17. [17]
    [PDF] Unbiased Estimators, Std Error - Engineering Statistics Section 6.1
    Mar 25, 2016 · i.e. A point estimator is unbiased if its sampling distribution is always. ”centered” at the true value of the population parameter.
  18. [18]
    What are parametric models? | DataRobot Blog
    A parametric model is any model that captures all the information about its predictions within a finite set of parameters.Missing: fully specified
  19. [19]
    Parametric and Nonparametric Tests in Spine Research: Why Do ...
    Parametric and nonparametric are 2 broad classifications of statistical procedures. Parametric tests are based on assumptions about the distribution of the ...
  20. [20]
    Normal Distribution - Overview, Parameters, and Properties
    The two main parameters of a (normal) distribution are the mean and standard deviation. The parameters determine the shape and probabilities of the distribution ...
  21. [21]
    What Is a Bernoulli Distribution? A Deep Dive - DataCamp
    Aug 22, 2024 · It is characterized by a single parameter, p, which represents the probability of success. The probability of failure is consequently 1 - p.
  22. [22]
    A Gentle Introduction to Probability Density Estimation
    Jul 24, 2020 · Parametric probability density estimation involves selecting a common distribution and estimating the parameters for the density function from a ...Probability Density · Parametric Density... · Nonparametric Density...
  23. [23]
    Under-parameterized Model of Sequence Evolution Leads to Bias in ...
    Under-parameterized Model of Sequence Evolution Leads to Bias in the ... model (JC+Γ) and the underparameterized model (JC). For each phylogeny with ...Simulation And Analyses · More Complex Models · Results
  24. [24]
    On identifiability of parametric statistical models
    This is a review article on statistical identifiability. Besides the definition of the main concepts, we deal with several questions relevant to the statis.
  25. [25]
    Parameter Identifiability in Statistical Machine Learning: A Review
    May 1, 2017 · Parameter identifiability is concerned with the theoretical uniqueness of model parameter determined from the statistical family ⁠.
  26. [26]
    1.4 - Method of Moments | STAT 415 - STAT ONLINE
    The method of moments involves equating sample moments with theoretical moments. So, let's start by making sure we recall the definitions of theoretical ...
  27. [27]
    Quantile-Parameterized Distributions for Expert Knowledge Elicitation
    Mar 31, 2025 · This paper provides a comprehensive overview of quantile-parameterized distributions (QPDs) as a tool for capturing expert predictions and parametric judgments.
  28. [28]
    The use of simple reparameterizations to improve the efficiency of ...
    In this paper we have focused on the application of reparameterization methods to the estimation of multilevel discrete time survival models, but the three ...<|separator|>
  29. [29]
    Navigating the landscape of parameter identifiability methods
    The model parameters are formally identifiable, or in short, the model is formally identifiable, if two different parameter vectors lead to two different ...
  30. [30]
    [PDF] Lecture 9: Exponential and location-scale families
    Examples of location families are normal and Cauchy with location parameter µ ∈ 勿 and the other parameter σ fixed. Other examples are given later.
  31. [31]
    Univariate Distribution Relationships
    A shape parameter changes the shape of the probability density function. An example of a location parameter is the mean of a normal random variable; an example ...
  32. [32]
    1.3.6.4. Location and Scale Parameters
    Location parameters shift a distribution's graph horizontally, while scale parameters stretch or compress the graph. For normal distribution, location is mean ...Missing: properties | Show results with:properties
  33. [33]
    1.3.6.6.2. Uniform Distribution
    where A is the location parameter and (B - A) is the scale parameter. The case where A = 0 and B = 1 is called the standard uniform distribution. The ...
  34. [34]
    The Logistic Distribution - R
    Density, distribution function, quantile function and random generation for the logistic distribution with parameters location and scale.
  35. [35]
    [PDF] Common Families of Distributions - Purdue Department of Statistics
    Then the family of pdfs f(x−µ), indexed by the parameter µ, −∞ <µ< ∞, is called the location family with standard pdf f(x) and µ is called the location.
  36. [36]
    1.3.6.6.7. Exponential Distribution - Information Technology Laboratory
    The exponential distribution has a probability density function with parameters μ and β. It's used in reliability to model constant failure rates.
  37. [37]
    [PDF] Families of Distributions - Andrew B. Nobel
    Definition: The location family generated by a density f is given by. P = {f(x|θ) = f(x − θ) : θ ∈ R}. This is the set of densities of Y = X + θ where X ...
  38. [38]
    1.3.6.6.17. Beta Distribution - Information Technology Laboratory
    The following is the plot of the beta probability density function for four different values of the shape parameters. plot of the Beta probability density ...
  39. [39]
    Kappa Cumulative Distribution Function
    Jul 7, 2009 · Compute the kappa cumulative distribution function with shape parameters h and k. Description: The general form of the kappa distribution ...
  40. [40]
    Probability Playground: The Gamma Distribution
    The parameter α is known as the shape parameter, and the parameter β is called the scale parameter. Increasing α leads to a more "peaked" distribution, while ...
  41. [41]
    Lesson 1: Point Estimation | STAT 415
    Point estimation involves estimating unknown population parameters using random samples. This lesson covers maximum likelihood and method of moments methods.
  42. [42]
    [PDF] 6 Point Estimation
    Point estimation aims to estimate a parameter (θ) using sample data. A point estimate is a sensible guess for θ based on a sample.
  43. [43]
    [PDF] Lecture 3 Properties of MLE: consistency, asymptotic normality ...
    Asymptotic normality says that the estimator not only converges to the unknown parameter, but it converges fast enough, at a rate 1/≥n. Consistency of MLE. To ...
  44. [44]
    METHOD OF MOMENTS AND METHOD OF MAXIMUM LIKELIHOOD
    KARL PEARSON, F.R.S; METHOD OF MOMENTS AND METHOD OF MAXIMUM LIKELIHOOD, Biometrika, Volume 28, Issue 1-2, 1 June 1936, Pages 34–47, https://doi.org/10.109.Missing: original | Show results with:original<|separator|>
  45. [45]
    [PDF] 6 Classic Theory of Point Estimation - Purdue Department of Statistics
    Point estimation is a starting point for inference. Key concepts include the likelihood function, maximum likelihood estimates, and sufficient statistics.
  46. [46]
    [PDF] On the Mathematical Foundations of Theoretical Statistics
    Apr 18, 2021 · On the illathematical Foundations of Theoretical Statistics. 1By R. A. FISHER, M.A., Fellow of Gonville and Caims College, Cambridge, Chief.
  47. [47]
    Maximum likelihood estimation | Theory, assumptions, properties
    Learn the theory of maximum likelihood estimation. Discover the assumptions needed to prove properties such as consistency and asymptotic normality.The sample and its likelihood · Maximum likelihood estimator · Asymptotic properties
  48. [48]
    Outline of a Theory of Statistical Estimation Based on the Classical ...
    1937Outline of a Theory of Statistical Estimation Based on the Classical ... Mathematical notations produced through Infty OCR. DOWNLOAD PDF. Figures ...
  49. [49]
    The Normal Distribution - Utah State University
    A normal distribution has two parameters, the mean μ μ , and the variance σ2 σ 2 . The mean can be any real number and the variance can be any non-negative ...Missing: parameterization source
  50. [50]
    Week 4: The Normal Distribution - NTNU
    The estimate \(\hat{\mu}\) is just the sample mean, and \(\hat{\sigma}^2\) is the sample variance. These two statistics can be used to summarise the whole ...
  51. [51]
    9.4 Estimating the parameters of a Normal distribution
    This chapter will introduce two models for inferring the parameters of a (single) normal distribution, both of which are set-up in such a way that it is ...Missing: parameterization | Show results with:parameterization
  52. [52]
    1.2 - Maximum Likelihood Estimation | STAT 415
    The first example on this page involved a joint probability mass function that depends on only one parameter, namely p , the proportion of successes. Now ...
  53. [53]
    15.1 - Exponential Distributions | STAT 414 - STAT ONLINE
    If λ (the Greek letter "lambda") equals the mean number of events in an ... θ = 1 λ and λ = 1 θ. For example, suppose the mean number of customers to ...
  54. [54]
    Maximum-likelihood estimation of the parameters of a multivariate ...
    This paper provides an exposition of alternative approaches for obtaining maximum- likelihood estimators (MLE) for the parameters of a multivariate normal ...