Fact-checked by Grok 2 weeks ago

Frequentist probability

Frequentist probability, also known as the objective interpretation of probability, defines the probability of an as the limiting relative with which it occurs in an infinite sequence of independent trials under identical conditions. This approach treats probability as an empirical property of repeatable processes in , such as flips or rolls, rather than a measure of subjective or . It requires the existence of a stable, long-run for the event, emphasizing objectivity by grounding probabilities in rather than personal degrees of confidence. The historical roots of frequentist probability trace back to the 17th century with early work on games of chance by and , but it gained formal structure in the through applications in and error theory in astronomy, where stable frequency ratios emerged from large datasets. In the early 20th century, proposed a frequency-based view in 1909, laying groundwork for modern developments. Richard von Mises advanced the theory in the 1920s and 1930s by axiomatizing probability as the limit of relative frequencies in "random collectives"—infinite sequences satisfying conditions of randomness and independence—published in his seminal 1928 German work Wahrscheinlichkeit, Statistik und Wahrheit, later translated as Probability, Statistics and Truth. and further refined it in the 1930s through their decision-theoretic framework, introducing concepts like hypothesis testing and confidence intervals to apply frequencies to inductive inference in experimental design. Andrey Kolmogorov's 1933 axiomatic provided a mathematical foundation compatible with frequentism, while extended it to sequential analysis during . Central to frequentist probability are tools for statistical inference, including hypothesis testing, which evaluates evidence against a by controlling the long-run error rate of falsely rejecting a true hypothesis (Type I error, at significance level α, often 0.05). Confidence intervals complement this by constructing ranges around point estimates that cover the true parameter with a specified long-run success rate (e.g., 95% confidence), derived from the under repeated experiments. These methods prioritize in experiments to ensure validity and have been pivotal in fields like , , and , though they face critiques for not directly quantifying uncertainty in single events or incorporating prior information. In contrast to , which updates beliefs using priors, frequentism maintains an objective stance by avoiding subjective elements, influencing much of modern empirical science.

Fundamentals

Definition

Frequentist probability is an interpretation of probability within the framework of that defines the probability of an as the limiting relative with which that occurs in an infinite sequence of , identically distributed trials under fixed conditions. This approach grounds probability in empirical, rather than subjective degrees of , emphasizing long-run in repeated experiments. To formalize this, consider the foundational concepts of a , which is the set of all possible outcomes of a random experiment, and an , which is a subset of the representing outcomes of interest. These elements provide the structure for interpreting probabilities based on frequencies within the axiomatic framework of . Unlike other interpretations that may incorporate personal priors or subjective assessments, frequentist probability insists on an objective basis derived solely from the data-generating process and its hypothetical infinite repetitions. This empirical focus distinguishes it from general , where probabilities can sometimes reflect about unique events, by restricting application to scenarios amenable to replication. The thus captures the stable proportion of occurrences, ensuring that claims about likelihood are tied to verifiable patterns rather than opinion. A classic illustration is the toss, where the consists of two outcomes—heads or tails—and the event of interest is heads. In this setup, the probability of heads is 0.5, meaning that over an infinite number of flips, heads would appear in exactly half of the trials on average. This example highlights how frequentist probability operationalizes the concept through the expectation of equal relative frequencies for symmetric outcomes, providing a concrete, testable foundation for more complex statistical analyses.

Interpretation of Probability

In the frequentist framework, probability is interpreted as an objective property inherent to the experimental setup or process, rather than a subjective degree of held by an observer. This view posits that probabilities exist independently of personal judgments, grounded instead in the empirical regularity of outcomes from repeatable experiments. The of probability in this is given by the limiting relative of an 's occurrence over an sequence of identical trials. Specifically, for a repeatable A, the probability is P(A) = \lim_{n \to \infty} \frac{\text{number of occurrences of } A \text{ in } n \text{ trials}}{n}, assuming the limit exists and remains stable. This formulation emphasizes that probability quantifies the long-run proportion of successes in a hypothetical series of repetitions under the same conditions. Central to this interpretation is the idealization of infinite repetitions, which assumes that such sequences are theoretically possible even if practically unattainable due to finite resources or time. This abstraction allows for a precise mathematical treatment but applies only to events that can be conceptualized as part of a , repeatable . Unlike propensity interpretations, which attribute probability to underlying causal tendencies or dispositions of the experimental setup independent of actual frequencies, the frequentist approach adheres strictly to observed or hypothetical relative frequencies without invoking such physical propensities.

Scope and Limitations

Frequentist probability is best suited for scenarios involving repeatable experiments under stable conditions, where the long-run relative frequency of an event can be meaningfully estimated, such as in games of chance or controlled clinical trials. This approach defines probability as the limit of relative frequencies in an infinite sequence of independent trials, making it applicable to processes like coin flips or sampling from a fixed population, provided the underlying mechanism remains unchanged. A primary limitation of frequentist probability is its inability to directly assign probabilities to unique or one-off events, such as the likelihood of on a specific future date, since these cannot be repeated under identical conditions to observe frequencies. Similarly, it does not incorporate prior information or subjective beliefs about parameters, relying solely on observed data from the experiment itself. In practice, frequentist methods face challenges from finite samples, which inevitably introduce estimation errors due to sampling variability, even as sample sizes increase; the ideal of trials remains an unattainable used to justify asymptotic properties. For boundary cases involving non-repeatable events, frequentist approaches handle them indirectly by , embedding the event within a reference class of similar repeatable trials to infer a probability, though the choice of reference class can introduce ambiguity.

Historical Development

Origins in the 19th Century

The foundations of frequentist probability can be traced to the , where early probabilistic ideas emerged from analyses of games of chance, providing the groundwork for later frequency-based interpretations. Jacob Bernoulli's posthumously published (1713) introduced the , demonstrating that under repeated independent trials, the relative frequency of an event's occurrence converges to its theoretical probability, thus linking empirical observation to mathematical . further developed these concepts in his Théorie Analytique des Probabilités (1812), applying probabilistic methods to astronomical data and error analysis, emphasizing the stability of frequencies in large samples as a basis for in natural phenomena. These works shifted focus from purely combinatorial calculations to empirical regularities, setting the stage for a more data-oriented view of probability. In the , the frequentist interpretation began to formalize, with key contributions emphasizing probability as an objective limit of relative . Siméon Denis Poisson's Recherches sur la probabilité des jugements en matière criminelle et en matière civile (1837) applied probabilistic reasoning to empirical judicial , interpreting probabilities as stable derived from observed patterns in legal outcomes, thereby grounding abstract in real-world repetitions. advanced this in his Exposition de la théorie des chances et des probabilités (1843), explicitly defining the probability of an as the limit to which its relative approaches as the number of trials increases indefinitely, providing a rigorous empirical criterion for probabilistic statements. Robert Leslie Ellis further solidified the frequency interpretation in his paper "On the Foundations of the of Probabilities," presented in 1842 and published in 1844, where he argued that probability measures the limit of the relative of an event in a series of analogous trials, critiquing subjective elements in favor of inductive, observation-based foundations. This work, collected in The Mathematical and Other Writings of Robert Leslie (1863), emphasized the objective nature of probability through long-run . John Venn elaborated on these ideas in his 1866 book The Logic of Chance, providing a comprehensive defense of the by interpreting probability as the limiting relative in an infinite sequence of trials, which helped popularize the objective, empirical approach among philosophers and scientists. The rise of frequentist ideas reflected a broader empiricist shift in the sciences during the , driven by the industrial revolution's demand for statistical methods to analyze vital records, quality, and social data. This era's emphasis on observable, repeatable phenomena—exemplified by Quetelet's application of probability to social averages in Sur l'homme (1835)—moved probability from rationalist, a priori assumptions toward data-driven approaches, aligning with empirical philosophy's focus on from experience.

Key Figures and Advancements

played a foundational role in bridging empirical frequencies to in the early . In 1900, he introduced the as a measure of goodness-of-fit for categorical data, allowing researchers to assess whether observed frequencies deviated significantly from expected values under a , thus formalizing the use of relative frequencies in hypothesis testing. also developed the around the same period, providing a quantitative measure of linear association between variables based on frequency distributions, which facilitated early inferential methods in and social sciences. His work emphasized the interpretation of probability as long-run relative frequencies, laying groundwork for later frequentist developments. Émile Borel contributed to the frequency-based view in 1909 through his book Éléments de la théorie des probabilités, where he connected the law of large numbers to empirical probabilities observed in repeated trials, helping to establish probability as an objective limit of frequencies in practical applications. Ronald A. Fisher advanced frequentist probability through innovations in estimation and experimental design during the 1920s and 1930s. In his 1922 paper, Fisher proposed maximum likelihood estimation as a method to find parameter values that maximize the probability of observing the given data, assuming the underlying probability model, which became a cornerstone of frequentist inference for its efficiency in large samples. He further developed significance testing in his 1925 book Statistical Methods for Research Workers, introducing the concept of p-values to quantify the strength of evidence against a null hypothesis based on the frequency of extreme outcomes under repeated sampling. Fisher's emphasis on randomization in experimental design, detailed in his 1935 book The Design of Experiments, ensured that inferences about treatment effects relied on the stability of relative frequencies across randomized trials, minimizing bias and enabling causal interpretations in agriculture and biology. Richard von Mises advanced the theory in the 1920s and 1930s by axiomatizing probability as the limit of relative frequencies in infinite sequences known as "random collectives," which satisfy conditions of randomness and independence. His seminal 1928 German work Wahrscheinlichkeit, Statistik und Wahrheit, later translated as Probability, Statistics and Truth, provided a philosophical and mathematical foundation for frequentism by emphasizing empirical stability over subjective beliefs. Jerzy Neyman and Egon S. Pearson refined hypothesis testing and in the 1930s, establishing a decision-theoretic framework for frequentist statistics. Their 1933 Neyman-Pearson lemma provided a unified approach to constructing the most powerful tests for simple hypotheses by maximizing power subject to a controlled type I error rate, derived from likelihood ratios and long-run error frequencies. Neyman later introduced in 1937, defining them as intervals that contain the true parameter with a specified long-run across repeated samples, offering a frequentist alternative to for quantifying . Following , frequentist methods integrated deeply into modern statistics, particularly in and experimental design. applied Shewhart's control charts—rooted in frequentist monitoring of process frequencies—to postwar Japanese manufacturing, promoting to reduce variability and improve product quality, which spurred Japan's economic recovery. Fisher's randomization principles and Neyman-Pearson testing frameworks also expanded into industrial experimental design, enabling efficient optimization of processes through factorial designs and response surface methods, as seen in George Box's postwar adaptations that standardized frequentist approaches in engineering and .

Etymology and Terminology Evolution

The term "probability" originates from the Latin probabilitas, denoting likelihood, credibility, or the quality of being probable, entering English via probabilité in the mid-15th century. Its application evolved in the with the foundational mathematical developments in , such as the 1654 correspondence between and on games of chance, which began incorporating notions of repeatable events that later aligned with frequency-based interpretations. The designation "frequentist" emerged in the mid-20th century as a label to differentiate this approach from Bayesian methods, with the term first appearing in 1949 in the work of statistician Maurice G. Kendall, amid ongoing critiques of paradigms. Key terminology within frequentist probability also developed progressively: the concept of "relative frequency" was articulated by in his 1843 treatise Exposition de la théorie des chances et des probabilités, where he defined probability as the limit of the ratio of favorable outcomes to total trials in repeated experiments. Similarly, "confidence interval" was introduced by in 1937 through his paper "Outline of a Theory of Statistical Estimation Based on the Classical Theory of Probability," formalizing intervals that cover unknown parameters with a specified long-run success rate. These terms, including Ronald Fisher's 1925 introduction of the "p-value" in Statistical Methods for Research Workers as a measure of against a via tail probabilities under repeated sampling, have profoundly shaped modern statistical discourse by establishing a standardized lexicon for and hypothesis testing.

Core Concepts and Principles

Long-Run Frequency Approach

The long-run frequency approach defines probability as the limiting value of the relative frequency of an 's occurrence in an infinite sequence of repeated trials under fixed conditions. This interpretation, central to frequentist probability, posits that the probability P(A) of an event A is the stable ratio \lim_{n \to \infty} \frac{n_A}{n}, where n_A is the number of favorable outcomes and n is the total number of trials, assuming the sequence can be extended indefinitely while maintaining the . This view treats probability not as a static property but as an empirical regularity emerging from asymptotic behavior, applicable only to repeatable processes where such a limit exists. A classic conceptual example illustrates this approach with fair dice rolls, where the probability of rolling a six is \frac{1}{6}, as the empirical of sixes in a long sequence of independent throws converges to this value. In practice, after many rolls—say, thousands—the proportion of sixes stabilizes near \frac{1}{6}, reflecting the underlying of the die despite short-term fluctuations. This example underscores how the approach grounds probability in observable patterns rather than abstract assignments, with the long-run limit providing a benchmark for verification. The approach relies on key assumptions to ensure the limit's existence and stability: independence of trials, meaning each outcome does not influence subsequent ones; identical conditions across trials, such as using the same fair die; and stationarity, where the underlying process remains unchanged over time. These assumptions imply , allowing the in a single extended to represent the expected proportion across hypothetical repetitions. Without them, the relative may not converge, limiting the approach to well-defined, repeatable experiments. By tying probability to verifiable long-run data, this method promotes objectivity, as probabilities can be tested empirically rather than derived from subjective degrees of or assumptions. It ensures that claims about probabilities, such as the fairness of a , are falsifiable through accumulated evidence from trials, aligning frequentist probability with scientific . This verifiability distinguishes the approach from more interpretive views, emphasizing data-driven confirmation over personal judgment.

Relative Frequency and Convergence

In frequentist probability, the relative frequency of an event is defined as the proportion of trials in which the event occurs, formally expressed as f_n = \frac{k}{n}, where k is the number of occurrences of the event in n independent trials. This measure serves as an empirical approximation to the underlying probability p of the event. , in his foundational work, emphasized this definition within the framework of repeatable experiments, where the sequence of trials forms a "" allowing for such frequencies to be observed. The concept of convergence posits that, as the number of trials n increases, the relative frequency f_n oscillates around the true probability p but tends to stabilize and approach it in the limit. This idea, central to the frequentist interpretation, relies on the notion that sufficiently large samples provide reliable indicators of p, even though finite observations may show variability due to random fluctuations. Von Mises articulated this convergence as a key property for defining probabilities in empirical settings, distinguishing it from purely mathematical abstractions. A practical implication of this is that unknown probabilities can be estimated from finite datasets by computing the observed relative , which becomes increasingly accurate with more trials. For instance, in trials such as repeated coin flips, the relative of heads estimates the probability p of heads landing face up, improving as the number of flips grows from a few dozen to thousands. This approach underpins assessment in fields like and risk analysis, where direct repetition is but large samples approximate the .

Role in Statistical Inference

Frequentist statistical inference provides a framework for drawing conclusions about unknown parameters from sample by relying on the properties of sampling distributions, which describe the behavior of under repeated sampling from the same . In this approach, parameters are treated as fixed but unknown constants, and probabilities are interpreted as long-run frequencies of events in hypothetical repetitions of the sampling process. This enables the evaluation of estimators and test through their sampling distributions, allowing inferences that generalize from observed samples to the broader without incorporating subjective beliefs. A central role of frequentist methods lies in procedures for parameter estimation and hypothesis testing that operate solely on the data and the assumed sampling model. For estimation, techniques such as yield point estimates and interval estimates, like , which achieve a specified over repeated samples—for instance, a 95% covers the true in 95% of repeated experiments in the long run. Hypothesis testing, formalized in the Neyman-Pearson , compares competing by selecting tests that maximize while controlling error probabilities, ensuring decisions are based on objective criteria derived from the rather than personal priors. These methods emphasize and calibration across potential samples, distinguishing frequentist inference from purely Bayesian approaches that update beliefs with priors. The objectivity of frequentist inference stems from its focus on controlling error rates in the long run, such as the Type I error rate (the probability of rejecting a true ), typically set at a level like α = 0.05, and the Type II error rate (failing to reject a false null). Procedures are designed so that, under the assumed model, the overall error rates remain bounded across repeated applications, providing a guarantee of performance without reliance on subjective inputs. This error control is achieved by tailoring tests and intervals to the under specified conditions, ensuring inferences are defensible in terms of frequentist coverage and levels. Unlike descriptive statistics, which merely summarize observed data without generalization, frequentist inference explicitly moves to inductive reasoning about unseen cases by leveraging the stability of relative frequencies in large samples to estimate population characteristics. This shift enables statements about the likelihood of future observations or the reliability of generalizations, grounded in the convergence properties of estimators to true parameters as sample size increases. By focusing on the behavior of procedures over the sampling process, frequentist methods facilitate robust decision-making in fields requiring evidence-based conclusions from limited data.

Mathematical Foundations

Probability Spaces and Measures

In the frequentist approach to probability, the underlying is formalized through the theory of probability spaces, providing a rigorous measure-theoretic foundation for defining probabilities as limits of observable frequencies in repeatable experiments. This axiomatic framework was established by in his seminal 1933 work, which recasts probability in terms of and measure, enabling the treatment of both discrete and continuous cases. A probability space is defined as a triple (\Omega, \mathcal{F}, P), where \Omega represents the encompassing all possible outcomes of a random experiment, \mathcal{F} is a \sigma-algebra of measurable subsets of \Omega (events), and P: \mathcal{F} \to [0,1] is a satisfying the condition P(\Omega) = 1. The \sigma-algebra \mathcal{F} ensures that events are closed under countable unions, intersections, and complements, allowing for the systematic assignment of probabilities to complex event structures. The probability measure P adheres to three fundamental axioms: non-negativity, which requires P(E) \geq 0 for every E \in \mathcal{F}; , as noted; and countable additivity, stating that for a countable collection of pairwise disjoint events \{E_n\}_{n=1}^\infty \in \mathcal{F}, P\left( \bigcup_{n=1}^\infty E_n \right) = \sum_{n=1}^\infty P(E_n). These properties imply continuity of P, meaning that if a non-decreasing sequence of events E_n converges to E, then P(E_n) \to P(E), which supports the consistent handling of limiting processes in probability calculations. Within the frequentist instantiation, the is conceptualized for scenarios involving infinite sequences of independent, identically distributed trials, where P(E) quantifies the asymptotic relative frequency of E across such repetitions, aligning the abstract measure with empirical long-run frequencies. For instance, in the uniform modeling a flip, \Omega = \{\text{heads}, \text{tails}\}, \mathcal{F} = 2^\Omega (the power set), and P(\{\text{heads}\}) = 1/2, reflecting the expected frequency of heads in numerous tosses.

Laws of Large Numbers

The laws of large numbers (LLNs) are fundamental theorems in that underpin the frequentist interpretation by establishing the of empirical frequencies to theoretical probabilities under repeated trials. These results demonstrate how the average outcome stabilizes as the number of observations increases, providing a rigorous justification for interpreting probabilities as limiting relative frequencies. The Weak Law of Large Numbers (WLLN) asserts that for a sequence of and identically distributed (i.i.d.) random variables X_1, X_2, \dots, X_n with finite \mu = \mathbb{E}[X_i] and finite variance \sigma^2 = \mathrm{Var}(X_i) < \infty, the sample mean \bar{X}_n = \frac{1}{n} \sum_{i=1}^n X_i converges in probability to \mu as n \to \infty. Formally, \bar{X}_n \xrightarrow{P} \mu, meaning that for any \epsilon > 0, \lim_{n \to \infty} P(|\bar{X}_n - \mu| \geq \epsilon) = 0. This convergence holds due to the application of to the variance of the sample , which scales as \sigma^2 / [n](/page/N+). The finite variance condition ensures the probability of large deviations diminishes sufficiently fast. The Strong Law of Large Numbers (SLLN) provides a more robust guarantee, stating that under similar i.i.d. assumptions with finite \mu, the sample \bar{X}_n converges to \mu, i.e., P\left( \lim_{n \to \infty} \bar{X}_n = \mu \right) = 1. This almost sure convergence implies that deviations from \mu occur only on a set of probability zero. A seminal generalization for i.i.d. random variables, extending beyond finite variance in some cases, was proved by Andrey Kolmogorov in 1933. In frequentist probability, the LLNs offer the theoretical basis for the relative frequency approach, where the probability of an event is defined as the limit of its observed frequency in an infinite sequence of independent repetitions, ensuring consistency between empirical data and probabilistic models. This framework, as articulated by Richard von Mises, relies on the SLLN to validate the stability of such limits in practical statistical inference.

Confidence Intervals and Hypothesis Testing

In frequentist statistics, confidence intervals offer a method to estimate an unknown population parameter while quantifying the uncertainty associated with the estimate. Developed by , these intervals are constructed such that, over repeated random sampling from the population, a specified proportion—known as the confidence level—will contain the true parameter value. A 95% , for instance, implies that if the sampling and construction process were replicated infinitely many times, approximately 95% of the resulting would encompass the true ; however, for any single , the true value either lies within it or does not, with no probability attached to that specific case. This frequentist interpretation avoids assigning probabilities to parameters themselves, focusing instead on the long-run performance of the procedure. The width of the reflects the precision of the estimate, narrowing as sample size increases due to reduced variability, a property underpinned by the . For estimating the mean μ of a normally distributed with known standard deviation σ, based on a sample of size n with sample mean \bar{x}, the 95% is given by: \bar{x} \pm z_{\alpha/2} \frac{\sigma}{\sqrt{n}} where α = 0.05 and z_{0.025} ≈ 1.96 is the 97.5th of the standard , ensuring the interval covers μ with 95% confidence in repeated sampling. Hypothesis testing provides a frequentist approach to decision-making about population parameters by evaluating evidence against a null hypothesis H_0 in favor of an alternative H_1. The null hypothesis typically posits no effect or a specific value (e.g., equality), while the alternative suggests deviation from it; the test assesses whether observed data are compatible with H_0 under the assumption it is true. Central to many tests is the p-value, defined by Ronald Fisher as the probability, under H_0, of obtaining a test statistic at least as extreme as the one observed in the sample. A small p-value (e.g., below 0.05) indicates data unlikely under H_0, leading to its rejection, though Fisher viewed it as a measure of evidential strength rather than a strict decision rule. In contrast, the Neyman-Pearson framework formalizes tests to maximize statistical power—the probability of correctly rejecting H_0 when H_1 is true—while controlling error rates, as outlined in their 1933 lemma, which identifies the most powerful test for simple hypotheses via the likelihood ratio. Error control is fundamental: a Type I error occurs when H_0 is rejected despite being true, with probability α (the significance level, often set at 0.05); a Type II error happens when H_0 is not rejected but H_1 is true, with probability β, where power equals 1 - β. The Neyman-Pearson approach balances these by fixing α and minimizing β for a given sample size, ensuring long-run error frequencies align with these probabilities across hypothetical repeated tests. A classic example is testing the fairness of a , modeled as a experiment with success probability p (heads). The is H_0: p = 0.5 (), against the alternative H_1: p ≠ 0.5 (biased). For n = 100 tosses yielding k = heads, the is computed as twice the tail probability from the : 2 \times \sum_{i=60}^{100} \binom{100}{i} (0.5)^{100}, approximately 0.057 under . Since this exceeds α = 0.05, H_0 is not , suggesting insufficient of , though increasing n would enhance to detect deviations.

Comparisons and Criticisms

Differences from Bayesian Probability

Frequentist probability treats model parameters as fixed but unknown constants, whereas regards them as random variables characterized by probability distributions that incorporate prior information. This foundational distinction arises because frequentists avoid assigning probabilities to parameters themselves, viewing them as deterministic entities inferred solely from data, while Bayesians update beliefs about parameters using to derive posterior distributions. In terms of , frequentists define probability as the long-run relative of an in repeated trials under identical conditions, emphasizing over subjective assessment. Conversely, Bayesians interpret probability as a degree of that can be updated with new , allowing for subjective priors that reflect existing before observing . This leads to frequentist probabilities being tied to hypothetical ensembles of experiments, whereas Bayesian probabilities quantify directly about hypotheses or parameters. For statistical inference, frequentists rely on procedures like p-values and confidence intervals, which are calibrated based on their long-run performance across repeated sampling before data collection, providing guarantees about coverage but not direct probabilities for specific values. In contrast, Bayesians derive posterior distributions after observing , enabling credible intervals that directly represent the probability that the parameter lies within the interval given the and . P-values in frequentist analysis assess the compatibility of observed with a under the assumption it is true, while Bayesian posteriors offer a full probabilistic description of post-data. A illustrative example is estimating the of a , say the probability p of heads. In the frequentist approach, observing 7 heads in 10 tosses yields a 95% for p around the sample proportion (e.g., approximately [0.35, 0.93]), interpreted as containing the true fixed p in 95% of repeated samples, but not as a probability statement about p itself given this . Bayesian estimation, however, starts with a for p (e.g., uniform on [0,1] for ), updates it with the to form a posterior (Beta ), and computes a 95% (e.g., [0.36, 0.92]) that directly gives the probability the true p falls within it, conditional on the observed tosses and . This contrast highlights how frequentist intervals emphasize procedural reliability, while Bayesian intervals provide substantive probability updates.

Other Interpretations

The propensity interpretation views probability as an objective, dispositional property or tendency inherent in physical systems or experimental setups, rather than a mere frequency of outcomes. This approach, developed by , posits that probabilities measure the strength of a propensity for certain events to occur under specific conditions, allowing for objective assignments even in non-repeatable scenarios. For instance, the probability of a die landing on six reflects its physical propensity to do so, independent of observed frequencies. The logical interpretation treats probability as a measure of partial entailment or the degree of logical support that evidence provides for a hypothesis, akin to an extension of deductive logic into inductive reasoning. John Maynard Keynes introduced this view in his 1921 work, arguing that probabilities are objective relations between propositions, determined by rational analysis rather than empirical frequencies or personal beliefs. Rudolf Carnap further formalized it through inductive logic, developing systems where confirmation functions assign degrees of probability based on the logical structure of language and evidence, as outlined in his 1950 book. Subjective interpretations, distinct from full Bayesian frameworks, emphasize probability as a coherent degree of personal belief, where coherence ensures that beliefs avoid contradictions such as sure losses in betting scenarios. articulated this in his foundational work on subjective probability, focusing on operational definitions through fair betting odds without mandating Bayesian updating rules for all revisions. Here, probability reflects an individual's rational stance, calibrated to maintain consistency across judgments. Unlike the frequentist approach, which limits probabilities to long-run frequencies in repeatable experiments and thus avoids assigning them to unique or single events, these interpretations enable direct probability assessments for one-off occurrences. The propensity view attributes tendencies to specific situations, logical probability evaluates evidential support for hypotheses in isolation, and subjective probability captures beliefs about any proposition, thereby addressing frequentism's scope limitations for non-repeatable cases.

Common Critiques and Responses

One major critique of frequentist probability is its perceived inability to handle unique events or incorporate information effectively, as exemplified by . In this scenario, with a large sample size and a modest , frequentist testing may reject a point based on a below a conventional threshold (e.g., 0.05), while a Bayesian analysis with a diffuse on the favors the null, leading to conflicting conclusions. This paradox arises because frequentist p-values focus solely on the compatibility of data with the null without considering prior plausibility or model comparison, rendering the approach inadequate for one-off or non-repeatable events where priors could provide necessary context. Critics argue that this limitation stems from frequentism's rejection of subjective priors, forcing reliance on long-run frequencies that do not apply to singular analyses. Another common criticism involves the frequent misinterpretation of p-values as posterior probabilities of the null hypothesis being true. In frequentist inference, a p-value represents the probability of observing data as extreme or more extreme than the sample, assuming the null is true, but it is often erroneously equated with the probability that the null is correct or the posterior probability of the hypothesis given the data. This confusion leads to overstatement of evidence against the null, as p-values confound effect size with sample size and do not directly quantify belief in hypotheses, unlike Bayesian posterior probabilities. Such misinterpretations undermine the reliability of frequentist results in practice, particularly when p-values are used to claim definitive proof rather than assessing long-run error rates. Frequentist probability also faces criticism for its over-reliance on long-run frequencies, which are deemed irrelevant to single-case analyses. The interpretation defines probability as the limiting relative frequency in an infinite sequence of repeatable trials, but for unique events—such as a specific scientific experiment or policy decision—this framework provides no direct applicability, as the "long run" cannot be observed or replicated. Critics contend that this renders frequentist tools philosophically disconnected from real-world inference, where decisions must be made based on finite, non-repeatable data without hypothetical infinite repetitions. The approach's emphasis on asymptotic properties thus prioritizes theoretical guarantees over practical utility for isolated instances. In response to these critiques, proponents of frequentism emphasize the framework's focus on procedures with desirable long-run behavior, such as controlling error rates across repeated applications, which ensures reliability in scientific practice even if not directly interpretable for single events. For instance, confidence intervals and are designed to achieve nominal coverage probabilities (e.g., 95%) in hypothetical replications, providing a standard for without subjective priors. Regarding and prior incorporation, some hybrid approaches integrate or non-informative priors into frequentist testing to reconcile differences, approximating Bayesian credible intervals while maintaining frequentist guarantees. On p-value misinterpretations, defenders clarify that proper use treats them as measures of strength rather than posterior , and simulations show high agreement with Bayesian results under weak priors, mitigating interpretive issues. Overall, frequentism's strength lies in its empirical for repeated use, with misapplications attributed to rather than inherent flaws.

Applications

In Statistics and Data Analysis

In statistics and , frequentist approaches underpin many core methods for modeling relationships and testing differences in data. , particularly ordinary (OLS) , exemplifies this paradigm by treating model parameters as fixed but unknown quantities, with the goal of minimizing the sum of squared residuals to obtain unbiased and efficient under assumptions like , , and homoscedasticity. Developed by in 1809 as a method for adjusting observations in astronomical data, OLS provides point estimates and standard errors for about these fixed parameters, enabling predictions and assessments of variable significance through t-statistics and associated p-values. This fixed-parameter view contrasts with probabilistic assignments to parameters, emphasizing long-run frequency properties of the estimator across repeated samples from the same . Analysis of variance (ANOVA) and experimental design further illustrate frequentist principles, where Ronald Fisher's introduction of in the 1920s ensured unbiased by treating assignments as random to control for factors. In ANOVA, this randomization allows partitioning of total variance into components attributable to treatments and errors, with F-tests evaluating whether observed differences exceed what randomization alone would produce. Fisher's framework, detailed in his 1935 work on experimental design, supports the validity of p-values for testing null hypotheses of no treatment effects, facilitating robust conclusions in controlled studies like agricultural trials. This approach prioritizes the experiment's design to guarantee the procedure's error rates in hypothetical repetitions. For scenarios where parametric assumptions fail, non- tests offer distribution-free alternatives within the frequentist tradition. The Wilcoxon rank-sum test, introduced by Frank Wilcoxon in 1945, compares two independent samples by ranking observations and summing ranks from one group, yielding a whose under the of identical distributions does not rely on normality or equal variances. This method computes exact or approximate p-values to assess shifts in location, providing a powerful option when deviate from parametric ideals, such as in skewed distributions or small samples. Its distribution-free nature ensures controlled Type I error rates across a broad class of underlying distributions, making it a staple for . Frequentist tools are widely implemented in statistical software, enabling routine computation of p-values and intervals for . In , the base stats package's t.test() function performs tests and constructs intervals based on the t-distribution, returning p-values for one- or two-sample comparisons alongside interval estimates that cover the true parameter with the nominal probability in repeated sampling. Similarly, Python's library, through the scipy.stats module, supports frequentist testing with functions like ttest_ind() for samples, which output p-values and optional intervals derived from asymptotic or exact methods for small samples. These implementations standardize frequentist procedures, allowing analysts to apply them scalably in data pipelines while adhering to classical error control.

In Scientific Experiments and Physics

In experiments, frequentist methods are employed to establish claims through stringent levels, particularly in high-energy . The 5σ criterion, corresponding to a of approximately 2.87 × 10^{-7}, serves as the gold standard for declaring a new particle or phenomenon, ensuring that the probability of a false positive due to background fluctuations is exceedingly low. This threshold was pivotal in the 2012 announcement of the by the ATLAS and collaborations at CERN's , where the combined excess over background reached 5σ after analyzing collision events. Frequentist intervals further quantify uncertainties in signal strength, providing bounds on parameters like particle masses without assuming distributions. In , the frequentist interpretation aligns the with the notion of probabilities as long-run relative frequencies observed in repeated measurements on identically prepared systems. According to this view, the squared modulus of the wave function coefficient, |⟨ψ|φ⟩|^2, represents the limiting frequency with which a system in state ψ collapses to outcome φ upon measurement, as the number of trials approaches infinity. This perspective underpins experimental validations of quantum predictions, such as patterns in double-slit setups, where frequency ratios match Born probabilities across numerous runs, reinforcing the rule's empirical foundation without invoking subjective beliefs. Frequentist approaches dominate the design and analysis of randomized controlled trials (RCTs) in , particularly for assessing drug efficacy through . Power calculations determine the minimum sample size needed to detect a clinically meaningful with a specified (typically 80-90%) at a significance level of α=0.05, under the of no treatment difference. In phase III trials, this involves simulating the probability of rejecting the null when an of efficacy holds, as seen in evaluations of antihypertensive drugs where frequentist tests confirmed superiority over based on reductions. In astronomy, frequentist hypothesis testing is integral to exoplanet detection using the method, which monitors stellar light curves for periodic dips indicative of planetary orbits. The posits no transit (pure or stellar variability), while the alternative assumes a planetary signal; detection is assessed via test statistics like the Box-fitting periodogram, yielding p-values to reject the null. For instance, Kepler mission analyses applied this framework to identify thousands of candidates, confirming exoplanets when transit depths exceeded 5σ thresholds relative to photometric .

Practical Examples and Case Studies

In election polling, frequentist probability underpins the estimation of candidate support through random sampling and the calculation of as part of confidence intervals. For instance, a poll reporting 48% support for a candidate with a ±3% at the 95% confidence level implies that, if the poll were repeated many times under identical conditions, 95% of such intervals would contain the true . This frequentist interpretation treats the interval as a long-run measure of coverage accuracy, rather than a probability statement about the specific parameter value. A practical example from the 2012 U.S. polls showed aggregated results with margins reflecting sampling variability, where a 5-point lead between candidates had an effective margin of ±6%, highlighting how differences in proportions require wider intervals to account for in estimates. In , Shewhart control charts apply frequentist principles to monitor manufacturing processes by tracking cumulative frequencies of measurements against empirical control limits derived from sample . Developed by Walter Shewhart in the 1920s, these charts plot sample means (X-bar) and ranges over time, with upper and lower control limits set at ±3 standard deviations from the process mean, corresponding to a low probability (about 0.27%) of points falling outside under stable conditions assuming . This setup allows detection of special cause variation signaling process shifts, as seen in early applications at Bell Laboratories for inspecting telephone wire quality, where out-of-control signals prompted immediate adjustments to maintain consistency. For example, if successive sample averages exceed the upper limit, it indicates a non-random increase in defect rates, enabling corrective action based on the chart's probabilistic thresholds. In , frequentist testing of Hardy-Weinberg uses the chi-squared goodness-of-fit to assess whether observed and frequencies in a match expected proportions under assumptions of random mating and no selection. The test compares observed counts of homozygous and heterozygous genotypes against expectations calculated as p², 2pq, and q², where p and q are frequencies, yielding a chi-squared value that follows a with 1 degree of freedom under the . A significant result (e.g., p < 0.05) suggests deviation due to factors like structure. An illustrative case involves analyzing single nucleotide polymorphisms (SNPs) in human populations; for a locus with observed genotypes AA: 990, Aa: 2550, aa: 1460 in a sample of 5000 individuals, the expected counts under (with p ≈ 0.453) lead to a chi-squared of approximately 1.3, failing to reject and supporting stable frequencies. Such tests are routine in genome-wide association studies to validate data quality before further analysis. These examples illustrate how frequentist probability relies on finite sample approximations, such as the for confidence intervals in polling or normality assumptions in control charts, which provide reliable inferences for large datasets but can introduce bias in small samples where distributions deviate from ideals. A common pitfall is multiple testing, where conducting numerous hypothesis tests (e.g., across many genetic loci) inflates the ; for instance, performing 20 independent tests at α = 0.05 yields a 64% chance of at least one false positive without correction methods like Bonferroni adjustment, which divides α by the number of tests to maintain overall control. Addressing these requires careful sample sizing and adjustments to ensure the long-run frequency properties hold in practice.

References

  1. [1]
    Probability Statistics And Truth : Richard Von Mises - Internet Archive
    Jan 19, 2017 · Probability Statistics And Truth Book Source: Digital Library of India Item 2015.189506 dc.contributor.author: Richard Von MisesMissing: definition primary
  2. [2]
  3. [3]
    [PDF] Frequentism as a positivism: a three-tiered interpretation of probability
    Jul 29, 2013 · Abstract. I explore an alternate clarification of the idea of frequency probabil- ity, called frequency judgment. I then distinguish three ...
  4. [4]
    [PDF] Interpreting Probability - Assets - Cambridge University Press
    In Germany, the philosopher-physicist Richard von Mises was developing his own frequentist interpretation of probability, but also regarded the concept as ...
  5. [5]
    Overview of Frequentist Confidence Intervals
    The reasoning used in frequentist (classical) inference depends on thinking about all possible suitable samples of the same size n.
  6. [6]
    [PDF] The Frequentist School of Statistics Class 17, 18.05 - MIT Mathematics
    The main philosophical difference concerns the meaning of probability. The term frequentist refers to the idea that probabilities represent longterm frequencies ...
  7. [7]
    What is Probability?
    These are called frequentist statistics. The frequentist view is what gives credibility to standard estimates based on sampling. For example, if we choose a ...
  8. [8]
    Reading: Frequentist Probability - Grinnell Domains
    In this reading, we'll introduce the fundamental definitions of this frequentist perspective on probability theory as well as some key concepts: expectation and ...Missing: historical development authoritative sources
  9. [9]
    Probability - Stat 210a
    Frequentist answer: Probability represents a relative frequency of an outcome when an experiment is repeated many times. Bayesian answer: Probability ...
  10. [10]
    The Basics of the Bayesian Approach: An Introductory Tutorial
    A key tenet of the frequentist framework is the frequentist interpretation of probability: the probability that an event occurs is defined by the long-run ...<|control11|><|separator|>
  11. [11]
    [PDF] what is Probability?
    Apr 3, 2020 · Non-repeatable events do not have a probability. Von Mise's response: “theory...not logically definable but sufficiently exact in practice”.
  12. [12]
    [PDF] The Propensity Interpretation of Probability - Pasquale Cirillo
    Popper. Source: The British Journal for the Philosophy of Science , May, 1959, Vol. 10, No. 37. (May, 1959), pp. 25-42. Published by: Oxford University Press ...
  13. [13]
    [PDF] Error Statistics and the Frequentist Interpretation of Probability1
    Frequentist is the interpretation that identifies the probability of an event A with the limit of the relative frequency of its occurrence in a large number of ...
  14. [14]
    [PDF] Von Mises' Frequentist Approach to Probability - StatLit.org
    Richard Von Mises (1883-1953), the originator of the birthday problem, viewed statistics and probability theory as a science that deals with and is based on ...Missing: primary | Show results with:primary
  15. [15]
    (PDF) Single Trial Probability Applications: Can Subjectivity Evade ...
    Single Trial Probability Applications: Can Subjectivity Evade Frequency Limitations ... unique events. concerning their probability characteristics while making ...
  16. [16]
    Probability and statistics - Causes, Events, Outcomes | Britannica
    Sep 15, 2025 · Laplace argued in his Philosophical Essay on Probabilities (1825) that man's dependence on probability was simply a consequence of imperfect ...Missing: ideas | Show results with:ideas
  17. [17]
    Poisson Statistics - SNEWS: SuperNova Early Warning System
    Siméon Denis Poisson (1781-1840) introduced the probability concept which bears his name in the book Recherches sur la probabilité des jugements en matière ...
  18. [18]
    Grundbegriffe - Project Euclid
    Cournot (1843, page 78) seems to have been the first to say explicitly that probability theory does gain empir- ical meaning only by declaring events of ...<|separator|>
  19. [19]
    Robert Leslie Ellis and the Frequency Theory - SpringerLink
    In 1842, Robert Leslie Ellis presented a paper to the Cambridge Philosophical Society, entitled 'On the foundations of the theory of probability'.Missing: date | Show results with:date
  20. [20]
    (PDF) Ellis on Mathematical Statistics and Probability - ResearchGate
    12 Robert Leslie Ellis, “On the foundations of the theory of probabilities ... 1866), Ellis. placed the inductive scienti c method – found primarily in ...
  21. [21]
    The objective and the subjective in mid-nineteenth-century British ...
    This paper provides a critical discussion of the historical and theoretical meaningfulness of the distinction between 'objective' and 'subjective' probability.
  22. [22]
    [PDF] Karl Pearson a - McGill University
    From these ~1 groups I have found X 2 by the method of this paper. By this reduction of groups I have given. Sir George Airy's curve even a better chance than ...Missing: test | Show results with:test
  23. [23]
    On the mathematical foundations of theoretical statistics - Journals
    On the mathematical foundations of theoretical statistics. R. A. Fisher.
  24. [24]
    IX. On the problem of the most efficient tests of statistical hypotheses
    The problem of testing statistical hypotheses is an old one. Its origin is usually connected with the name of Thomas Bayes.
  25. [25]
    Outline of a Theory of Statistical Estimation Based on the Classical ...
    Outline of a Theory of Statistical Estimation Based on the Classical Theory of Probability. J. Neyman.
  26. [26]
  27. [27]
    Probability - Etymology, Origin & Meaning
    From mid-15c. Old French and Latin origin, "probabilite" means likelihood, credibility, or quality of being probable.
  28. [28]
    This Month in Statistics History: June - Amstat News
    Jun 2, 2025 · 1949 The term “frequentist” is first used by Maurice Kendall (ASA Fellow, 1950): “It might be thought that the differences between the ...
  29. [29]
    Exposition de la théorie des chances et des probabilités
    Oct 19, 2008 · Exposition de la théorie des chances et des probabilités. by: Cournot, A. A. (Antoine Augustin), 1801-1877. Publication date: 1843. Topics ...Missing: relative frequency
  30. [30]
    [PDF] Outline of a Theory of Statistical Estimation Based on the Classical ...
    0 (E), the confidence interval corresponding to the confidence coefficient ct. It is obvious that the form of the functions 0 (E) and 0 (E) must depend upon the.Missing: Jerzy | Show results with:Jerzy
  31. [31]
    Fisher (1925) Index - Classics in the History of Psychology
    STATISTICAL METHODS FOR RESEARCH WORKERS. By Ronald A. Fisher (1925) ... p. 75 · IV. TESTS OF GOODNESS OF FIT, INDEPENDENCE AND HOMOGENEITY; WITH TABLE OF ...
  32. [32]
    [PDF] Probability, Statistics and Truth
    RICHARD VON MISES. Probability, Statistics and Truth ... PROBABILITY, STATISTICS AND TRUTH meaning according as it is used scientifically or colloquially.Missing: interpretation | Show results with:interpretation
  33. [33]
    The Project Gutenberg eBook of The Logic of Chance by John Venn
    that the events really would occur with equal frequency in the long run. The ground of this assertion would probably be found in past experience, and it ...
  34. [34]
    [PDF] Ergodic Theory, Interpretations of Probability and the Foundations of ...
    He concludes that frequentist probability can only be applied to ergodic systems. Thus, the role ergodic theory now plays in the foundations of statistical ...
  35. [35]
    Interpretations of Probability - Stanford Encyclopedia of Philosophy
    Oct 21, 2002 · Applicability to frequencies: an interpretation should render perspicuous the relationship between probabilities and (long-run) frequencies.
  36. [36]
    Statistical Paradigms: Frequentist, Bayesian, Likelihood & Fiducial
    Frequentist Inference​​ Inference is based on the sampling distribution— how results would behave if we repeated the experiment many times. Key Tools: Confidence ...
  37. [37]
    [PDF] Stat 5102 Lecture Slides: Deck 2 Statistical Inference, Bias ...
    Frequentist statistics says sampling distributions are the correct measure of uncertainty of estimators. Frequentist probability says probabilities can only ...
  38. [38]
    [PDF] Review on Statistical Inference 5.1 Introduction 5.2 Frequentist ...
    The central idea of hypothesis test is to control the type-1 ... Thus, hypothesis test problem and confidence intervals are highly related to each other.
  39. [39]
  40. [40]
    [PDF] FOUNDATIONS THEORY OF PROBABILITY - University of York
    THEORY OF PROBABILITY. BY. A.N. KOLMOGOROV. Second English Edition. TRANSLATION EDITED BY. NATHAN MORRISON. WITH AN ADDED BIBLIOGRPAHY BY. A.T. BHARUCHA-REID.
  41. [41]
    [PDF] A Tricentenary history of the Law of Large Numbers - arXiv
    The Weak Law of Large Numbers is traced chronologically from its inception as Jacob Bernoulli's. Theorem in 1713, through De Moivre's Theorem, ...
  42. [42]
    7.1.1 Law of Large Numbers - Probability Course
    It states that if you repeat an experiment independently a large number of times and average the result, what you obtain should be close to the expected value.
  43. [43]
    [PDF] 8 Laws of large numbers - Arizona Math
    (Weak law of large numbers) Let Xj be an i.i.d. sequence with finite mean and variance. Let µ = E[Xj]. Then. Xn = 1 n n. X j=1. Xj → µ in probability. There ...
  44. [44]
    [PDF] An Appreciation of Kolmogorov's 1933 Paper - DTIC
    Jun 15, 1992 · In 1933, A. N. Kolmogorov (1933a) published a short but landmark paper in the Italian Actuarial Journal. He formally defined the empirical.
  45. [45]
    Hypothesis testing, type I and type II errors - PMC - NIH
    A type I error (false-positive) occurs if an investigator rejects a null hypothesis that is actually true in the population; a type II error (false-negative) ...
  46. [46]
    Binomial Hypothesis Testing | Real Statistics Using Excel
    Provides various examples demonstrating how to use Excel functions to perform hypothesis testing using the binomial distribution.
  47. [47]
    A Gentle Introduction to Bayesian Analysis - PubMed Central - NIH
    For example, the frequentist-based 95% confidence interval is often interpreted as meaning that there is a 95% chance that a parameter of interest lies between ...
  48. [48]
    Karl Popper: Philosophy of Science
    Popper defends a propensity theory of probability, according to which probabilities are interpreted as objective, mind-independent properties of experimental ...
  49. [49]
    [PDF] The Project Gutenberg eBook #32625: A treatise on probability
    Title: A Treatise of Probability. Author: John Maynard Keynes. Release Date: February 9, 2014 [EBook #32625]. Language: English. Character set encoding: ISO ...
  50. [50]
    [PDF] Logical Foundations of Probability
    PREFACE. The purpose of this work. This book presents a new approach to the old problem of induction and probability. The theory here developed is.
  51. [51]
    [PDF] BRUNO DE FINETTI - Foresight: Its Logical Laws, Its Subjective ...
    probability theory is not an attempt to describe actual behavior; its subject is coherent behavior, and the fact that people are only more or less coherent is.Missing: aspects | Show results with:aspects
  52. [52]
    De Finetti's Contribution to Probability and Statistics - Project Euclid
    Abstract. This paper summarizes the scientific activity of de Finetti in probability and statistics. It falls into three sections: Section 1 includes.<|separator|>
  53. [53]
    History and nature of the Jeffreys–Lindley paradox - PMC
    The Jeffreys–Lindley paradox exposes a rift between Bayesian and frequentist hypothesis testing that strikes at the heart of statistical inference.
  54. [54]
    Testing a Precise Null Hypothesis: The Case of Lindley's Paradox
    Jan 1, 2022 · Lindley's Paradox exposes a salient divergence between subjective Bayesian and frequentist reasoning when a parametric point null hypothesis is ...
  55. [55]
    Valid P-Values Behave Exactly as They Should: Some Misleading ...
    Among fair criticisms of P-values are that they are too easily confused with posterior probabilities, and that they are distortive evidence measures that need ...
  56. [56]
    [PDF] Twenty-One Arguments Against Propensity Analyses of Probability
    (Note also that the frequentist cannot explain the plausibility of their methodological precept, whereas the propensity theorist can.) This shifts probability ...
  57. [57]
    A Pragmatic View on the Frequentist vs Bayesian Debate | Collabra ...
    Aug 24, 2018 · Frequentists call such an interval a confidence interval, Bayesians call it a credible interval. These two types of intervals are, from a ...
  58. [58]
    Why do physicists mention “five sigma” in their results? - CERN
    Five sigma is considered the “gold standard” in particle physics because it guarantees an extremely low likelihood of a claim being false.
  59. [59]
    Five sigma revisited - CERN Courier
    Jul 3, 2023 · The standard criterion for claiming a discovery in particle physics is that the observed effect should have the equivalent of a five standard- ...
  60. [60]
    [1903.12027] Born's Rule from Quantum Frequentism - arXiv
    Mar 28, 2019 · Our result therefore demonstrates that purely unitary evolution can co-exist with Born's rule under more relaxed conditions. In addition, we ...Missing: frequentist mechanics
  61. [61]
    On the Concepts, Methods, and Use of “Probability of Success ... - NIH
    Jan 24, 2025 · For clinical trials, the power of the study is typically used to determine the sample size. Power, which tends to be set at 80% or 90%, is ...
  62. [62]
    Reassessment of Kepler's habitable zone Earth-like exoplanets with ...
    The standard frequentist approach toward hypothesis testing is to first design a test statistic: a number that is high in the presence of the signal and low ...
  63. [63]
    5 key things to know about the margin of error in election polls
    Sep 8, 2016 · In presidential elections, even the smallest changes in horse-race poll results seem to become imbued with deep meaning.
  64. [64]
    6.3.2. What are Variables Control Charts?
    Shewhart proposed a general model for control charts as follows: Shewhart Control Charts for variables, Let w be a sample statistic that measures some ...Missing: frequentist | Show results with:frequentist
  65. [65]
    12.6 - Hardy-Weinberg Equilibrium | STAT 555
    It is assessed by considering whether the percentage of samples having each genotype are in the correct proportion, which can be tested using a Chi-square ...Missing: frequentist | Show results with:frequentist
  66. [66]
    Common pitfalls in statistical analysis: The perils of multiple testing
    It can be calculated that if two groups are compared 5 times, the probability of a false positive finding is as high as 23%; if they are compared 20 times, the ...Missing: frequentist finite approximations