Fact-checked by Grok 2 weeks ago

Convergence of measures

Convergence of measures is a central in measure theory and that formalizes the limiting behavior of sequences of measures on a , particularly probability measures on metric spaces. It encompasses several modes of convergence, with being the most prominent: a sequence of probability measures \{\mu_n\} converges weakly to a probability measure \mu if \int f \, d\mu_n \to \int f \, d\mu for every bounded continuous function f on the space. This definition, equivalent via the Portmanteau theorem to convergence on continuity sets (where \mu(\partial A) = 0), captures the idea that the measures become increasingly similar in their effects on smooth test functions. Other notable types include convergence in total variation, where the total variation distance d_{TV}(\mu_n, \mu) = \sup_A |\mu_n(A) - \mu(A)| \to 0, implying strong uniform agreement on all measurable sets and thus weak convergence. Set-wise convergence requires \mu_n(A) \to \mu(A) for every measurable set A, a stricter condition than weak convergence. For non-probability measures, vague convergence generalizes weak convergence to sub-probability measures on spaces like \mathbb{R}^d, often used in point process theory. These notions are metrized, for instance, by the Lévy-Prokhorov metric for weak convergence, which quantifies the infimum of \varepsilon > 0 such that \mu_n(A) \leq \mu(A^\varepsilon) + \varepsilon and \mu(A) \leq \mu_n(A^\varepsilon) + \varepsilon for Borel sets A, where A^\varepsilon is the \varepsilon-enlargement. The theory's importance stems from its role in proving asymptotic results, such as the , where empirical distributions converge weakly to a . Tightness is a key property ensuring relative compactness of measure families: for every \varepsilon > 0, there exists a compact set K with \mu_n(K) \geq 1 - \varepsilon for all n, which, by in spaces, guarantees the existence of weakly convergent subsequences. Applications extend to functional limit theorems like , approximating random walks by in the Skorokhod space D[0,1], and to empirical processes and martingales. The further propagates under continuous transformations, broadening its utility in stochastic analysis.

Overview and Motivation

Informal Descriptions

Convergence of measures shares intuitive parallels with the convergence of sequences of functions, where measures act like generalized "distributions" that approximate a limiting distribution in varying degrees of closeness. For instance, just as a sequence of functions might converge —matching the limit at every individual point but varying unevenly across the —measures can converge weakly, aligning averages over or continuous test functions without uniform control over all possible subsets. In contrast, uniform convergence of functions requires a consistent bound on the difference everywhere, akin to stronger forms of measure convergence that demand near-exact matching across the entire , such as in distance. This analogy highlights how weaker convergences suffice for broad behavioral similarities, while stronger ones capture precise structural preservation. A straightforward probability example illustrates this sharpening : consider a sequence of Bernoulli measures, each corresponding to a random variable taking value 1 with probability p_n and 0 otherwise, where p_n approaches 1 as n increases. Initially spread between 0 and 1, these measures gradually concentrate most of their mass near 1, resembling a narrowing peak that "sharpens" into a Dirac delta measure at 1 in the limit. Visually, one can picture the probability bar at 0 shrinking while the one at 1 grows, transforming a spread into a pinpoint focus, though the total mass stays fixed at 1 for these probability measures. This convergence demonstrates how distributions can evolve toward extreme concentration without altering their overall scale. Multiple notions of convergence arise because real-world applications demand different levels of rigor: strong convergence, like , ensures that integrals (or expectations) against all bounded measurable functions align closely, safeguarding details such as probabilities of irregular events, much like preserves function values uniformly. Weak convergence, however, only requires agreement for continuous bounded functions, allowing "smooth" integrals to match while potentially overlooking jumps or discontinuities—sufficient for many limit theorems in probability, where exact fine-grained control is unnecessary. This flexibility explains why is the most common type in probabilistic contexts, balancing tractability with essential distributional properties.

Historical Context and Applications

The concept of convergence of measures emerged in the mid-20th century as sought rigorous frameworks to handle limits of distributions in abstract spaces. A foundational result in stronger forms of convergence was provided by Scheffé's 1947 , which establishes that if a of probability densities converges to another , then the distance between the measures also converges to zero, linking convergence directly to measure convergence. This laid groundwork for understanding convergence when densities exist, influencing subsequent developments in . The theory advanced significantly with Prohorov's 1956 work, which introduced for probability measures on general metric spaces and proved a criterion (now known as Prohorov's theorem) relating tightness to relative in the . Building on this, Varadarajan extended the framework in 1958 to separable complete metric spaces (Polish spaces), showing that can be metrized and that the space of probability measures inherits desirable topological properties from the underlying space. These milestones formalized as a central tool, enabling proofs of limit theorems beyond finite-dimensional settings. Convergence of measures plays a pivotal role in applications across probability and statistics. For instance, weak convergence underpins modern proofs of the in metric spaces, allowing asymptotic normality for sums of random variables in general domains. It is essential in , where exponential rates of rare events are analyzed through weak limits of scaled measures. In , convergence concepts justify the existence of stationary distributions in Markov chains, facilitating long-term behavior analysis via invariant measures. In , convergence of measures justifies limit theorems for integrals with respect to approximating measures and underpins results for solutions to stochastic differential equations, where weak limits preserve path properties. Notably, induces a natural structure on spaces of probability measures, which is crucial for optimal transport problems, enabling stability of transport costs and plans under perturbations.

Strong Forms of Convergence

Convergence

Total variation convergence represents the strongest form of for measures, quantifying the maximum discrepancy in their assignments to measurable sets. For two finite signed measures \mu and \nu on a (\Omega, \mathcal{A}), the distance is defined as \|\mu - \nu\|_{TV} = \sup_{A \in \mathcal{A}} |\mu(A) - \nu(A)|. This distance equals half the of the signed measure \mu - \nu, that is, \|\mu - \nu\|_{TV} = \frac{1}{2} |\mu - \nu|(\Omega), where |\lambda| denotes the measure of a signed measure \lambda. Equivalent formulations highlight its dual interpretations. The distance can be expressed as \|\mu - \nu\|_{TV} = \sup_{\|f\|_\infty \leq 1} \left| \int f \, d(\mu - \nu) \right|, where the supremum is over all measurable functions f: \Omega \to \mathbb{R} bounded by 1 in . For probability measures absolutely continuous with respect to a measure \lambda with densities f_n and f, total variation convergence \|\mu_n - \mu\|_{TV} \to 0 is equivalent to \int |f_n - f| \, d\lambda \to 0. This mode exhibits robust properties. The norm induces a on the of finite signed measures, which is complete and thus a . in implies both setwise convergence, where \mu_n(A) \to \mu(A) for every measurable set A, and , where \int f \, d\mu_n \to \int f \, d\mu for all bounded continuous functions f. Scheffé's theorem provides a precise characterization for densities: a of probability densities f_n with respect to \lambda converges in to a density f if and only if \int |f_n - f| \, d\lambda \to 0 and \int f_n \, d\lambda \to 1. Since the f_n are densities, the second condition holds automatically, reducing to L^1 convergence. An illustrative example arises with densities on the unit interval [0,1]. Consider the target density f(x) = 2 \mathbf{1}_{[0,1/2]}(x), which is the on [0,1/2] renormalized. A sequence of approximating densities f_n(x) can be constructed as linear functions that rise linearly from 0 to 2 over [0,1/(2n)], remain constant at 2 until $1/2, and then fall linearly to 0 over [1/2, 1/2 + 1/(2n)], extending uniformly to 1 if needed; as n \to \infty, f_n converges uniformly to f on [0,1] (with supremum distance O(1/n)), implying L^1 convergence and thus total variation convergence of the corresponding measures.

Setwise Convergence

Setwise convergence of a sequence of signed measures \{\mu_n\} on a measurable space (X, \mathcal{A}) to a signed measure \mu is defined by the condition that \mu_n(A) \to \mu(A) for every set A \in \mathcal{A}. This preserves integrals of bounded measurable functions, as any such function can be approximated in the sup norm by simple functions, and the integrals of simple functions converge by from the on indicators of measurable sets. In the context of measures on metric spaces, setwise implies . A sufficient condition for setwise arises in the context of : if \mu_n \ll \mu for each n and the Radon-Nikodym derivatives f_n = d\mu_n / d\mu converge \mu- to f = d\mu / d\mu, with |f_n| \leq g for some g \in L^1(\mu), then the implies \int_A f_n \, d\mu \to \int_A f \, d\mu for every A \in \mathcal{A}, yielding setwise . Setwise convergence does not imply convergence in the total variation norm, as shown by counterexamples with oscillating measures; for instance, on [0,1] equipped with \lambda, the measures \mu_n with densities f_n(x) = 1 + \sin(2\pi n x) with respect to \lambda satisfy \mu_n(A) \to \lambda(A) for every measurable A by the Riemann-Lebesgue lemma, since \int_A \sin(2\pi n x) \, d\lambda(x) \to 0, but \|\mu_n - \lambda\|_{\mathrm{TV}} = \frac{1}{2} \int_0^1 |\sin(2\pi n x)| \, d\lambda(x) = \frac{1}{\pi} remains constant and positive. For signed measures, an example where the \mu_n^+ and \mu_n^- converge setwise to those of the limit but the does not converge in norm involves sequences where the supports of the parts separate in a way that prevents uniform control, such as certain Dirac combinations shifting positions.

Weak Convergence

Definition and Basic Properties

of measures, also known as convergence in distribution, is a fundamental concept in and measure theory that describes how a sequence of probability measures \mu_n on a topological space approaches a limiting measure \mu. In the context of a complete separable metric space (S, d), a sequence of probability measures \{\mu_n\} is said to converge weakly to \mu if, for every bounded continuous function f: S \to \mathbb{R}, \lim_{n \to \infty} \int_S f \, d\mu_n = \int_S f \, d\mu. This definition captures the idea that the expectations of continuous observables under \mu_n converge to those under \mu. The notion extends to more general topological spaces using bounded continuous functions C_b(S). This applies to spaces like \mathbb{R}^d, ensuring compatibility with the topology. A key basic property of weak convergence for probability measures (i.e., those with total mass 1) is the preservation of total mass. Since the constant function f \equiv 1 is bounded and continuous on any topological space, weak convergence implies \mu_n(S) = 1 \to \mu(S) = 1, preventing mass from escaping to infinity in the limit. Another fundamental property is the Skorokhod representation theorem, which provides a probabilistic realization of . If \{\mu_n\} converges weakly to \mu on a (a separable ), then there exists a (\Omega, \mathcal{F}, P) and random elements X_n, X: \Omega \to S such that the law of X_n is \mu_n, the law of X is \mu, and X_n(\omega) \to X(\omega) for P-almost every \omega \in \Omega. This theorem facilitates proofs by embedding the measures into a single space where almost sure convergence holds. On the specific space \mathbb{R}^d, weak convergence admits a useful characterization via characteristic functions. By Lévy's continuity theorem, \mu_n \to \mu weakly if and only if the characteristic functions \phi_n(t) = \int_{\mathbb{R}^d} e^{i \langle t, x \rangle} \, d\mu_n(x) converge to \phi(t) = \int_{\mathbb{R}^d} e^{i \langle t, x \rangle} \, d\mu(x) for all t \in \mathbb{R}^d. This equivalence leverages the fact that characteristic functions uniquely determine probability distributions and are themselves continuous bounded functions. A classic example of weak convergence arises in the (CLT). Consider independent identically distributed random variables X_1, X_2, \dots with mean 0 and variance \sigma^2 > 0. The standardized partial sums S_n / \sqrt{n}, where S_n = X_1 + \cdots + X_n, have distributions that converge weakly to the standard N(0, \sigma^2) on \mathbb{R}. This illustrates how weak convergence underlies the asymptotic normality of sample means, with the limiting preserving the total mass of 1. Weak convergence of measures is closely related to convergence in distribution of random variables, where the distribution of a sequence of random variables converges weakly to that of a limiting variable.

Convergence of Random Variables

In , the convergence in distribution of a sequence of random variables \{X_n\} to a limiting random variable X is defined as the weak convergence of their respective laws (or probability distributions): \mathcal{L}(X_n) \Rightarrow \mathcal{L}(X), where \mathcal{L}(Y) denotes the induced by Y on the underlying space. This notion bridges the measure-theoretic concept of weak convergence to probabilistic applications, providing a framework for asymptotic behavior in processes and . For real-valued random variables on \mathbb{R}, convergence in distribution is equivalent to pointwise convergence of the cumulative distribution functions: F_n(x) := P(X_n \leq x) \to F(x) := P(X \leq x) at every continuity point x of the limiting function F. However, this form of convergence is weaker than almost sure convergence or L^p convergence for p \geq 1, as it controls only the behavior of probabilities for bounded continuous events and does not preserve moments or pathwise properties in general. A key property arises from the continuity theorem for characteristic functions: if the characteristic functions \phi_n(t) := E[e^{itX_n}] converge pointwise to \phi(t) := E[e^{itX}] for all t \in \mathbb{R} and \phi is continuous at t=0, then X_n \to X in distribution (Lévy's continuity theorem, often associated with Helly-Bray principles in this context). Illustrative examples highlight these concepts. The binomial random variable X_n \sim \text{Binomial}(n, \lambda/n) converges in distribution to a random variable X \sim \text{Poisson}(\lambda) as n \to \infty, demonstrating the , which approximates rare events in large trials. Similarly, if X_n is uniformly distributed on [0, 1/n], then X_n \to X = 0 in distribution, where X is degenerate at 0, as the mass concentrates at the origin while the support shrinks. In the multidimensional setting on \mathbb{R}^d, of \mathcal{L}(X_n) to \mathcal{L}(X) holds if and only if the one-dimensional projections converge in distribution for all directions: t^\top X_n \to t^\top X in distribution for every t \in \mathbb{R}^d. This characterization, known as the Cramér-Wold theorem or projection theorem, reduces multivariate verification to univariate cases and is fundamental for central limit theorems in higher dimensions.

Portmanteau Theorem and Equivalent Characterizations

The Portmanteau theorem provides a collection of equivalent conditions for the of a of probability measures \{\mu_n\} on a to a limiting probability measure \mu. Specifically, \mu_n \to \mu weakly if and only if any one of the following holds: (i) \limsup_{n \to \infty} \mu_n(F) \leq \mu(F) for every closed set F; (ii) \liminf_{n \to \infty} \mu_n(O) \geq \mu(O) for every open set O; or (iii) \mu_n(A) \to \mu(A) for every Borel set A such that \mu(\partial A) = 0, where \partial A denotes the boundary of A. These set-based criteria offer practical tools for verifying weak convergence without directly computing integrals against all bounded continuous functions. Additional equivalent characterizations include the convergence \int f \, d\mu_n \to \int f \, d\mu for all bounded continuous functions f, or more restrictively, for all bounded continuous functions f. For probability measures on \mathbb{R}^d, this extends to convergence of the characteristic functions, though the latter is often treated separately. For positive measures on [0, \infty), weak convergence is equivalently characterized by pointwise convergence of the Laplace transforms \int_0^\infty e^{-t x} \, d\mu_n(x) \to \int_0^\infty e^{-t x} \, d\mu(x) for all t > 0, provided the total masses converge. These integral conditions leverage the density of continuous functions with compact support in the space of test functions for vague convergence, bridging to weak convergence under tightness. A high-level proof of the Portmanteau theorem relies on the regularity of Borel measures and the Stone-Weierstrass theorem to approximate indicator functions of continuity sets by bounded continuous functions, ensuring equivalence to the integral definition of . The implications from set conditions to integrals use monotone class arguments and applied to level sets \{f > t\}, while the reverse directions exploit the density of continuous functions and of the measures (inherent for probabilities). For non-probability measures, the conditions require supplementary or tightness to ensure the total masses converge, as the Portmanteau criteria alone may fail when mass escapes to . As a counterexample illustrating the necessity of tightness, consider \mu_n = \delta_n, the Dirac measure at n \in \mathbb{N}, on \mathbb{R}. Here, \limsup_n \mu_n(F) \leq 0 for any fixed closed F (taking \mu = 0), and similarly for open sets, but the sequence does not converge weakly to the zero measure without tightness, as the mass drifts to infinity and integrals against non-compact-support functions diverge. This highlights how varying total mass or lack of tightness invalidates the Portmanteau conditions for general measures.

Tightness and Prohorov's Theorem

In the context of weak convergence of probability measures on a , tightness provides a crucial criterion for ensuring that a family of measures has compact in the . A family of probability measures \{\mu_n\}_{n \in \mathbb{N}} on a X is said to be tight if, for every \varepsilon > 0, there exists a compact K \subset X such that \mu_n(X \setminus K) < \varepsilon for all n. This condition controls the mass escaping to infinity, preventing the measures from spreading out uncontrollably along subsequences. Tightness is particularly relevant in spaces like Polish spaces, where it facilitates the extraction of weakly convergent subsequences. Prohorov's theorem establishes the precise relationship between tightness and relative compactness in the weak topology. Specifically, on a Polish space (a separable completely metrizable topological space), a family of probability measures is relatively compact in the space of measures equipped with the weak topology if and only if it is tight. The direct implication—that tightness implies relative compactness—holds generally on Polish spaces, while the converse requires the space to be complete and separable to ensure that every relatively compact family is tight. This bidirectional equivalence underscores tightness as both a necessary and sufficient condition for the existence of weakly convergent subsequences, making it indispensable for proving weak convergence in infinite-dimensional settings. The weak topology on the space of probability measures can be metrized using the Prohorov metric, which quantifies the distance between measures in a manner compatible with weak convergence. The Prohorov metric is defined as d(\mu, \nu) = \inf\left\{\varepsilon > 0 : \mu(A) \leq \nu(A^\varepsilon) + \varepsilon \text{ for all closed } A \subset X \right\}, where A^\varepsilon denotes the \varepsilon-neighborhood of A, and the definition is symmetrized by interchanging \mu and \nu. On separable spaces, convergence in this metric is equivalent to of measures, thereby confirming the metrizability of the weak topology. This metric provides a concrete way to verify compactness and properties derived from Prohorov's theorem. An illustrative example of tightness arises in the study of empirical measures from i.i.d. samples. Consider i.i.d. random variables X_1, \dots, X_n in \mathbb{R}^d drawn from a P with finite second , \mathbb{E}[\|X_1\|^2] < \infty. The empirical measure \mu_n = n^{-1} \sum_{i=1}^n \delta_{X_i} then forms a tight family \{\mu_n\}_{n \in \mathbb{N}} in the weak topology, as the moment condition ensures that the mass does not escape to infinity with high probability. This tightness underpins the Glivenko-Cantelli theorem and Donsker's theorem for empirical processes, enabling weak convergence results in nonparametric statistics.

Vague Convergence

Vague convergence is a mode of convergence for measures defined on a locally compact , particularly suited for subprobability measures where the total mass need not be preserved. A sequence of measures \{\mu_n\} on such a space is said to converge vaguely to a measure \mu if \int f \, d\mu_n \to \int f \, d\mu for every continuous function f with compact support. This definition arises in the study of limits where mass may dissipate without accumulating at finite points, making it applicable to spaces like \mathbb{R}^d equipped with the . Vague convergence is weaker than weak convergence, as it tests against continuous functions with compact support rather than all bounded continuous functions, thereby allowing mass to escape to infinity while still capturing local behavior near compact sets. An equivalent characterization is that \limsup_n \mu_n(F) \leq \mu(F) for every compact set F. This property ensures that vague limits are unique and that the limiting measure is a , which is inner regular and finite on compact sets. For sequences of , vague convergence extends naturally via the Hahn-Jordan decomposition for signed measures, where positive and negative parts converge separately if the total variation satisfies the limsup condition on compacts. Additionally, vague convergence implies the convergence of integrals against polynomial test functions when restricted to compact supports, relating to the preservation of low-order moments under suitable growth controls, though higher moments may diverge due to mass loss at infinity. A representative example is the sequence of Gaussian measures \mu_n on \mathbb{R} with mean zero and variance n, which converges vaguely to the zero measure as n \to \infty. For any continuous function f with compact support, the integral \int f(x) \, d\mu_n(x) tends to zero because the density spreads out, concentrating negligible mass within the support of f./Lecture%20Notes.pdf) This illustrates how vague convergence accommodates the "escape to infinity" without requiring tightness.

Comparison Between Weak and Vague Convergence

Weak convergence of measures implies vague convergence, as the class of bounded continuous functions includes those with compact support. This inclusion holds in general metric spaces, but the converse requires additional conditions: vague convergence plus tightness of the sequence implies weak convergence. Tightness ensures that the measures do not lose mass at infinity, preventing the limit from having total mass less than the approximating sequence. A key difference between the two notions is that vague convergence does not preserve total mass, allowing sequences where mass escapes to infinity to converge to a limit with smaller mass. For instance, the sequence of Dirac measures \delta_n (unit mass at point n) converges vaguely to the zero measure on \mathbb{R}, since for any continuous function f with compact support, \int f \, d\delta_n = f(n) \to 0 as n \to \infty. However, this sequence does not converge weakly, as the total mass remains 1 while the vague limit has mass 0, violating the requirement for convergence against all bounded continuous functions. Full equivalence between weak and vague convergence, including portmanteau-type characterizations, typically requires the underlying space to be Polish (complete separable metric). Consider the sequence of uniform probability measures \mu_n on the interval [n, n+1], normalized to have total mass 1. This converges vaguely to the zero measure, as any compactly supported continuous test function eventually vanishes on [n, n+1] for large n. For probability measures, such vague convergence to a subprobability limit can be interpreted conceptually as weak convergence in the one-point compactification of the space, where the "missing" mass accumulates at the point at infinity. Vague convergence is particularly useful for analyzing point processes, where measures may have unbounded support or intensity, as in Poisson point process limits or cluster processes. It also applies to diffusions with possible explosion (finite-time blow-up to infinity), capturing local behavior while allowing mass loss at infinity without normalizing to probabilities. In contrast, weak convergence is preferred for sequences of normalized probability measures on compact or tight families, ensuring preservation of total probability mass. The distinction arises from the test function classes: weak uses bounded continuous functions, while vague restricts to compactly supported ones.

Weak Convergence in Dual Spaces

The space of probability measures on a topological space X, denoted \mathcal{M}(X), can be viewed as a subset of the dual space of C_b(X), the Banach space of bounded continuous real-valued functions on X equipped with the supremum norm \|\cdot\|_\infty. Each probability measure \mu \in \mathcal{M}(X) acts as a linear functional on C_b(X) via integration: \mu(f) = \int_X f \, d\mu for f \in C_b(X). The total variation norm on this dual space, defined as \|\mu\|_{TV} = \sup \{ |\mu|(A) - |\mu|(A^c) : A \subset X \}, where |\mu| is the total variation measure, endows \mathcal{M}(X) with a norm under which it is a closed convex subset of the unit ball. This embedding highlights the functional analytic structure underlying measure convergence, where the norm topology corresponds to total variation convergence. Weak convergence of a sequence of probability measures \mu_n \to \mu in \mathcal{M}(X) is equivalent to weak-* convergence in the dual of C_b(X), meaning \int f \, d\mu_n \to \int f \, d\mu for every f \in C_b(X). This is convergence in the weak-* topology \sigma(\mathcal{M}(X), C_b(X)), the coarsest topology making all functionals in C_b(X) continuous. The weak-* topology ensures that the unit ball in the dual is metrizable on relatively compact subsets when X is Polish, facilitating the study of sequential compactness. Alaoglu's theorem asserts that the closed unit ball in the dual of any Banach space, including C_b(X)^*, is compact in the weak-* topology. For probability measures, this compactness result underpins Prohorov's theorem: a family of measures in \mathcal{M}(X) is relatively compact in the weak-* topology if and only if it is tight, meaning for every \varepsilon > 0, there exists a compact K \subset X such that \mu(X \setminus K) < \varepsilon for all \mu in the family. The Banach-Alaoglu theorem thus provides the topological foundation for Prohorov's characterization of weak convergence via tightness in separable metric spaces. This dual perspective generalizes to signed measures, where the space ba(X) of bounded finitely additive signed measures on the power set of X (or regular Borel measures on Borel \sigma-algebras) coincides with the full dual of C_b(X) under the total variation norm. Vague convergence of signed measures in ba(X) corresponds to weak-* convergence against C_c(X), the space of continuous functions with compact support, allowing for potential mass escape to infinity unlike the tight control in weak convergence. This distinction arises because C_c(X) is a proper subspace of C_b(X), inducing a coarser topology suitable for non-probability measures.

References

  1. [1]
    Convergence of Probability Measures, 2nd Edition | Wiley
    A new look at weak-convergence methods in metric spaces-from a master of probability theory. In this new edition, Patrick Billingsley updates his classic work.
  2. [2]
    [PDF] Summary1 of Convergence for Probability Measures. - USC Dornsife
    Jun 11, 2023 · Probability measures can converge set-wise, in total variation, or weakly. Weak convergence means limn→∞ µn[f] = µ[f] for all bounded ...
  3. [3]
    245A, Notes 4: Modes of convergence - Terry Tao - WordPress.com
    Oct 2, 2010 · As an undergraduate, one learns of the following two basic modes of convergence: Uniform convergence implies pointwise convergence, but not conversely.
  4. [4]
    [PDF] Weak Convergence of Measures: - UC Davis Math
    The theory of weak convergence of probability measures on metric spaces provides such conditions. When applied in appropriate spaces of functions, the theory ...
  5. [5]
    Convergence of Random Processes and Limit Theorems in ...
    The convergence of stochastic processes is defined in terms of the so-called “weak convergence” (w. c.) of probability measures in appropriate functional spaces ...
  6. [6]
    On the Convergence of Sample Probability Distributions - jstor
    Vabadarajan, V. S. (1958) : Weak convergence of measures in separable metric spaces. Sankhy?, 19, 15. Paper received : October, 1957. 26.
  7. [7]
    [PDF] Total variation distance between measures
    Feb 15, 2005 · When P is dominated by Q we can choose λ equal to Q, giving D(P Q) = Q(p0 log p0), where p0 := dP/dQ. The following properties of Hellinger ...Missing: formulations | Show results with:formulations
  8. [8]
    A Useful Convergence Theorem for Probability Distributions
    A useful convergence theorem for probability distributions. Henry Scheffe. Download PDF + Save to my library. Ann. Math. Statist. 18(3): 434-438.
  9. [9]
    [PDF] 11 Radon-Nikodym derivatives - 11.1 Signed measures
    Notice that the finite signed measures on a measurable space (X,M) form a vector space, denoted M(X). Theorem 11.2. M(X) is a complete normed space under the ...Missing: metrizes | Show results with:metrizes
  10. [10]
  11. [11]
    [PDF] Convergence of Probability Measures - Ceremade
    Convergence of probability measures / Patrick Billingsley. - 2nd ed. Probability and statistics) p. cm. - (Wiley series in probability and statistics. “ ...
  12. [12]
    [PDF] Convergence of Probability Measures - CERMICS
    Convergence of probability measures / Patrick Billingsley. - 2nd ed. Probability and statistics) p. cm. - (Wiley series in probability and statistics. “ ...
  13. [13]
    [PDF] Lecture 7 Weak Convergence
    Nov 17, 2013 · Weak convergence, or convergence in distribution, depends on distributions. A sequence of probability measures converges weakly if ∫ fdµn → ∫ ...
  14. [14]
    [PDF] Contents
    ... weak convergence, we have a 'Portmanteau theorem'. Theorem 2.3.3. The ... Laplace transform of random variables is equivalent to their weak convergence.
  15. [15]
    [PDF] Weak Convergence of Probability Measures - arXiv
    Jul 20, 2020 · Theorem 1.16 Portmanteau's theorem. The following five statements are equivalent. (i) Pn ⇒ P. (ii) R f(x)Pn(dx) → R f(x)P(dx) for all ...
  16. [16]
    Yu. V. Prokhorov, “Convergence of Random Processes and Limit ...
    Convergence of Random Processes and Limit Theorems in Probability Theory, Teor. Veroyatnost. i Primenen., 1:2 (1956), 177–238; Theory Probab. Appl., 1:2 (1956) ...
  17. [17]
    [PDF] A Gentle Introduction to Empirical Process Theory and Applications
    where δx denotes the Dirac measure at x. For each n ≥ 1, Pn denotes the ... Example 8.6 (When Xi's are i.i.d. Bernoulli). Suppose that Xi's are i.i.d. ...
  18. [18]
    [PDF] Probability and Measure - University of Colorado Boulder
    Measure and integral are used together in Chapters 4 and 5 for the study of random sums, the Poisson process, convergence of measures, characteristic functions ...
  19. [19]
    [PDF] Vague and weak convergence of signed measures - arXiv
    Aug 2, 2022 · The total variation of a signed measure µ is denoted by kµk := |µ| (Ω), and we say that µ is finite if kµk < ∞. A finite signed measure µ on (Ω, ...<|separator|>