Fact-checked by Grok 2 weeks ago

Empirical process

In statistics and probability theory, an empirical process is a stochastic process defined on a class of functions \mathcal{F}, typically arising from an independent and identically distributed (i.i.d.) sample X_1, \dots, X_n drawn from a probability measure P, where the process is given by \sqrt{n}(P_n - P), with P_n f = n^{-1} \sum_{i=1}^n f(X_i) denoting the empirical measure applied to functions f \in \mathcal{F}. This formulation captures the centered and scaled fluctuations between the empirical distribution and the true underlying distribution, often converging weakly to a Gaussian process in the space \ell^\infty(\mathcal{F}) under suitable conditions on \mathcal{F}. The theory of empirical processes originated in the early 20th century with foundational results on of empirical distributions, notably the Glivenko-Cantelli theorem, which establishes that \sup_t |F_n(t) - F(t)| \to 0 almost surely for the empirical F_n and true distribution function F on the real line. This was extended by Donsker's invariance principle in 1952, proving of the empirical process to a for one-dimensional indices, laying the groundwork for functional central limit theorems in higher dimensions. Subsequent developments in the 1970s and 1980s, influenced by Vapnik-Chervonenkis theory and metric entropy methods, generalized these results to broader classes of functions, addressing complexities like measurability and tightness in Banach spaces. Empirical processes play a central role in modern nonparametric and semiparametric statistics, enabling the analysis of uniform rates of convergence for estimators such as the or kernel density estimates. Key applications include constructing confidence bands for distribution functions, validating bootstrap methods for complex dependence structures, and deriving efficiency bounds in models like the proportional hazards, where infinite-dimensional nuisance parameters are present alongside finite-dimensional targets. Advanced tools, such as chaining arguments, bracketing numbers, and maximal inequalities, facilitate proofs of asymptotic normality and for Z-estimators and M-estimators in high-dimensional settings.

Introduction

Definition

In probability theory, the empirical process arises in the study of the fluctuations of empirical measures around their population counterparts. Consider independent and identically distributed random variables X_1, \dots, X_n taking values in a sample space \mathcal{X} and drawn from an unknown probability measure P. The empirical measure P_n is defined as the random probability measure that assigns to each Borel set A \subset \mathcal{X} the mass P_n(A) = n^{-1} \sum_{i=1}^n 1_{\{X_i \in A\}}, where $1_{\{\cdot\}} denotes the indicator function. For a measurable function f: \mathcal{X} \to \mathbb{R}, this extends to P_n(f) = \int f \, dP_n = n^{-1} \sum_{i=1}^n f(X_i), providing a natural nonparametric estimator of the expectation Pf = \int f \, dP. The empirical process is then obtained by centering and scaling the empirical measure to capture its asymptotic behavior. Specifically, for a class of functions \mathcal{F} consisting of measurable functions f: \mathcal{X} \to \mathbb{R} (typically bounded or with finite Pf), the empirical process is the stochastic process \{G_n(f) : f \in \mathcal{F}\} indexed by \mathcal{F}, where G_n(f) = \sqrt{n} \bigl( P_n(f) - Pf \bigr) = n^{-1/2} \sum_{i=1}^n \bigl( f(X_i) - Pf \bigr). Equivalently, in terms of measures, G_n = \sqrt{n} (P_n - P), viewed as an element of a suitable function space such as \ell^\infty(\mathcal{F}), the space of bounded functions on \mathcal{F} equipped with the supremum norm. The scaling factor \sqrt{n} is chosen because, by the central limit theorem, the variance of n^{-1/2} \sum_{i=1}^n (f(X_i) - Pf) stabilizes to \mathrm{Var}_P(f(X_1)) as n \to \infty, facilitating asymptotic normality for fixed f. This formulation distinguishes the empirical process from the unscaled , which serves primarily as a without the needed for . In the finite-dimensional case, the process reduces to a of centered and scaled sums for a finite collection of functions, akin to a multivariate . However, the infinite-dimensional perspective treats G_n as a random element in a indexed by the possibly uncountable class \mathcal{F}, enabling results over \mathcal{F} and generalizations of classical limit theorems to functional settings.

Historical development

The theory of empirical processes originated in the early , drawing from foundational results in probability such as the , which provided the basis for understanding the convergence of sample averages to expected values. In 1933, published a seminal paper demonstrating the almost sure uniform convergence of the empirical distribution function to the true under continuity assumptions, marking an early milestone in shifting focus from pointwise to uniform convergence properties. Independently in the same year, Vladimir Glivenko established a similar result for the empirical distribution, emphasizing its uniform behavior across the real line and laying groundwork for broader empirical approximations. During the 1930s, Francesco Paolo Cantelli extended these ideas by proving the for arbitrary distribution functions, without requiring continuity, which formalized the Glivenko-Cantelli theorem as a cornerstone of empirical process theory. These early contributions highlighted the reliability of empirical measures in nonparametric settings and connected to central limit theorems by extending scalar convergence to functional forms. Following , the field advanced significantly with Michel Donsker's 1952 invariance principle, which showed that a properly scaled version of the empirical process converges weakly to a in the space of cadlag functions, bridging empirical deviations to Gaussian processes. This functional invigorated research by enabling asymptotic analysis of uniform statistics. The and 1980s saw expansive generalizations through the Vapnik-Chervonenkis (VC) theory, initiated by and Alexey Chervonenkis in their 1971 work on rates for empirical means over classes of events, introducing the VC dimension to control complexity in function spaces. Subsequent developments by Richard Dudley and others refined these tools for abstract index sets, solidifying empirical processes as a framework for high-dimensional and applications. Contemporary syntheses of the field appear in authoritative texts, including "Weak Convergence and Empirical Processes" by Aad W. van der Vaart and Jon A. Wellner (1996), which unifies convergence results, and "Introduction to Empirical Processes and Semiparametric Inference" by Michael R. Kosorok (2008), emphasizing practical statistical extensions.

Mathematical foundations

Empirical measure

The empirical measure is a fundamental concept in the study of empirical processes, serving as the nonparametric estimator of an unknown based on observed . Given a sample X_1, \dots, X_n of independent and identically distributed random variables taking values in a (\mathcal{X}, \mathcal{A}) with common distribution P, the P_n is defined as P_n = n^{-1} \sum_{i=1}^n \delta_{X_i}, where \delta_x denotes the Dirac measure concentrated at x \in \mathcal{X}. This definition positions P_n as a random on (\mathcal{X}, \mathcal{A}), assigning mass $1/n to each observation X_i. A key property of the empirical measure arises from the law of large numbers, which ensures that P_n(f) \to P(f) almost surely for every bounded measurable function f: \mathcal{X} \to \mathbb{R}. In particular, for an event A \in \mathcal{A}, the empirical measure evaluates to P_n(A) = n^{-1} \sum_{i=1}^n \mathbf{1}_{\{X_i \in A\}}, which provides the sample proportion as an unbiased estimator of P(A). This connection highlights how the empirical measure generalizes the sample mean to arbitrary sets or functions, facilitating estimation in abstract spaces beyond the real line. Regarding convergence modes, the empirical measure P_n converges weakly to P in probability when \mathcal{X} is a separable , meaning that \int f \, dP_n \to \int f \, dP in probability for all bounded continuous functions f on \mathcal{X}. This weak convergence forms the basis for more advanced results in empirical process theory, such as those involving centered and scaled versions of P_n. Unlike the empirical distribution function, which specializes to cumulative probabilities on \mathbb{R}, the empirical measure applies broadly to general probability spaces.

Empirical distribution function

The empirical distribution function, often denoted F_n, provides a nonparametric estimate of the (CDF) for real-valued random variables. Consider a sample of independent and identically distributed (i.i.d.) random variables X_1, \dots, X_n drawn from an unknown distribution with CDF F. The empirical distribution function is defined as F_n(x) = P_n((-\infty, x]) = \frac{1}{n} \sum_{i=1}^n 1_{\{X_i \leq x\}}, where P_n denotes the , which assigns mass $1/n to each observation X_i, and $1_{\{ \cdot \}} is the . This definition directly links F_n(x) to the order statistics of the sample. Let X_{(1)} \leq X_{(2)} \leq \dots \leq X_{(n)} be the ordered observations; then F_n(x) equals the proportion of data points less than or equal to x, specifically F_n(x) = k/n where k is the number of X_i \leq x. The function remains constant between consecutive order statistics and increases by $1/n at each X_{(j)} (with adjustments for ties). Graphically, F_n appears as a right-continuous , starting at 0 for x < X_{(1)}, jumping by $1/n (or multiples thereof in case of ties) at each distinct observed value, and reaching 1 for x \geq X_{(n)}. This stepwise form visually approximates the underlying CDF F based solely on the sample, without assuming any parametric form for F. Regarding asymptotic behavior, the law of large numbers implies that F_n(x) converges almost surely to F(x) pointwise for each fixed x, yielding consistency at individual points. However, uniform convergence in the supremum norm requires additional theoretical machinery beyond the basic law of large numbers.

Key theorems and properties

Glivenko-Cantelli theorem

The Glivenko-Cantelli theorem asserts that for independent and identically distributed real-valued random variables X_1, \dots, X_n with common cumulative distribution function F, the empirical cumulative distribution function F_n(x) = n^{-1} \sum_{i=1}^n \mathbf{1}_{\{X_i \leq x\}} satisfies \sup_{x \in \mathbb{R}} |F_n(x) - F(x)| \to 0 almost surely as n \to \infty. This result provides a uniform strong law of large numbers for the empirical distribution, extending the pointwise almost sure convergence of F_n(x) to F(x) for each fixed x. The theorem holds without assuming continuity of F and applies to any distribution on the real line. Originally proved by Glivenko for continuous distributions and extended by Cantelli to the general case, the theorem forms a cornerstone of empirical process theory. A standard proof begins by establishing almost sure convergence of F_n to F at dyadic points using the strong law of large numbers and maximal inequalities to control fluctuations between points. Uniform convergence over the real line then follows from the monotonicity of F_n and F, combined with the Borel-Cantelli lemma to show that the probability of large deviations sums to a finite value, ensuring almost sure boundedness. This approach leverages the submartingale property of the empirical process deviations. The theorem generalizes beyond the empirical distribution function to supremum convergence over classes of functions \mathcal{F} with finite Vapnik-Chervonenkis (VC) dimension, where \sup_{f \in \mathcal{F}} |n^{-1} \sum_{i=1}^n (f(X_i) - \mathbb{E}[f(X_1)])| \to 0 almost surely. Such classes exhibit combinatorial structure that bounds the complexity of empirical realizations, enabling uniform laws via VC theory. This extension, developed by , underpins modern applications in learning theory by controlling generalization error through uniform consistency. Regarding rates, the Glivenko-Cantelli theorem implies almost sure convergence but does not specify speed; however, the uniform deviation is of probabilistic order O_p(n^{-1/2}), as established by the Dvoretzky-Kiefer-Wolfowitz inequality, which bounds P(\sup_x |F_n(x) - F(x)| > \epsilon) \leq 2 e^{-2n \epsilon^2} for \epsilon > 0. For more general VC classes, the uniform rate depends on the entropy number of the class, typically scaling as O_p( ( \log n / n )^{1/2} ) under suitable conditions.

Donsker's theorem

Donsker's theorem, also known as the functional central limit theorem for empirical processes, establishes weak convergence of a suitably scaled empirical process to a Gaussian limiting process. In its classical form, consider independent and identically distributed random variables X_1, \dots, X_n with common continuous distribution function F supported on [0,1], and let F_n denote the empirical distribution function. The theorem asserts that the process \sqrt{n}(F_n(t) - F(t)), viewed as a random element in the space \ell^\infty[0,1] equipped with the supremum norm, converges weakly to a Brownian bridge B^0 as n \to \infty. The Brownian bridge B^0 is a zero-mean Gaussian process with covariance \mathbb{E}[B^0(s)B^0(t)] = \min(s,t) - st. This result extends to more general empirical processes indexed by classes of functions \mathcal{F}. Specifically, for i.i.d. observations from a P on a \mathcal{X}, define the empirical process G_n(f) = \sqrt{n}(P_n f - P f) for f \in \mathcal{F}, where P_n f = n^{-1} \sum_{i=1}^n f(X_i) is the empirical . Under suitable conditions on \mathcal{F}, such as the existence of a measurable F with \mathbb{P}(F > 1) < \infty and finite bracketing entropy integral \int_0^\infty \sqrt{\log N_{[]}(\epsilon, \mathcal{F}, L_2(P))} \, d\epsilon < \infty, the process G_n converges weakly in \ell^\infty(\mathcal{F}) to a mean-zero Gaussian process G with covariance structure \mathbb{E}[G(f)G(g)] = \mathrm{Cov}_P(f(X), g(X)) for f, g \in \mathcal{F}. Such classes \mathcal{F} are termed Donsker classes. The proof of Donsker's theorem combines finite-dimensional convergence with tightness of the sequence of processes. Finite-dimensional distributions of G_n converge to those of the Gaussian limit by the multivariate central limit theorem applied to the means P_n f for finite subsets of \mathcal{F}. Tightness in \ell^\infty(\mathcal{F}) is established using chaining arguments, which control the supremum of the process by decomposing it over dyadic levels of approximation and bounding oscillations via entropy numbers of \mathcal{F}. Alternatively, bracketing entropy conditions provide moment bounds on the uniform deviation \sup_{f \in \mathcal{F}} |G_n(f)| sufficient for tightness. This weak convergence has key implications for statistical inference, particularly in deriving the asymptotic distribution of functionals like \sup_{t \in [0,1]} |\sqrt{n}(F_n(t) - F(t))|, which converges to \sup_{t \in [0,1]} |B^0(t)| and underpins uniform goodness-of-fit tests. In the general setting, it facilitates asymptotic analysis of suprema over \mathcal{F}, such as \sup_{f \in \mathcal{F}} |G_n(f)|, whose limiting distribution is \mathbb{E}[\sup_{f \in \mathcal{F}} |G(f)|], enabling construction of uniform confidence bands and hypothesis tests.

Applications

Nonparametric statistics

Empirical processes play a central role in nonparametric goodness-of-fit testing, where they provide the theoretical foundation for assessing how well an empirical distribution matches a hypothesized one. The is a prominent example, evaluating the maximum deviation between the empirical cumulative distribution function F_n and a specified null distribution F_0. The test statistic is given by \sup_x |\sqrt{n} (F_n(x) - F_0(x))|, and under the null hypothesis, its asymptotic distribution arises from the weak convergence of the empirical process to a , as established by . Another key goodness-of-fit statistic is the Cramér-von Mises (CvM) criterion, which measures the integrated squared difference between the empirical and null distributions. Defined as n \int (F_n(x) - F_0(x))^2 dF_0(x), the CvM statistic converges in distribution to the integral of the square of a under the null, yielding a limiting distribution that is a weighted sum of independent chi-squared random variables with one degree of freedom. This convergence follows from the functional limit theorem for . In nonparametric density estimation, empirical processes enable uniform consistency results by bounding deviations over function classes. For kernel density estimation, where the estimator is \hat{f}_h(x) = \frac{1}{nh} \sum_{i=1}^n K\left(\frac{x - X_i}{h}\right) with bandwidth h and kernel K, uniform rates of convergence are derived using empirical process bounds on the integrals involved. These bounds, often relying on entropy conditions for the class of kernel functions, yield rates of order O_p(\sqrt{\frac{\log n}{nh}} + h^2) under suitable smoothness assumptions on the true density. Histogram estimators, which partition the support into bins and assign uniform density within each, achieve uniform consistency through Vapnik-Chervonenkis (VC) classes. The class of histogram functions over fixed partitions forms a VC class, allowing entropy controls that bound the supremum deviation \sup_x |\hat{f}(x) - f(x)| with rates depending on the number of bins and sample size. Adaptive partitions, selected via data-driven methods, maintain consistency by ensuring the effective VC dimension remains controlled, leading to near-optimal uniform rates of O_p\left(\sqrt{\frac{k \log n}{n}}\right) for k bins.

Semiparametric inference

In semiparametric models, which combine finite-dimensional parametric components with infinite-dimensional nonparametric elements, empirical processes provide essential tools for deriving efficient estimators and their asymptotic properties. These models allow for flexible specification of nuisance functions while focusing inference on parameters of interest, such as regression coefficients. Empirical process theory facilitates the analysis of estimators through weak convergence results, enabling the establishment of consistency, normality, and efficiency bounds under mild regularity conditions like and entropy bounds. Z-estimators form a cornerstone of efficient estimation in semiparametric settings, defined as solutions \hat{\theta}_n to the estimating equations \mathbb{P}_n \psi_{\theta, \eta}(X_i) = 0, where \mathbb{P}_n is the empirical measure, \theta is the finite-dimensional parameter, and \eta is the infinite-dimensional nuisance parameter. Under of the parameter map and conditions ensuring the function class \{\psi_{\theta, \eta} : \theta \in \Theta\} is a , the Z-estimator achieves \sqrt{n}-consistency and asymptotic normality: \sqrt{n}(\hat{\theta}_n - \theta_0) \rightsquigarrow -\dot{\Psi}_{\theta_0}^{-1} Z(\theta_0), where Z is a mean-zero Gaussian process arising from the . This normality follows from the applied to the zero-extraction map, with the inverse information matrix \dot{\Psi}_{\theta_0}^{-1} yielding the semiparametric efficiency bound when the \tilde{\psi}_{\theta_0, \eta_0} is used. The representation of Z-estimators often relies on partial sums and , which express the estimator as an average over an . Specifically, the influence function \tilde{\psi}_{\theta, \eta} captures the efficient score, and \sqrt{n}(\mathbb{P}_n - P)(\psi_{\theta_0, \eta_0}) \rightsquigarrow Z via the for empirical processes indexed by the parameter space. This decomposition allows for the projection of the estimating equation onto the tangent space of the nonparametric component, ensuring orthogonality and robustness to nuisance parameter estimation. A prominent example is the Cox proportional hazards model for survival data, where the hazard function is \lambda(t | Z) = \lambda_0(t) \exp(\beta_0^T Z), with \beta_0 parametric and \lambda_0 the nonparametric baseline hazard. The partial likelihood estimator \hat{\beta}_n solves a Z-estimating equation based on the score function, achieving \sqrt{n}(\hat{\beta}_n - \beta_0) \rightsquigarrow N(0, \mathcal{I}_{\beta_0}^{-1}), where empirical processes handle the nonparametric risk set sums in the denominator. For baseline hazard estimation, the Nelson-Aalen-type estimator \hat{\Lambda}_n(t) = \int_0^t \frac{dN_n(s)}{Y_n(s) \exp(\hat{\beta}_n^T \bar{Z}_n(s))} uses empirical process weak convergence to yield uniform consistency and \sqrt{n}(\hat{\Lambda}_n - \Lambda_0) \rightsquigarrow a Gaussian process, facilitating inference on cumulative hazards. In semiparametric models, convergence rates for nonparametric components are typically slower than the parametric \sqrt{n} rate due to the curse of dimensionality, often achieving n^{1/3} or similar under optimal smoothing. These rates are controlled by entropy conditions on the function classes, such as the uniform entropy integral \int_0^\infty \sqrt{\log N(\epsilon, \mathcal{F}, L_2(P))} d\epsilon < \infty, which ensure the empirical process fluctuations remain manageable and preserve asymptotic normality of the parametric part.

Examples and extensions

Basic uniform convergence example

To illustrate the Glivenko-Cantelli theorem, which states that for i.i.d. random variables X_1, \dots, X_n with common CDF F, the supremum norm \sup_x |F_n(x) - F(x)| converges to 0 almost surely as n \to \infty where F_n(x) = n^{-1} \sum_{i=1}^n \mathbf{1}_{\{X_i \leq x\}}, consider samples from the uniform distribution on [0,1]. The true CDF is F(x) = x for x \in [0,1]. For n=100 i.i.d. samples from this distribution, the empirical CDF F_n(x) forms a step function jumping by $1/100 at each ordered sample value. When plotted against the true line F(x) = x, F_n(x) typically stays close to the diagonal, demonstrating the uniform convergence. Simulations at varying sample sizes show the supremum deviation decreasing with n. The Dvoretzky–Kiefer–Wolfowitz inequality provides a quantitative bound: P(\sup_x |F_n(x) - F(x)| > t) \leq 2 e^{-2 n t^2} for any t > 0. For example, this implies that for n=100, the probability of the supremum exceeding approximately 0.14 is small. These observations highlight the almost sure as n grows. However, for small n such as 10, the fluctuations in F_n(x) emphasize the importance of uniform convergence theory, as pointwise convergence alone would not capture the maximum discrepancy over all x.

Bootstrap empirical processes

The bootstrap empirical process provides a resampling-based method to approximate the of empirical processes, particularly useful when asymptotic distributions are complex or unavailable. Given an independent and identically distributed (i.i.d.) sample of size n from an unknown P, the P_n is computed as the average of Dirac measures at the observations. A bootstrap sample is then drawn with replacement from this empirical measure, yielding the bootstrap empirical measure P_n^*. The associated bootstrap empirical process is defined as G_n^* = \sqrt{n} (P_n^* - P_n), considered conditionally on the original data, which mimics the centering and scaling of the original empirical process G_n = \sqrt{n} (P_n - P). This construction allows for Monte Carlo estimation of functionals of the empirical process by generating B independent bootstrap replicates and averaging over them. Under Donsker conditions—ensuring that the class of functions indexing the empirical process satisfies properties like finite bracketing entropy—the bootstrap empirical process G_n^* converges weakly, conditionally on the data, to the same tight Gaussian limit process as the unconditional G_n. This bootstrap for empirical processes guarantees the consistency of the bootstrap approximation with respect to the data-generating P. The result was rigorously established for general classes of functions by Giné and Zinn (), building on earlier work for specific cases like the uniform empirical process. A prominent application arises in hypothesis testing, such as approximating p-values for the Kolmogorov-Smirnov (KS) goodness-of-fit test, where the test statistic is the supremum norm \sup_t |G_n(t)| over the unit interval for the uniform empirical process. By computing \sup_t |G_n^*(t)| for each of B bootstrap samples and estimating the p-value as the proportion exceeding the observed KS statistic, the method yields reliable inference without relying on asymptotic tables, which can be inaccurate for small samples. This bootstrap KS test performs well in simulations and maintains nominal size across various distributions. The bootstrap approach excels in handling complex dependencies and non-i.i.d. settings through adaptations like block bootstrapping for , where consistency is again justified via empirical process theory. Its validity stems from the equivalence between the bootstrap measure and the true P in the large-sample limit, as proved in Giné and Zinn (1990), making it a versatile tool for semiparametric and robust .

References

  1. [1]
    [PDF] Introduction to Empirical Processes and Semiparametric Inference
    The goal of this book is to introduce statisticians, and other researchers with a background in mathematical statistics, to empirical processes and.
  2. [2]
    [PDF] EMPIRICAL PROCESSES: THEORY AND APPLICATIONS
    It also eliminates an unnecessary nota- tional distinction between empirical processes and partial-sum processes, bringing both closer to the theory for sums of ...
  3. [3]
    Weak Convergence and Empirical Processes - SpringerLink
    This book tries to do three things. The first goal is to give an exposition of certain modes of stochastic convergence, in particular convergence in ...
  4. [4]
    [PDF] EMPIRICAL PROCESSES: Theory and Applications
    g(Zn) →d g(Z) for any continuous function g : D(R, k·k∞) → R. 1. Page 4. 2. CHAPTER 1. EMPIRICAL PROCESSES: THEORY. Remark: ...
  5. [5]
    [PDF] Empirical Processes: Notes 1
    Mar 9, 2010 · Empirical Processes on General Sample Spaces: The modern theory of empirical processes aims to generalize the classical results to empirical ...
  6. [6]
    Kolmogorov, A.N. (1933) Sulla determinazione empirica di una ...
    Kolmogorov, A.N. (1933) Sulla determinazione empirica di una legge di distribuzione. Giornale dell' Istituto Italianodegli Attuari, 4, 83-91.Missing: Glivenko | Show results with:Glivenko
  7. [7]
    Empirical Processes (Chapter 19) - Asymptotic Statistics
    19 - Empirical Processes. Published online by Cambridge University Press: 05 June 2012. A. W. van der Vaart ... The empirical distribution function is defined as.
  8. [8]
    [2110.13236] A Constructive Proof of the Glivenko-Cantelli Theorem
    Oct 25, 2021 · The Glivenko-Cantelli theorem states that the empirical distribution function converges uniformly almost surely to the theoretical distribution for a random ...
  9. [9]
    On the Uniform Convergence of Relative Frequencies of Events to ...
    Vapnik, A. Ja. Červonenkis, The uniform convergence of frequencies of the appearance of events to their probabilities, Dokl. Akad. Nauk SSSR, 181 (1968), ...
  10. [10]
    Justification and Extension of Doob's Heuristic Approach to the ...
    In this paper we give a justification of Doob's approach to these theorems and show that the method can be extended to a wide class of such asymptotic theorems.
  11. [11]
    [PDF] 1 The Glivenko-Cantelli Theorem
    Under this condition, we can show using the Glivenko-Cantelli Theorem that Med( ˆFn) → Med(F) a.s. We will now prove this result. Suppose Fn is a (nonrandom) ...
  12. [12]
    Bootstrapping General Empirical Measures - Project Euclid
    It is proved that the bootstrapped central limit theorem for empirical processes indexed by a class of functions F F and based on a probability measure P P ...
  13. [13]
    Permutation and Bootstrap Kolmogorov-Smirnov Tests for the ... - jstor
    With this setup we have the following result concerning the limiting distribution of the permutation and bootstrap empirical processes. The proof is contained ...
  14. [14]
    Application of bootstrap method in Kolmogorov-Smirnov test
    The experimental results show that by using bootstrap method, not only the optimal model can be estimated but also the confidence of the K-S test conclusion can ...
  15. [15]
    The bootstrap for empirical processes based on stationary ...
    Giné and Zinn, 1990. E. Giné, J. Zinn. Bootstrapping general empirical measures. Ann. Probab., 18 (1990), pp. 851-869. Google Scholar. Hoffmann-Jørgensen, 1984.