Fact-checked by Grok 2 weeks ago

Classical Wiener space

The classical Wiener space is the probability space consisting of the separable W = C_0([0,1]; \mathbb{R}^d) of continuous functions w: [0,1] \to \mathbb{R}^d with w(0) = 0, equipped with the supremum norm \|w\|_\infty = \sup_{t \in [0,1]} |w(t)|, and endowed with the \mu, a centered Gaussian on W under which the coordinate process W_t(w) = w(t) is a standard d-dimensional starting at the origin. This construction provides a rigorous mathematical framework for modeling the paths of as random elements in a . Introduced by in his 1923 paper "Differential Space," the classical Wiener space formalized the existence of a version of with continuous sample paths, resolving earlier physical and probabilistic observations of irregular particle motion into a measure-theoretic object. A key feature is the Cameron-Martin space H, a separable of absolutely continuous functions h: [0,1] \to \mathbb{R}^d with square-integrable derivatives (i.e., H = \{ h : h(t) = \int_0^t \dot{h}(s) ds, \ \dot{h} \in L^2([0,1]; \mathbb{R}^d) \}, with inner product \langle h, g \rangle_H = \int_0^1 \dot{h}(s) \cdot \dot{g}(s) ds), which is densely and continuously embedded in W. The Wiener measure \mu is quasi-invariant under translations by elements of H, enabling the development of tools like the for changing measures. The classical Wiener space serves as a foundational model in stochastic analysis, underpinning Malliavin calculus for studying the regularity of random variables and solving stochastic differential equations, as well as applications in filtering theory (e.g., the Zakai equation), financial mathematics (e.g., option pricing via the Black-Scholes model), and infinite-dimensional measure theory. It also motivates the broader theory of abstract Wiener spaces (W, H, \mu), where W is a Banach space, H its Hilbertian subspace, and \mu a non-degenerate Gaussian measure, extending the classical case to more general Gaussian processes.

Definition and Foundations

Historical Development

The classical Wiener space emerged from efforts to mathematically model the erratic, continuous paths of , a phenomenon first observed by Robert Brown in 1827 and theoretically analyzed by in 1905 as the random displacement of particles suspended in a due to molecular collisions. This motivation drove the need to represent these paths not as deterministic curves but as random elements in an infinite-dimensional , enabling probabilistic analysis of their irregular, non-differentiable nature. Norbert Wiener laid the foundational work in his 1923 paper "Differential-Space," where he introduced a rigorous construction of as a with continuous sample paths. In this seminal contribution, Wiener defined a —now called the Wiener measure—on the space of continuous functions from [0, ∞) to the reals, ensuring independent Gaussian increments over disjoint time intervals while addressing the challenges of infinite-dimensional integration via Daniell's method. This approach marked the birth of the classical Wiener space, transforming physical observations into a precise mathematical framework. During the 1940s, and advanced the formalization of the Wiener measure on continuous paths, building on earlier insights to establish criteria for path regularity and extensions to broader stochastic processes. Kolmogorov's contributions, including the 1934 continuity theorem later refined in collaborative contexts, provided tools to verify the existence of continuous modifications for processes like , solidifying the measure's role in infinite-dimensional probability spaces. These developments, intertwined with wartime applications in prediction and filtering, elevated the Wiener space from a novel construct to a cornerstone of modern analysis.

Formal Construction of the Space

The classical Wiener space, denoted C_0[0,1], consists of all continuous functions f: [0,1] \to \mathbb{R}^d satisfying f(0) = 0. This space captures the sample paths of , providing a mathematical framework for representing the continuous trajectories of the process starting from the origin. As a , C_0[0,1] is equipped with pointwise (f + g)(t) = f(t) + g(t) and (\alpha f)(t) = \alpha f(t) for \alpha \in \mathbb{R} and t \in [0,1]. It forms a when normed by the supremum \|f\|_\infty = \sup_{t \in [0,1]} |f(t)|, ensuring completeness under . Elements of the space are typically denoted by \omega \in C_0[0,1], where each \omega symbolizes a realization of a continuous path in the modeling of stochastic processes like .

Topological Structure

Uniform Topology on C_0([0,1]; \mathbb{R}^d)

The uniform topology on C_0([0,1]; \mathbb{R}^d), the space of continuous \mathbb{R}^d-valued functions on the compact interval [0,1] with w(0) = 0, is generated by the supremum norm \|w\|_\infty = \sup_{t \in [0,1]} |w(t)|, where |\cdot| denotes the norm on \mathbb{R}^d. This norm measures the maximum length of the function values over the , providing a natural way to quantify the size of elements in the space. The supremum norm induces a metric d(w, \eta) = \|w - \eta\|_\infty = \sup_{t \in [0,1]} |w(t) - \eta(t)|, which defines the open sets of the uniform topology through the standard construction of balls B(w, \epsilon) = \{\eta \in C_0([0,1]; \mathbb{R}^d) : d(w, \eta) < \epsilon\}. Under this , the topology ensures that C_0([0,1]; \mathbb{R}^d) behaves as a complete metric space, where Cauchy sequences converge uniformly. Convergence in the uniform topology is equivalent to uniform convergence of function sequences: a sequence \{w_n\} \subset C_0([0,1]; \mathbb{R}^d) converges to w \in C_0([0,1]; \mathbb{R}^d) if and only if \sup_{t \in [0,1]} |w_n(t) - w(t)| \to 0 as n \to \infty. This form of convergence preserves the continuity of limits, as the uniform limit of continuous functions is continuous. The uniform topology is especially apt for continuous functions on [0,1] because every such function is uniformly continuous on the compact domain, allowing the supremum norm to capture global uniformity without local discrepancies. For Brownian paths in the classical , this topology is natural due to the paths' almost sure continuity and their specific modulus of continuity; in particular, standard satisfies |B_t - B_s| \leq \sqrt{2 d |t-s| \log(1/|t-s|)} with probability approaching 1 as |t-s| \to 0, ensuring that uniform convergence aligns with the paths' regularity properties.

Separability and Completeness

The classical Wiener space W = C_0([0,1]; \mathbb{R}^d), consisting of continuous \mathbb{R}^d-valued functions on the compact interval [0,1] with w(0)=0 equipped with the supremum norm \|w\|_\infty = \sup_{t \in [0,1]} |w(t)|, forms a complete metric space under the induced uniform topology. Completeness follows from the fact that C_0([0,1]; \mathbb{R}^d) is a closed subspace of the Banach space C([0,1]; \mathbb{R}^d): any Cauchy sequence \{w_n\} in C_0([0,1]; \mathbb{R}^d) converges uniformly to a limit function w with w(0)=0, since the pointwise limit of a uniformly Cauchy sequence of continuous functions is continuous and thus belongs to C_0([0,1]; \mathbb{R}^d). This uniform convergence preserves the continuity essential to the space, ensuring the limit lies within it. The space is also separable, possessing a countable dense subset. By the Stone-Weierstrass theorem, the algebra of polynomials with \mathbb{R}^d-valued coefficients (or equivalently, componentwise polynomials) is dense in C([0,1]; \mathbb{R}^d) with respect to the supremum norm, as it separates points and contains the constants in each component. Restricting to polynomials with rational coefficients yields a countable subalgebra that remains dense, providing the required countable dense set for the subspace C_0([0,1]; \mathbb{R}^d) by considering those vanishing at 0. (Stone, 1948) This separability and completeness have significant implications for infinite-dimensional analysis on the Wiener space, enabling the construction of countable bases for approximations and the application of techniques like spectral theory that rely on second-countability. In contrast, non-separable spaces such as L^\infty([0,1]; \mathbb{R}^d), the space of essentially bounded measurable \mathbb{R}^d-valued functions with the essential supremum norm, lack a countable dense subset—evidenced by the uncountable family of characteristic functions of distinct subintervals, which are at distance 1 from each other—highlighting why C_0([0,1]; \mathbb{R}^d) is particularly suited for probabilistic and analytical constructions. (Folland, 1999, p. 136)

Wiener Measure and Probability

Construction of the Wiener Measure

The Wiener measure is constructed on the space C_0([0,1]) of continuous functions \omega: [0,1] \to \mathbb{R}^d with \omega(0) = 0, endowed with the uniform topology, by defining its finite-dimensional distributions and ensuring their consistency for extension to the full space. This approach builds on the probabilistic structure of paths, where the measure assigns probabilities to sets of paths based on their behavior at finitely many time points. The Wiener measure is the product of d independent one-dimensional Wiener measures. The finite-dimensional distributions specify that, for any n \in \mathbb{N}, $0 \leq t_1 < \cdots < t_n \leq 1, the random vector (\omega(t_1), \dots, \omega(t_n)) \in \mathbb{R}^{n d} follows a multivariate Gaussian distribution with mean zero and covariance structure where the covariance between the components at times t_i and t_j is \min(t_i, t_j) I_d, with I_d the d \times d identity matrix. This covariance structure captures the independent increments property of , scaled by time, ensuring that the variance at time t is t I_d and covariances reflect shared path segments up to the minimum time, with independence across dimensions. These distributions are defined on the cylinder sets \{\omega \in C_0([0,1]) : (\omega(t_1), \dots, \omega(t_n)) \in B\}, where B is a Borel set in \mathbb{R}^{n d}. Consistency of these finite-dimensional distributions holds because the covariance function \min(s,t) I_d is continuous and positive semi-definite, satisfying the compatibility conditions: for any m < n and times $0 \leq s_1 < \cdots < s_m \leq 1, the marginal distribution on (\omega(s_1), \dots, \omega(s_m)) matches the specified Gaussian for those times. This consistency on nested cylinder sets allows the family of measures to be extended coherently. laid the groundwork for this measure by providing the first rigorous probabilistic model for , though the modern finite-dimensional approach formalized the marginal specifications. The Kolmogorov extension theorem then guarantees the existence of a probability measure \mu on the Borel \sigma-algebra \mathcal{B}(C_0([0,1])) generated by the uniform topology, such that the finite-dimensional distributions of the coordinate process under \mu coincide with the specified Gaussians. Specifically, for any n, times t_1, \dots, t_n, and Borel B \subset \mathbb{R}^{n d}, \mu(\{\omega : (\omega(t_1), \dots, \omega(t_n)) \in B\}) = the Gaussian probability of B. Uniqueness follows from the theorem, as any two measures agreeing on all cylinder sets (which generate \mathcal{B}(C_0([0,1]))) must coincide on the entire \sigma-algebra; this is the Wiener measure, providing the canonical probability space for standard d-dimensional Brownian motion on [0,1].

Properties of the Wiener Measure

The Wiener measure \mu on the space C_0([0,1]) of continuous functions from [0,1] to \mathbb{R}^d is supported entirely on the set of continuous paths, meaning \mu(C_0([0,1])) = 1. This holds despite the fact that almost all sample paths under \mu exhibit unbounded variation, as they are nowhere differentiable and possess a modulus of continuity that precludes bounded variation over any interval. A key scaling property of the Wiener measure arises from the self-similarity of : for any \lambda > 0, the transformed process defined by B(t) = \lambda^{-1/2} W(\lambda t), where W is a standard d-dimensional under \mu, has the same distribution as W. This invariance under time-space rescaling, \omega \mapsto \lambda^{-1/2} \omega(\lambda \cdot), preserves the measure \mu, reflecting the diffusive scaling inherent to the underlying Gaussian structure. The Wiener measure induces and independent increments on the coordinate process X_t(\omega) = \omega(t): for $0 \leq s < t \leq 1, the increment X_t - X_s is independent of the sigma-algebra generated by \{X_u : 0 \leq u \leq s\}, and its distribution is Gaussian with mean zero and covariance (t - s) I_d, depending only on the interval length. This property ensures that the process exhibits the Markovian behavior characteristic of Brownian motion. In finite dimensions, the marginal distributions of the Wiener measure are absolutely continuous with respect to Lebesgue measure on \mathbb{R}^{n d}, given explicitly by multivariate Gaussian densities derived from the covariance kernel \min(s,t) I_d. However, in the infinite-dimensional path space, the Wiener measure is singular, concentrating on an irregular subset where no equivalent Lebesgue-like measure exists, underscoring the measure's pathological support relative to smoother function spaces.

Coordinate Process and Maps

In the classical Wiener space, the coordinate process is defined as X_t(\omega) = \omega(t) for t \in [0,1] and \omega \in C_0([0,1]), the space of continuous functions on [0,1] to \mathbb{R}^d vanishing at 0, where this process is realized as a standard d-dimensional under the P. This canonical construction identifies the paths of with elements of the function space, providing a probabilistic framework for analyzing continuous-time stochastic processes. The evaluation maps \pi_t: C_0([0,1]) \to \mathbb{R}^d, given by \pi_t(\omega) = \omega(t), serve as the coordinate projections that extract the value of a path at time t. These maps are continuous with respect to the uniform topology on C_0([0,1]) and hence measurable with respect to the Borel \sigma-algebra generated by the open sets in this topology. Their measurability ensures that the coordinate process \{X_t\}_{t \in [0,1]} consists of random variables adapted to the natural filtration generated by the cylinder sets of the space. The finite-dimensional distributions of the coordinate process, for fixed times $0 \leq t_1 < \cdots < t_n \leq 1, are those of a multivariate Gaussian random vector (X_{t_1}, \dots, X_{t_n}) \in \mathbb{R}^{n d} with mean zero and covariance where the block corresponding to times s and t is \min(s,t) I_d. This Gaussian structure arises directly from the properties imposed by the Wiener measure, ensuring consistency across all finite collections of times and facilitating the extension to the full path measure. Through this setup, the coordinate process embeds the abstract definition of —characterized by independent Gaussian increments and continuous paths—directly into the path space C_0([0,1]), allowing the study of as a measurable process on a canonical probability space without reference to an underlying probability measure beyond the itself.

Key Properties and Substructures

Tightness and Compactness

Tightness plays a crucial role in the study of probability measures on the classical Wiener space C_0([0,1]), the space of continuous functions on [0,1] vanishing at the origin equipped with the uniform metric. A family \mathcal{P} of probability measures on C_0([0,1]) is said to be tight if for every \epsilon > 0, there exists a compact K \subset C_0([0,1]) such that P(K^c) < \epsilon for all P \in \mathcal{P}. This property ensures that the mass of the measures is concentrated on "small" sets in a probabilistic sense, facilitating convergence arguments. Prokhorov's theorem provides a fundamental link between tightness and compactness in the weak topology on the space of probability measures. Specifically, since C_0([0,1]) is a complete separable metric space, a sequence of probability measures \{P_n\} on C_0([0,1]) is tight if and only if it is relatively compact in the weak topology, meaning every subsequence has a further subsequence converging weakly to some probability measure on C_0([0,1]). This result, due to Billingsley, is pivotal in the construction and approximation of measures on infinite-dimensional spaces like the . The Wiener measure itself is tight on C_0([0,1]), which follows from probabilistic bounds on the of sample paths. The Kolmogorov-Chentsov criterion guarantees this by establishing that for the Brownian motion increments, \mathbb{E}[|W_t - W_s|^4] = 3 |t - s|^2 for t, s \in [0,1], satisfying the moment condition \mathbb{E}[|X_t - X_s|^\alpha] \leq C |t - s|^{1 + \beta} with \alpha = 4 and \beta = 1 > 0. This implies that for any \epsilon > 0, the probability that the w(\delta) = \sup_{|t-s| \leq \delta} |W_t - W_s| exceeds \epsilon tends to zero as \delta \to 0, uniformly under the Wiener measure, allowing construction of compact sets where paths remain close to zero at the origin and exhibit controlled oscillations. Compact subsets of C_0([0,1]) relevant to tightness are precisely those characterized by the Arzelà-Ascoli theorem. This theorem states that a F \subset C_0([0,1]) is relatively compact in the uniform topology if and only if it is bounded (i.e., \sup_{f \in F} \|f\|_\infty < \infty) and equicontinuous (i.e., for every \epsilon > 0, there exists \delta > 0 such that |t - s| < \delta implies |f(t) - f(s)| < \epsilon for all f \in F). In the context of Wiener measure tightness, such sets are formed by intersecting balls around the zero function with sets enforcing small moduli of continuity, ensuring the measure assigns high probability to these compacta.

Cameron-Martin Subspace

The Cameron-Martin subspace H, also known as the Cameron-Martin , is defined as the set of all functions h \in C_0([0,1]) that are absolutely continuous with square-integrable , i.e., H = \{ h \in C_0([0,1]) : h(t) = \int_0^t h'(s) \, ds, \, h' \in L^2[0,1] \}. This space is equipped with the inner product \langle h, g \rangle_H = \int_0^1 h'(t) g'(t) \, dt, which induces a \|h\|_H = \sqrt{\int_0^1 |h'(t)|^2 \, dt}. The Cameron-Martin theorem establishes that for any h \in [H](/page/H+), the translated Wiener measure \mu_h(A) = \mu(A - h) for Borel sets A \subseteq C_0([0,1]) is equivalent to the original Wiener measure \mu, meaning they are mutually absolutely continuous with a Radon-Nikodym given by the martingale \exp\left( \int_0^1 h'(t) \, d\omega(t) - \frac{1}{2} \|h\|_H^2 \right). Conversely, if h \notin [H](/page/H+), then \mu_h is singular with respect to \mu, implying that the supports of \mu_h and \mu are disjoint , which serves as the orthogonal complement characterization in the sense of measure-theoretic singularity. This subspace plays a pivotal role in the Girsanov theorem, which facilitates changing the to account for a deterministic drift in processes; specifically, if the drift function \mu \in L^2[0,1], corresponding to an element of H, the law of the process W_t + \int_0^t \mu(s) \, ds is absolutely continuous with respect to the measure, with the Radon-Nikodym derivative expressed via the Doléans-Dade exponential. The Cameron-Martin space H is dense in C_0([0,1]) with respect to the uniform , as polynomials vanishing at 0 (which belong to H) approximate any continuous path vanishing at 0 uniformly by the Stone-Weierstrass theorem.

Reproducing Kernel Hilbert Space

The reproducing kernel Hilbert space (RKHS) associated with the classical Wiener space is the Hilbert space H of functions on [0,1] in which point evaluation at any t \in [0,1] is a continuous linear functional, uniquely determined by the covariance kernel K(s,t) = \min(s,t) of the coordinate process under the Wiener measure. By the Moore–Aronszajn theorem, H (often denoted H_1) is the completion of the linear span of the functions e_t(\cdot) = \min(t, \cdot) for t \in [0,1], equipped with the inner product \langle f, g \rangle_H = \sum_{i,j} a_i b_j \min(t_i, s_j) whenever f = \sum_i a_i e_{t_i} and g = \sum_j b_j e_{s_j}. Equivalently, H_1 consists of all absolutely continuous functions h: [0,1] \to \mathbb{R} with h(0) = 0 and square-integrable derivative h' \in L^2([0,1]), under the inner product \langle h, g \rangle_{H_1} = \int_0^1 h'(u) g'(u) \, du. The reproducing property holds: for any h \in H_1 and t \in [0,1], h(t) = \langle h, K(\cdot, t) \rangle_{H_1}, where K(\cdot, t)(s) = \min(s, t). This structure ensures that H_1 embeds continuously and densely into the space C_0([0,1]) of continuous functions vanishing at 0, reflecting the Gaussian nature of the . For the on [0,1], H_1 is isometric to the Cameron-Martin subspace, providing the natural for translations under the Wiener measure. This isometry facilitates analysis of sample path properties, as Brownian paths are continuous but lack the required for membership in H_1. Additionally, H_1 forms the first Wiener chaos in the orthogonal decomposition of L^2 over the Wiener space, with higher chaoses constructed as symmetric tensor powers of H_1, enabling chaos expansions for smooth functionals of Brownian paths.

References

  1. [1]
    [PDF] Analysis on Wiener Space and Applications - arXiv
    Mar 13, 2016 · In particular, if we deal with the classical Wiener space, the path defined by. Tλ(w) = w + λη(w), is a Brownian motion under the new ...
  2. [2]
    [PDF] A guide to Brownian motion and related stochastic processes
    The modern mathematical treatment of Brownian motion (abbrevi- ated to BM), also called the Wiener process is due to Wiener in 1923 [436]. ... Differential space.
  3. [3]
    [PDF] The concept of velocity in the history of Brownian motion - arXiv
    Nov 2, 2021 · Wiener measure - Differential space 1923. In 1923, Wiener published Differential Space, which would later be cited as the article in which.
  4. [4]
    [PDF] probability theory - part 4 brownian motion - IISc Math
    Norbert Wiener took inspiration from Einstein's theory to ask ... This is why, his 1923 paper that introduced Brownian motion is titled Differential space.
  5. [5]
    Kolmogorov's Early Work on Convergence Theory and Foundation
    integral to give a rigorous treatment of the Brownian motion process, now accordingly sometimes called the Wiener process, but this work, like his pioneer-.
  6. [6]
    Prodigy of probability - MIT News
    Jan 19, 2011 · During World War II, Wiener received a government contract to help build a system that improved the accuracy of antiaircraft guns by predicting ...
  7. [7]
    Random fields and the geometry of Wiener space - Project Euclid
    pressed using F :C0[0,1] → R, where C0[0,1] is the space of continuous func- tions f :[0,1] → R, such that f(0) = 0, also referred to as the classical Wiener.
  8. [8]
    [PDF] Banach Spaces - UC Davis Mathematics
    The set {f ∈ C ([0, 1]) | f(0) = 0} is a closed linear subspace of C ([0, 1]), and is a Banach space equipped with the sup-norm. Page 5. 94. Banach Spaces.
  9. [9]
    Abstract Wiener spaces and applications to analysis - MSP
    Let C denote the space of real-valued continuous functions on [0,1] which vanish at zero, and let C* be the subspace of. C consisting of functions whose ...
  10. [10]
    [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. “ ...
  11. [11]
    [PDF] Brownian Motion and Stochastic Calculus - On-Line.It
    Brownian motion and stochastic calculus / loannis Karatzas, Steven. E. Shreve ... Modulus of continuity. 2.10. Solutions to Selected Problems. 2.11. Notes.
  12. [12]
    Completeness in Metric Spaces
    Theorem Let C[a,b] denote the normed linear space of continuous functions on the interval [a,b] equipped (as before) with the sup-norm, then C[a,b] is complete.
  13. [13]
    [PDF] PARTIAL SOLUTIONS TO REAL ANALYSIS, FOLLAND
    Nov 18, 2019 · This following are partial solutions to exercises on Real Analysis, Folland, written ... However, L∞(Rn,m) is not separable. Proof. We ...
  14. [14]
    [PDF] Wiener measure - USC Dornsife
    The Wiener space C[0,T] is a separable Banach space under the uniform norm. Separability comes from approximation by Bernestein polynomials of rational ...Missing: supremum classical sources
  15. [15]
    [PDF] Lecture 11 Stochastic Process and Brownian Motion
    The following theorem of Kolmogorov states that an abstract probability space (Ω,F,P) can be established for a stochastic process X by knowing its all finite ...
  16. [16]
    [PDF] gaussian processes; kolmogorov-chentsov theorem
    A standard (one-dimensional) Wiener process (also called Brownian motion) is a stochastic process {Wt}t≥0+ indexed by nonnegative real numbers t with the ...Missing: Centsov | Show results with:Centsov
  17. [17]
    Differential‐Space - Wiener - 1923 - Wiley Online Library
    Differential-Space. Norbert Wiener,. Norbert Wiener. Search for more papers by ... Download PDF. back. Additional links. About Wiley Online Library. Privacy ...
  18. [18]
    [PDF] Wiener Measure and Brownian Motion
    Wiener measure is a probability measure on paths undergoing Brownian motion, a random motion. It's a positive Borel measure on a space of paths.
  19. [19]
    [PDF] 212a1416 Wiener measure. - Harvard Mathematics Department
    Nov 11, 2014 · property that for almost all ω, the sample path X(t)(ω) is continuous. Let φ be a continuous function of compact support. Then the Riemann ...
  20. [20]
    Brownian Motion and Stochastic Calculus
    ### Summary of Wiener Space, Coordinate Process, Evaluation Maps, Measurability, and Finite-Dimensional Distributions from Karatzas and Shreve's "Brownian Motion and Stochastic Calculus"
  21. [21]
    [PDF] Lecture 21: Tightness of measures - MIT OpenCourseWare
    Nov 27, 2013 · Theorem 4 (Existence of Wiener measure). A sequence of measures Pn has a weak limit P∗ which satisfies the property of Wiener measure on C[0,T].
  22. [22]
    [PDF] The Arzel`a-Ascoli Theorem
    Feb 20, 2017 · The Arzel`a-Ascoli theorem is a foundational result in analysis, and it gives necessary and sufficient conditions for a collection of ...