Fact-checked by Grok 2 weeks ago

Weak topology

In mathematics, the weak topology on a topological space X with respect to a family of continuous maps \{f_\alpha: X \to Y_\alpha\}_{\alpha \in A} from X to other topological spaces \{Y_\alpha\}_{\alpha \in A} is defined as the coarsest topology on X (i.e., the initial topology) that renders each f_\alpha continuous. This construction ensures that a subbasis for the weak topology consists of sets of the form f_\alpha^{-1}(U_\alpha), where U_\alpha is open in Y_\alpha, making it the weakest topology compatible with the given maps. In the specific context of topological vector spaces (TVS), the weak topology—often denoted \sigma(X, X^*)—arises when the family consists of all continuous linear functionals from the X^* to the scalars (typically \mathbb{R} or \mathbb{C}). It is generated by the seminorms \rho_\mu(x) = |\mu(x)| for \mu \in X^*, and is locally convex and Hausdorff provided X^* separates points on X (as guaranteed by the Hahn-Banach theorem for normed spaces). The weak topology is strictly coarser than the original on X (e.g., the norm topology in Banach spaces), meaning every weakly open set is originally open, but not conversely; for instance, in infinite-dimensional spaces, norm-open balls are not weakly open. A related but distinct notion is the weak topology* on the dual space X^*, denoted \sigma(X^*, X), which is the initial topology induced by the evaluation maps ev_x: X^* \to \mathbb{K} given by ev_x(\mu) = \mu(x) for x \in X. This topology is weaker than the weak topology restricted to X^* and plays a crucial role in functional analysis, as the closed unit ball in X^* is compact in the weak* topology by the Banach–Alaoglu theorem, enabling the study of bounded operators and duality in infinite-dimensional settings. If X is separable, the weak* topology on the unit ball is metrizable. Weak convergence in these topologies is characterized pointwise: a net x_\lambda converges weakly to x in X if \mu(x_\lambda) \to \mu(x) for all \mu \in X^*, while weak* convergence in X^* requires \mu_\lambda(x) \to \mu(x) for all x \in X. These topologies are fundamental in for proving existence results (e.g., via ) without relying on stronger norms, and they extend naturally to more general settings like product topologies or constructions.

Historical Development

Origins and Early Concepts

The study of weak topology traces its roots to early 20th-century efforts in solving integral equations and problems in the , where traditional proved insufficient for handling infinite-dimensional function spaces. In 1903, Ivar Fredholm developed a general theory for linear integral equations of the second kind, expressed as \phi(s) = f(s) + \lambda \int K(s,t) \phi(t) \, dt, introducing concepts like the Fredholm determinant and resolvent kernel that necessitated examining pointwise limits rather than ones to ensure solvability in broader classes of functions. This approach highlighted the limitations of classical in addressing operator invertibility and spectral properties. David Hilbert built upon Fredholm's ideas between 1904 and 1910, applying them to the calculus of variations and boundary value problems, such as those arising in physics. Hilbert's work on quadratic forms and symmetric integral operators emphasized the role of orthogonal expansions and eigenfunction series, where sequences of approximating functions converged pointwise or in the mean to minimizers of variational functionals, rather than uniformly across the domain. These considerations arose from the need to handle ill-posed problems in infinite dimensions, paving the way for topologies that prioritize integral or bilinear pairings over norm-based uniformity. Earlier informal notions of "weak" convergence appeared in the analysis of and orthogonal expansions during the late 19th and early 20th centuries. , in the 1880s, investigated the of representations for solutions to partial differential equations like the , often relying on or variational convergence to justify expansions in eigenfunctions, which avoided the pitfalls of failures observed in classical trigonometric series. extended this in the 1900s through his development of the Lebesgue integral, enabling rigorous treatments of mean-square for orthogonal series, where limits were understood in an averaged sense over the domain rather than everywhere, thus accommodating non-uniform behaviors in expansions. This progression from concrete analysis to abstraction was advanced by Maurice Fréchet in 1906 and in 1910. Fréchet's introduction of metric spaces and abstract sets of functions shifted focus from specific convergence types to general topological structures, allowing weaker limits that preserved distances in non-uniform ways. Riesz, in turn, unified these ideas by applying them to L^p spaces and integral equations, explicitly considering weak limits of sequences in abstract linear spaces to resolve convergence issues in orthogonal projections and operator theories. These developments marked a critical transition, embedding early convergence intuitions into the framework of infinite-dimensional vector spaces.

Key Milestones and Contributors

The formalization of in the was significantly advanced by Stefan Banach's work spanning 1920 to 1932, culminating in his seminal monograph Théorie des opérations linéaires. In this text, Banach introduced the concept of for sequences in \ell_p spaces and extended it to general normed linear spaces, defining a sequence \{x_n\} in a space E as weakly convergent to x \in E if \lim_{n \to \infty} f(x_n) = f(x) for every continuous linear functional f in the E^*. John von Neumann contributed crucially in 1932 through his book Mathematische Grundlagen der Quantenmechanik, where he developed weak topologies specifically in Hilbert spaces and provided the first explicit use of dual pairings to define them. Von Neumann's approach emphasized the weak operator topology on bounded operators, leveraging the inner product structure of Hilbert spaces to study convergence properties essential for . The Hahn-Banach theorem, independently established by Hans Hahn in 1927 and in 1932, proved pivotal for weak topologies by enabling the extension of linear functionals from subspaces to the entire space while preserving boundedness. This result ensures the separation of points by hyperplanes in the weak topology, underpinning the Hausdorff property and facilitating proofs of and . In the early , Stanisław Mazur advanced the theory by establishing key results on weak compactness in reflexive spaces, notably proving in 1933 that the closed unit ball of a reflexive is weakly compact. Mazur's work connected weak topological properties to reflexivity, showing that weakly convergent sequences in such spaces admit convex combinations that converge in norm. These milestones laid the groundwork for applications in , where weak topologies facilitate the study of bounded operators on Hilbert spaces.

Core Definitions

Weak Topology via Bilinear Pairings

In the context of topological vector spaces, the weak topology can be defined abstractly using a bilinear pairing between two vector spaces. Consider vector spaces X and Y over the complex numbers \mathbb{C}, equipped with a bilinear pairing \langle \cdot, \cdot \rangle: X \times Y \to \mathbb{C} that is linear in each argument separately and non-degenerate, meaning that if \langle x, y \rangle = 0 for all y \in Y, then x = 0, and similarly for Y. Such a setup forms a dual system or dual pair (X, Y). The weak topology on X, denoted \sigma(X, Y), is the initial topology induced by the family of seminorms p_y(x) = |\langle x, y \rangle| for all y \in Y. This is the coarsest topology on X that renders each map x \mapsto \langle x, y \rangle continuous from X to \mathbb{C} for every y \in Y. Equivalently, in the category of topological vector spaces, it is the with respect to this separating family of continuous linear functionals on X, distinguishing it from more general initial or final topologies that might not restrict to linear maps. The resulting space (X, \sigma(X, Y)) is Hausdorff and locally , with a base of neighborhoods at the consisting of sets of the form \{x \in X : |\langle x, y_i \rangle| < \epsilon_i \text{ for } i=1,\dots,n\}, where y_1, \dots, y_n \in Y and \epsilon_i > 0. A net (x_\lambda)_{\lambda \in \Lambda} in X converges to x \in X in this weak topology if and only if it converges with respect to the , that is, \langle x_\lambda, y \rangle \to \langle x, y \rangle for every y \in Y. This characterization emphasizes the topology's focus on in the "directions" defined by Y. When Y = X' is the continuous dual space of a topological vector space X, the pairing is the canonical evaluation map \langle x, f \rangle = f(x) for f \in X', and the resulting weak topology \sigma(X, X') is the standard weak topology on X. In this case, the seminorms are p_f(x) = |f(x)| for f \in X', yielding the coarsest topology making all continuous linear functionals continuous. This construction provides the general framework for weak topologies in , encompassing various dualities beyond normed spaces.

Weak Topology on Normed Spaces

In a normed linear space X, the weak topology, denoted \sigma(X, X'), is the coarsest topology that renders every continuous linear functional f \in X' continuous, where X' is the continuous of X. This topology equips X with the structure of a locally Hausdorff , with X' serving as its dual. The weak topology is generated by the family of seminorms p_f(x) = |f(x)| for all f \in X', which form a separating family on X. A net (x_\lambda) in X converges weakly to x \in X, written x_\lambda \rightharpoonup x, f(x_\lambda) \to f(x) for every f \in X'. x_\lambda \rightharpoonup x \quad \Leftrightarrow \quad f(x_\lambda) \to f(x) \quad \forall f \in X'. In the context of normed spaces, the weak \sigma(X, X') stands as the weakest Mackey topology, meaning it is the coarsest locally topology on X for which the continuous dual coincides precisely with X'; finer topologies, such as the Mackey topology itself, share this dual but are strictly stronger in general. Regarding boundedness, a B \subset X is weakly bounded if \sup_{x \in B} |f(x)| < \infty for every f \in X'. By the uniform boundedness principle (also known as the Banach-Steinhaus theorem), every weakly bounded set in a normed space is norm-bounded. This equivalence underscores the weak topology's utility in controlling boundedness without invoking the full norm structure, particularly in Banach spaces where reflexivity or separability may further refine properties.

Weak* Topology

In the context of a normed vector space X, the weak* topology on the dual space X' (also denoted X^*) is the initial topology induced by the family of evaluation maps \hat{x}: X' \to \mathbb{K} (where \mathbb{K} is the scalar field, typically \mathbb{R} or \mathbb{C}) defined by \hat{x}(f) = f(x) for each fixed x \in X. This topology, often denoted \sigma(X', X), is generated by the seminorms p_x(f) = |f(x)| for all x \in X, making it the coarsest topology on X' such that each \hat{x} is continuous when X is equipped with its norm topology. Equivalently, the weak* topology can be constructed via the bilinear pairing \langle f, x \rangle = f(x), where the subbasis of neighborhoods of the origin in X' consists of sets \{f \in X' : |\langle f, x_i \rangle| < \epsilon_i \text{ for } i=1,\dots,n\} for finite collections x_1, \dots, x_n \in X and \epsilon_i > 0. Convergence in the weak* topology is characterized pointwise on the predual: a net (f_\lambda)_{\lambda \in \Lambda} in X' converges weak* to f \in X' if and only if f_\lambda(x) \to f(x) for every x \in X. This form of aligns with the topology's role in duality theory, where it ensures that the dual pairing remains continuous without requiring stronger uniformity conditions imposed by the on X'. The weak* topology differs from the full weak topology on X', which is \sigma(X', X'') and generated by seminorms |g(f)| for g \in X'' (the bidual of X). Since the canonical j: X \hookrightarrow X'' identifies X as a of X'', the family of seminorms from X forms a of those from X'', rendering \sigma(X', X) coarser than \sigma(X', X''). The topologies coincide X is reflexive, i.e., X = X''. A significant advantage of the weak* topology is that it equips X' with a locally topological vector space structure regardless of whether X is complete, as the defining seminorms are continuous on X under its norm topology and the resulting space admits continuous addition and . This property underscores its utility in duality theory for normed spaces, where completeness of X is not presupposed.

Comparisons with Other Topologies

Strong Topology

In a normed space X equipped with a \|\cdot\|, the topology is the norm topology induced by the open balls B(x, r) = \{ y \in X : \|y - x\| < r \} for x \in X and r > 0, which endows X with the structure of a . This topology ensures that addition and are continuous operations, with the norm providing a d(x, y) = \|x - y\| that generates the open sets. Strong convergence in this topology is defined such that a net (or sequence) \{x_\alpha\} in X converges strongly to x \in X if and only if \|x_\alpha - x\| \to 0. The strong topology is finer than the weak topology on X, meaning every open set in the weak topology is open in the strong topology, but the converse does not hold. Consequently, every strongly convergent net is weakly convergent, but the reverse implication fails in infinite-dimensional spaces. The strong and weak topologies coincide precisely in finite-dimensional normed spaces, where all norms are equivalent and induce the same . In such cases, the norm topology aligns fully with the weak topology generated by the continuous linear functionals.

Norm-Induced Topology

The norm-induced topology on a normed linear space X with \|\cdot\| is the topology generated by the d(x, y) = \|x - y\|, which defines open balls B(x, r) = \{y \in X : \|y - x\| < r\} as a basis for the open sets. This structure ensures that the topology is metrizable by construction and Hausdorff, as the separates distinct points: for x \neq y, there exists \epsilon > 0 such that B(x, \epsilon) \cap B(y, \epsilon) = \emptyset. In the case where X is complete with respect to this , it forms a , endowing the space with completeness properties that facilitate the study of Cauchy sequences and fixed-point theorems. Unlike the weak topology, which is generally non-metrizable in infinite-dimensional spaces, the norm-induced topology remains metrizable regardless of , providing a stronger separation of points and sets. This metrizability allows for sequential characterizations of and , contrasting with the weak topology's reliance on nets or filters for in infinite dimensions. The norm topology arises from the uniform structure induced by the family of entourages \{(x, y) : \|x - y\| < \epsilon : \epsilon > 0\}, leading to of sequences or nets in the norm. In contrast, the weak topology corresponds to with respect to the X^*, resulting in a coarser uniform structure that does not metrize uniform convergence. In reflexive Banach spaces, the weak closure of a coincides with its norm closure, a consequence of the Hahn-Banach separation theorem applied in locally spaces.

Convergence and Properties

Weak Convergence

In a topological vector space equipped with the weak topology induced by its continuous dual X', a net (x_\lambda)_{\lambda \in \Lambda} in X converges weakly to x \in X, denoted x_\lambda \rightharpoonup x, f(x_\lambda) \to f(x) for every continuous linear functional f \in X'. This definition captures the coarsest topology making all elements of the dual continuous, ensuring that weak convergence tests continuity against the . In Banach spaces, which are complete normed spaces, this notion extends naturally, where the dual X' consists of all bounded linear functionals. For sequences in separable s, weak convergence coincides with the net definition, as the separability ensures that sequential criteria suffice for many topological properties, including arguments. The Eberlein–Šmulian theorem states that in a , a is relatively weakly compact if and only if every in the has a weakly convergent subsequence. This equivalence bridges general net-based with more tractable sequential in the weak topology. In Banach spaces, weak convergence preserves convex combinations through Mazur's lemma, which asserts that if x_n \rightharpoonup x, then there exists a sequence of convex combinations of the x_n that converges in norm to x. However, it does not generally preserve norms: \|x\| \leq \liminf \|x_n\| holds, with equality if and only if the space is reflexive. Reflexive spaces, where the canonical embedding into the bidual is surjective, thus ensure that weak limits retain norm information more robustly. A concrete criterion for weak convergence appears in L^p spaces for $1 < p < \infty, where the dual is L^q with $1/p + 1/q = 1: a sequence (f_n) converges weakly to f in L^p(\mu) if \int f_n g \, d\mu \to \int f g \, d\mu for all g \in L^q(\mu). \int f_n g \, d\mu \to \int f g \, d\mu \quad \forall g \in L^q(\mu). This integral test leverages the duality pairing, providing a measurable verification of weak limits in function spaces.

General Properties of Weak Topologies

The weak topology on a topological vector space X with respect to its continuous dual X^*, denoted \sigma(X, X^*), is Hausdorff provided that the bilinear pairing \langle \cdot, \cdot \rangle: X \times X^* \to \mathbb{K} (where \mathbb{K} is the scalar field) separates points in X, meaning that if \langle x, f \rangle = 0 for all f \in X^*, then x = 0. In the case of normed spaces, the guarantees this separation property, ensuring that the weak topology is Hausdorff. A key feature of weak topologies is their preservation of convexity: the weak closure of a convex subset of X is itself convex. This follows from the fact that the weak topology is compatible with the convex structure of the space, as finite linear combinations and limits in the weak sense maintain convexity. The further leverages this property, stating that in a , every compact convex subset is the closure (in the given topology, which may be the weak topology) of the convex hull of its extreme points. Weak topologies on infinite-dimensional normed spaces are generally not complete. Specifically, for a Banach space X, the weak topology \sigma(X, X^*) coincides with the norm topology if and only if X is finite-dimensional, and thus the weak topology is complete precisely in finite dimensions; in infinite dimensions, it fails to be complete due to the existence of Cauchy sequences that do not converge weakly to elements in X. Regarding boundedness, the uniform boundedness principle establishes a close link between weak and norm boundedness: a subset A \subseteq X is weakly bounded (meaning that for every f \in X^*, the set \{ \langle a, f \rangle : a \in A \} is bounded in \mathbb{K}) if and only if it is bounded in the norm topology. This equivalence holds because pointwise bounded families of continuous linear functionals on X are uniformly bounded in the dual norm.

Examples and Applications

Hilbert and Banach Spaces

In Hilbert spaces, the weak topology is the coarsest topology making all inner product functionals continuous, meaning a sequence \{x_n\} converges weakly to x if and only if \langle x_n, y \rangle \to \langle x, y \rangle for every y in the space. This definition leverages the , which identifies the dual space with the space itself via inner products. A classic illustration is the orthonormal basis \{e_n\} in a Hilbert space H, where e_n \rightharpoonup 0 weakly since \langle e_n, y \rangle \to 0 for all y \in H by , yet \|e_n\| = 1 prevents strong convergence. Reflexive Banach spaces, such as \ell_p for $1 < p < \infty, exhibit particularly well-behaved weak topologies. These spaces are reflexive, meaning the natural embedding into the bidual is surjective, and their unit balls are weakly compact by the Eberlein–Šmulian theorem. On bounded sets, the weak topology is metrizable when the dual space is separable, as a countable dense subset of the dual generates a metric d(x, y) = \sum_{i=1}^\infty 2^{-i} |\Lambda_i(x - y)| for \{\Lambda_i\} dense in the dual. James' theorem characterizes reflexivity by stating that a Banach space is reflexive if and only if every continuous linear functional attains its norm on the closed unit ball. In contrast, non-reflexive spaces like c_0, the space of sequences converging to zero under the sup norm, show stark differences in weak topology behavior. The closed unit ball of c_0 is not weakly compact, violating a key property of reflexive spaces and highlighting how weak and norm topologies diverge more significantly here. A concrete equation for weak convergence in the Hilbert space L_2(\Omega) is that f_n \rightharpoonup f if \int_\Omega f_n g \, d\mu \to \int_\Omega f g \, d\mu for all g \in L_2(\Omega), directly tying to the inner product structure.

Distribution Spaces

In the theory of distributions, the space \mathcal{D}'(\Omega) consists of all continuous linear functionals on the space of compactly supported smooth test functions \mathcal{C}_c^\infty(\Omega), and it is endowed with the weak* topology. This topology, also known as the weak topology on the dual space, ensures that a sequence of distributions \{T_n\} converges to a distribution T if and only if \langle T_n, \phi \rangle \to \langle T, \phi \rangle for every test function \phi \in \mathcal{C}_c^\infty(\Omega). This pointwise convergence on test functions provides a natural framework for handling generalized functions that arise in partial differential equations (PDEs), where classical solutions may not exist but distributional solutions do. Sobolev spaces H^s(\Omega), which are Hilbert spaces of functions with square-integrable weak derivatives up to order s, inherit weak convergence from the underlying L^2 structure, allowing sequences to converge in energy norms while preserving integrals against smooth functions. In variational methods for PDEs, weak convergence in H^s is crucial for proving the existence of minimizers of energy functionals, as it enables the application of compactness theorems like the to extract strongly convergent subsequences. This weak topology facilitates the passage to the limit in weak formulations, ensuring that solutions satisfy the equation in a distributional sense without requiring pointwise regularity. A classic example of weak convergence in distribution spaces is the approximation of the Dirac delta distribution \delta by sequences of smooth functions, such as Gaussian kernels \phi_\epsilon(x) = \frac{1}{\sqrt{2\pi\epsilon}} e^{-x^2/(2\epsilon)} as \epsilon \to 0^+, where \langle \phi_\epsilon, \psi \rangle \to \psi(0) = \langle \delta, \psi \rangle for any test function \psi. This convergence highlights how the weak* topology captures the "concentration" of mass at a point without strong convergence in norms like L^1 or L^2. Such approximations are fundamental in regularization techniques for singular sources in PDEs. The weak topology on distribution spaces is indispensable for defining weak solutions to elliptic PDEs, such as the Dirichlet problem for the Laplace equation -\Delta u = f on a bounded domain \Omega with boundary data g, where solutions in H^1(\Omega) satisfy the equation in the distributional sense \int_\Omega \nabla u \cdot \nabla v \, dx = \int_\Omega f v \, dx for all test functions v \in \mathcal{C}_c^\infty(\Omega). This approach, relying on weak convergence to establish existence via the Lax-Milgram theorem, extends classical theory to irregular data and non-smooth boundaries, forming the basis for modern variational analysis in PDEs.

Algebraic Dual Spaces

The algebraic dual of a vector space X over a field \mathbb{K} (typically \mathbb{R} or \mathbb{C}), denoted X^*, consists of all linear functionals on X, without requiring continuity with respect to any topology on X. This contrasts with the continuous dual X' used in normed or topological vector spaces, where only continuous linear functionals are considered. The weak topology \sigma(X, X^*) on X is the initial topology generated by the evaluation maps x \mapsto f(x) for all f \in X^*, making it the coarsest topology under which every functional in X^* is continuous. This topology is defined by the subbasis of open sets consisting of sets of the form \{x \in X : |f(x) - f(x_0)| < \epsilon\} for f \in X^*, x_0 \in X, and \epsilon > 0. Equivalently, it is the locally convex topology induced by the seminorms p_f(x) = |f(x)| for each f \in X^*. Unlike the weak topology \sigma(X, X') from the continuous dual in normed spaces, \sigma(X, X^*) incorporates a larger family of seminorms and is thus finer, as the additional functionals refine the open sets. The \sigma(X, X^*) separates points on X: for any distinct x, y \in X, there exists f \in X^* such that f(x) \neq f(y). Consequently, it is Hausdorff when \mathbb{K} is equipped with its standard Hausdorff topology. This topology is particularly useful in settings without a or pre-existing topology on X, such as purely algebraic spaces or inductive limits, where it provides a natural locally convex structure compatible with the full linear structure. In finite-dimensional spaces, say \dim X = n < \infty, the algebraic dual X^* is isomorphic to X as vector spaces, and \sigma(X, X^*) coincides with the standard Euclidean topology on X, which is the unique Hausdorff topology up to equivalence. For instance, on \mathbb{R}^n, the seminorms from X^* generate the usual norm topology, rendering all singletons closed but not open unless n=0. This equivalence highlights how \sigma(X, X^*) aligns with classical topologies in low dimensions while diverging in infinite dimensions, where it induces a product-like structure with respect to a Hamel basis. The topology \sigma(X, X^*) is coarser than the full pointwise convergence topology in certain extended senses (e.g., uniform on finite sets) but remains finer than \sigma(X, X') in normed settings, emphasizing its role in algebraic rather than analytic contexts.

Extensions and Advanced Topics

Operator Topologies

In the context of , the space B(H) of bounded linear operators on a H is equipped with several natural topologies that are coarser than the operator norm topology. The weak operator topology (WOT) on B(H) is the coarsest among these, defined by convergence T_\lambda \to T if and only if \langle T_\lambda x, y \rangle \to \langle T x, y \rangle for all x, y \in H, where \langle \cdot, \cdot \rangle denotes the inner product on H. This topology arises from the seminorms p_{x,y}(T) = |\langle T x, y \rangle| for x, y \in H, making B(H) a locally . The strong operator topology (SOT) on B(H) is defined by convergence T_\lambda \to T if and only if T_\lambda x \to T x in the norm of H for every x \in H. Equivalently, it is generated by the seminorms q_x(T) = \| T x \| for x \in H. The SOT is strictly finer than the WOT but still coarser than the topology, as uniform boundedness on bounded sets distinguishes the norm topology from the SOT. In particular, a net converges in the SOT if it does so pointwise in the strong topology of H, but WOT convergence requires only scalar convergence via matrix elements. These topologies play a crucial role in the study of operator algebras, such as algebras, which are defined as weakly closed *-subalgebras of B(H). A classic example illustrating the distinction between WOT and SOT convergence is provided by the powers of the unilateral S on the \ell^2(\mathbb{N}), where (S e_n) = e_{n+1} for the standard orthonormal basis \{e_n\}. The sequence S^n converges to the zero in the WOT, since \langle S^n x, y \rangle \to 0 for all x, y \in \ell^2(\mathbb{N}) as the supports shift to infinity, but it does not converge in the SOT because \| S^n e_1 \| = 1 for all n. This example highlights how WOT allows for "pointwise" dissipation that SOT detects as persistent action on individual vectors.

Weak Topologies in Topological Vector Spaces

In a E, the weak topology \sigma(E, E') is the generated by the family of all continuous linear functionals E' \to \mathbb{K}, where E' denotes the continuous dual of E and \mathbb{K} is the underlying field. This makes \sigma(E, E') the coarsest on E such that every element of E' remains continuous. For locally convex topological vector spaces, the Mackey-Arens theorem provides a fundamental characterization of topologies compatible with the duality (E, E'): every such topology lies between the weak topology \sigma(E, E') and the Mackey topology \tau(E, E'), where the latter is defined by uniform convergence on the absolutely , weakly compact subsets of E'. This theorem implies the existence of a finest locally convex topology compatible with the duality, ensuring that the weak topology serves as the minimal such structure while preserving duality properties. Bornological spaces, which are locally convex topological vector spaces in which every convex bornivore is a neighborhood of the origin, exhibit compatibility between the weak topology and their inductive limit representations. Specifically, a bornological space admits a description as an inductive limit of normed spaces where the weak topology aligns with the bornology, facilitating the study of bounded sets and continuity in limit constructions. Post-1960s advancements have generalized weak topologies to non-Hausdorff topological vector spaces and ind-completed spaces, such as countable inductive limits of Fréchet spaces (LF-spaces), where the induced by the continuous enables analysis of without requiring separation axioms. In LF-spaces, the weak topology supports sequential by ensuring that sequences Cauchy in the weak sense converge within the inductive structure, provided each constituent Fréchet space is sequentially complete.

References

  1. [1]
    [PDF] An introduction to some aspects of functional analysis, 6: Weak and ...
    The weak topology on V is automatically weaker than the original topology on V , in the sense that every open set in V with respect to the weak topology is an ...Missing: mathematics | Show results with:mathematics
  2. [2]
    [PDF] E.6 The Weak and Weak* Topologies on a Normed Linear Space
    The “weak topology” corresponding to such a class is the weakest (smallest) topology such that each map fα is continuous. Another example of such a weak ...Missing: mathematics | Show results with:mathematics
  3. [3]
    [PDF] 4. Dual spaces and weak topologies Recall that if X is a Banach ...
    The weak-∗ topology Tw* on X∗ is defined as the weak topology generated by X, viewed as a subset of X∗∗. Put differently, Tw* is the weakest topology that turns ...Missing: mathematics | Show results with:mathematics
  4. [4]
    [PDF] On the origin and early history of functional analysis - DiVA portal
    In this report we will study the origins and history of functional analysis up until 1918. We begin by studying ordinary and partial differential equations ...
  5. [5]
    [PDF] The Establishment of Functional Analysis
    This article surveys the evolution of functional analysis, from its origins to its establish- ment as an independent discipline around 1933.
  6. [6]
    None
    Below is a merged summary of the weak topology history from Pietsch’s "History of Banach Spaces and Linear Operators," consolidating all information from the provided segments into a comprehensive response. To retain maximum detail, I will use a table in CSV format for clarity and density, followed by a narrative summary that integrates key points and additional context not easily captured in the table. The table will focus on specific contributions, page references, key facts, and quotes, while the narrative will provide an overarching synthesis.
  7. [7]
    [PDF] Duality Theory I - KSU Math
    The two topologies constructed above are called the weak topologies associated with the dual pairing. Remarks 1-2. Use the notations as above.
  8. [8]
    Dual Pairs and the Weak Topology | 16 - Taylor & Francis eBooks
    ... dual pair under the natural bilinear form: < u, x ' > = x ' ( u ) ( for all ... weak topology on X, denoted by σ(X,Y), is defined by the family { p y ...
  9. [9]
    [PDF] FUNCTIONAL ANALYSIS | Second Edition Walter Rudin
    Rudin, Walter, (date). Functional analysis/Walter Rudin.-2nd ed. p. em. -(international series in pure and applied mathematics). Includes bibliographical ...
  10. [10]
    [PDF] FUNCTIONAL ANALYSIS - ETH Zürich
    Jun 8, 2017 · The subject of Chapter 3 are the weak topology on a Banach space X and the weak* topology on its dual space X∗. With these topologies X and ...
  11. [11]
    weak-* topology - PlanetMath
    Mar 22, 2013 · The weak-* topology on X* is defined by seminorms {px|x∈X}, where X* is the continuous dual of X, and is denoted by σ(X*,X).
  12. [12]
    245B, Notes 11: The strong and weak topologies - Terry Tao
    Feb 21, 2009 · Certainly many results in functional analysis involving the continuous dual space will break down if the algebraic dual space is used instead, ...
  13. [13]
    [PDF] Chapter 13: Metric, Normed, and Topological Spaces - UC Davis Math
    A special type of metric space that is particularly important in analysis is a normed space, which is a vector space whose metric is derived from a norm. On the ...Missing: metrizable | Show results with:metrizable
  14. [14]
    [PDF] FUNCTIONAL ANALYSIS1 Douglas N. Arnold2 References
    ... weak* topology and is a weaker topology than the weak topology. If X is reflexive, the weak and weak* topologies coincide. Examples of weak and weak ...
  15. [15]
    [PDF] Functional Analysis Princeton University MAT520 Lecture Notes
    Aug 18, 2023 · ... weak topology on X∗ and the weak-star topology on X∗ coincide. Proof. The weak topology on X∗ is characterized by the initial topology ...
  16. [16]
    [PDF] “Banach spaces and topology (I) (For Encyclopedia on General ...
    Jun 26, 2001 · The topology of the norm metric is called the norm topology, and the normed space is always assumed to have the norm topology unless other ...
  17. [17]
    [PDF] GEOMETRIES OF TOPOLOGICAL GROUPS Contents 1. Banach ...
    The above results indicate that the uniform structure of a Banach space is more rigid than the coarse structure. However, once we pass to the underlying ...
  18. [18]
    [PDF] The weak topology of locally convex spaces and the ... - Jordan Bell
    Apr 3, 2014 · 22Walter Rudin, Functional Analysis, second ed., p. 94, Theorem 4.3. 23Robert E. Megginson, An Introduction to Banach Space Theory, p. 248, ...
  19. [19]
    [PDF] Weak convergence in Banach Spaces - Joel H. Shapiro
    Jan 22, 2022 · (b) If dim X < ∞ then every weakly con- vergent seq. in X is norm-convergent. Question Is weak convergence always the same as norm convergence?Missing: 1929 | Show results with:1929
  20. [20]
    [PDF] Banach Spaces - UC Davis Math
    A third type of convergence of operators, weak convergence, may be defined using the notion of weak convergence in a Banach space, given in Definition 5.59.
  21. [21]
    [PDF] Weak Convergence in Banach Spaces - Joel H. Shapiro
    Feb 28, 2022 · The discussion revolves around Issai Schur's remark- able discovery: In the classical sequence space `1, every weakly convergent sequence is ...Missing: l_p | Show results with:l_p
  22. [22]
    The Eberlein-Smulian Theorem
    Theorem (Eberlein-Smulian). A subset of a Banach space is relatively weakly compact if and only if it is relatively weakly sequentially compact. In particular, ...
  23. [23]
    Weak convergence implying norm convergence - MathOverflow
    Feb 23, 2012 · When a sequence converges weakly then there is another sequence made of (finite) convex combination which converges in norm (to the same element).Weak convergence (probability theory) and weak ... - MathOverflowAre weak and strong convergence of sequences not equivalent?More results from mathoverflow.net
  24. [24]
    [PDF] Chapter 8. The L Spaces: Duality and Weak Convergence
    Feb 25, 2023 · Spaces: Duality and Weak Convergence. Section 8.1. The Riesz Representation for the Dual of L p. , 1 ≤ p < ∞. Note. In this section, we ...
  25. [25]
    [PDF] an introduction to functional analysis - UChicago Math
    Aug 7, 2010 · Weak limits are unique, and weakly convergent sequences are bounded. Proof. If xn * x and xn * y then it follows that f(x) = f(y) for every f ∈ ...
  26. [26]
    [PDF] Extreme points and the Krein–Milman theorem - Caltech
    Definition An extreme point of a convex set, A, is a point x ∈ A, with the property that if x = θy + z with y, z ∈ A and θ ∈ [0, 1], then y = x and/or z = x ...
  27. [27]
    [PDF] Banach and Hilbert space analysis - DPMMS
    Thus a Hilbert space is a special case of a Banach space, however the presence of ... known as the weak topology. With respect to the weak topology, the elements ...<|separator|>
  28. [28]
    [PDF] Section 3.8. Properties of Orthonormal Systems
    Apr 21, 2019 · (x, xn) = 0 and so all orthonormal sequences weakly converge to 0. Definition 3.8.1. An orthonormal sequence {xn} in a Hilbert space H is ...
  29. [29]
    [PDF] Weak Topology. A weak open set around x ∈ X is given by N(x
    Since X ⊂X∗∗ one is weaker than the other. The weak topology on X∗ can come from considering either X or X∗∗. One hardly ever chooses X∗∗. In many examples ...Missing: mathematics | Show results with:mathematics<|control11|><|separator|>
  30. [30]
    [PDF] arXiv:2102.06452v1 [math.FA] 12 Feb 2021
    Feb 12, 2021 · The famous James theorem states that a Banach space E is reflexive if and only if every bounded linear functional on E attains its norm.
  31. [31]
    [PDF] E.7 Alaoglu's Theorem
    Let X be a separable normed space, and fix K ⊆ X∗. If K is weak*-compact, then the weak* topology of X∗ restricted to K is metrizable. Proof. Let {xn} ...Missing: norm | Show results with:norm
  32. [32]
    Sobolev Spaces - Robert A. Adams, John J. F. Fournier
    Jun 26, 2003 · Sobolev Spaces presents an introduction to the theory of Sobolev Spaces and other related spaces of function, also to the imbedding characteristics of these ...
  33. [33]
    The Analysis of Linear Partial Differential Operators I - Google Books
    May 1, 1983 · The presentation then pro ceeded directly to the most general results available on partial differ ential operators.
  34. [34]
    Partial Differential Equations - Lawrence C. Evans - Google Books
    It offers a comprehensive survey of modern techniques in the theoretical study of PDE with particular emphasis on nonlinear equations.
  35. [35]
    [PDF] Topological Vector Spaces
    Jan 2, 2015 · ... algebraic dual of X. X0 is a vector subspace of X∗ and is ... usually called the finest locally convex topology on the given vector space.
  36. [36]
    [PDF] II.1 General weak topologies and duality
    (1) The space X is a LCS if it is equipped by the topology σ(X, M). (2) The topology σ(X, M) is Hausdorff if and only if M separates points of X, i.e., if and ...Missing: properties | Show results with:properties
  37. [37]
    [PDF] norm, strong, and weak operator topologies on b(h)
    In this lecture we will consider 3 topologies on B(H), the space of bounded linear operators on a Hilbert space H. In sections 1 and 2, we shall be reminded ...
  38. [38]
    [PDF] Properties of operator systems, corresponding to channels - arXiv
    May 26, 2020 · Topology on B (H) generated by seminorms ·x,y : B ↦→ |〈x, By〉| is called weak operator topology and is denoted wo. The space B (H) is dual to ...
  39. [39]
    [PDF] Notes on operator algebras
    Mar 17, 2014 · Let H be a Hilbert space. On B(H) we define the following locally convex topologies: • The weak operator topology (WOT) is defined by the family ...
  40. [40]
    [PDF] Lectures on C∗-algebras - arXiv
    Oct 11, 2021 · (i) The sequence {Sn} weakly converges to zero, but it is not strongly convergent ... the weak operator topology. For given x, y ∈ H, let Dx,y be ...
  41. [41]
    [PDF] 1 Lecture 14 - Notes for Functional Analysis
    Notes for Functional Analysis. Wang Zuoqin. (typed by Xiyu Zhai). Oct 27, 2015. 1 Lecture 14. 1.1 The weak topology defined by maps. Recall: • Let F1,F2 be ...Missing: normed textbook
  42. [42]
    A locally quasi-convex abelian group without a Mackey group topology
    Feb 28, 2018 · The Mackey–Arens theorem suggests the following general question posed in [4]: Is every locally quasi-convex abelian group a pre-Mackey group?<|separator|>
  43. [43]
    Lecture Notes in Mathematics
    Corollary 3: A quotient of a bornological space is bornological. Theorem 4.2. Any Bornological space (E,j) is inductive limit of semi-normed spaces. If E is ...Missing: source | Show results with:source
  44. [44]
    [PDF] Countable inductive limits - Open UCT - University of Cape Town
    This theref'ore shows that the inductive limit of' a sequence of' spaces with the wealc topologies need not yield a weak topology. University of Cape Town ...