Fact-checked by Grok 2 weeks ago

Filters in topology

In , a on a set X is a non-empty family \mathcal{F} of subsets of X such that the empty set is not in \mathcal{F}, \mathcal{F} is closed under finite intersections, and if A \in \mathcal{F} and A \subseteq B \subseteq X, then B \in \mathcal{F}. This structure generalizes the concept of "eventual" or "large" sets, analogous to the tails of sequences, and serves as a foundational tool for defining , , and in arbitrary topological spaces where sequences alone are insufficient. Introduced by in 1937, filters were developed to extend limit theories from metric spaces to general topologies, replacing sequences with a more flexible notion that captures adherence and neighborhood behaviors without countable indexing. Filters encompass principal filters generated by fixed subsets, Fréchet filters of cofinite sets on infinite spaces, and ultrafilters, which are maximal filters to which every proper filter can be extended and pivotal for non-constructive proofs like stating that arbitrary products of compact spaces are compact. In practice, the neighborhood filter at a point x in a (X, \tau) consists of all subsets containing an open neighborhood of x, enabling precise characterizations of topological properties: a filter \mathcal{F} converges to x if every member of the neighborhood filter at x belongs to \mathcal{F}. This framework, further refined by in their 1940 treatise on , underpins modern abstract approaches to uniform structures, initial topologies, and categorical limits, influencing areas from to .

Motivation

Historical Development

The development of filters in topology emerged as part of broader efforts to generalize notions of beyond sequences in arbitrary topological spaces. In 1922, and H. L. Smith introduced the concept of nets through their Moore-Smith , which extended sequential limits to functions on directed sets, addressing limitations in non-first-countable spaces. This framework laid groundwork for subsequent generalizations, leading to the introduction of filters by in 1937. Cartan defined filters as a tool to replace sequences in studying within general topological spaces, presenting the foundational in his "Théorie des filtres." His approach emphasized filters' role in capturing limit points more flexibly. In the 1940s, the Bourbaki group further developed and popularized filters, integrating them into their axiomatic treatise Topologie Générale (first chapters published in 1940). Bourbaki's rigorous, structure-based methodology unified filters with nets, establishing them as a cornerstone of modern and facilitating convergence theory in abstract settings. This synthesis highlighted filters' equivalence to nets in defining limits, bridging earlier ideas and influencing subsequent topological advancements.

Role in Convergence and Limits

In topological spaces that are not first-countable, sequences often fail to fully characterize , , or , as they rely on countable index sets that cannot capture the complexity of in spaces with uncountable local bases. For instance, in non-first-countable spaces such as uncountable spaces or certain topologies, convergent sequences may only detect isolated points or eventually constant behaviors, missing broader adherence properties that require approaching a from uncountably many directions. Filters address these limitations by generalizing the notion of directed sets beyond countable sequences, providing a uniform framework for defining limits in arbitrary topological spaces through the adherence of filter bases to neighborhoods. Introduced by in 1937 to extend convergence theory beyond sequences, filters allow for the description of limits using arbitrary families of sets that refine toward smaller and smaller neighborhoods, without dependence on a countable ordering. A key advantage of filters is their ability to describe adherence and accumulation points in spaces where countable approximations are insufficient, enabling the uniform treatment of topological properties like and Hausdorff separation across all spaces. This avoids the "narrow view" of sequences, which are "too short and too thin" to probe uncountable structures effectively. For example, consider the on \{0,1\}^I where I is uncountable; here, sequences—corresponding to countable supports—cannot converge to points with uncountable support, as they remain confined to finite or countable variations, whereas filter can capture such limits by refining through bases that intersect all relevant neighborhoods.

Fundamentals

Prefilters and Filters

A prefilter on a set X (also called a filter base) is a non-empty family \mathcal{F} of non-empty subsets of X that is closed under finite intersections. Formally, for all A, B \in \mathcal{F}, A \cap B \in \mathcal{F}. The empty set is not an element of \mathcal{F}. A filter on X is a prefilter that is also closed under supersets: if A \in \mathcal{F} and A \subseteq B \subseteq X, then B \in \mathcal{F}. Every filter contains X (take any A \in \mathcal{F} and superset to X) and excludes the , ensuring it is proper and non-trivial for describing limits in topological spaces. Some prefilters, such as the collection of all cofinite subsets, are already filters since they are upward closed. The Fréchet filter on the natural numbers \mathbb{N} is the filter consisting of all cofinite subsets of \mathbb{N}, i.e., all A \subseteq \mathbb{N} such that \mathbb{N} \setminus A is finite. Given a non-empty \mathcal{G} of subsets of X, the filter generated by \mathcal{G} is the smallest filter containing \mathcal{G}, formed by taking all supersets of finite intersections of elements from \mathcal{G}. A principal filter generated by a single set A \subseteq X (with A \neq \emptyset) is the collection \{B \subseteq X \mid A \subseteq B\}, which is upward closed and thus a filter. For example, the cofinite filter on an infinite set is non-principal, unlike principal filters fixed by a specific subset.

Basic Examples and Ultrafilters

A fundamental example of a filter in a topological space is the neighborhood filter at a point x, denoted \mathcal{N}(x), which consists of all neighborhoods of x in the space. This filter captures the local structure around x, where a set U belongs to \mathcal{N}(x) if x \in U and U contains an open set containing x. In metric spaces like \mathbb{R}, \mathcal{N}(x) includes all open intervals (x - \epsilon, x + \epsilon) for \epsilon > 0. Another standard example is the Fréchet filter on the natural numbers \mathbb{N}, defined as the collection of all cofinite subsets of \mathbb{N}, i.e., sets whose complements are finite. This filter is free, meaning the intersection of all its members is empty, and it is not principal since no single point fixes it. On the real line \mathbb{R}, the filter of neighborhoods of infinity consists of all subsets U \subseteq \mathbb{R} that contain some tail interval (r, \infty) for r \in \mathbb{R}. This filter models divergence to +\infty and is generated by the base \{(r, \infty) \mid r \in \mathbb{R}\}. Ultrafilters extend the notion of filters to maximal elements under inclusion. An ultrafilter on a set X is a filter \mathcal{U} that cannot be properly extended to a larger filter, equivalently, for every subset A \subseteq X, exactly one of A or X \setminus A belongs to \mathcal{U}. By , every filter on X is contained in some ultrafilter. Ultrafilters are classified as principal or (non-principal). A principal ultrafilter is fixed by a point p \in X, consisting of all subsets of X containing p, i.e., \hat{p} = \{A \subseteq X \mid p \in A\}. In contrast, a ultrafilter has empty total intersection \bigcap \mathcal{U} = \emptyset and contains no finite sets, extending the Fréchet filter on infinite sets like \mathbb{N}. The neighborhood \mathcal{N}(x) in a is an ultrafilter if and only if x is an . A key application of ultrafilters on \mathbb{N} arises in the Stone-Čech compactification \beta \mathbb{N}, where the points of \beta \mathbb{N} are identified with the ultrafilters on \mathbb{N}, and \mathbb{N} embeds densely via principal ultrafilters. Each ultrafilter \mathcal{U} on \mathbb{N} converges to the unique point in \beta \mathbb{N} it represents, with the basis topology ensuring that every such ultrafilter defines a convergent structure in this compactification.

Kernels and Equivalence

The kernel of a prefilter \mathcal{F} on a set X, denoted \ker \mathcal{F}, is defined as the intersection of all sets belonging to \mathcal{F}, that is, \ker \mathcal{F} = \bigcap_{A \in \mathcal{F}} A. This set represents the collection of elements common to every member of the prefilter. (Bourbaki, 1966) The kernel \ker \mathcal{F} possesses key properties in the context of convergence and adherence. It serves as the smallest set that is contained within every element of \mathcal{F}, thereby capturing the "fixed" portion of the space under the prefilter's structure. Moreover, a prefilter \mathcal{F} is termed free if \ker \mathcal{F} = \emptyset, meaning no point belongs to all sets in \mathcal{F}; otherwise, it is fixed. For instance, principal ultrafilters have singleton kernels. (Bourbaki, 1966) Equivalence relations on prefilters provide a means to identify those that induce identical limiting behaviors. Two prefilters \mathcal{F} and \mathcal{G} on X are , denoted \mathcal{F} \sim \mathcal{G}, if they generate the same and possess the same . This equivalence ensures that \mathcal{F} and \mathcal{G} yield the same convergent structures in topological settings. A stronger form of equivalence holds if \mathcal{F} = \mathcal{G} as collections of sets (ultra-strong equivalence). Weaker variants include having the same "tails," where for every A \in \mathcal{F}, there exists B \in \mathcal{G} such that B \subseteq A, and vice versa, implying the generated filters coincide. (Bourbaki, 1966)

Relations and Orderings

Finer and Coarser Structures

In the of filters and prefilters on a set X, these structures are partially ordered by . A prefilter \mathcal{F} is coarser than a prefilter \mathcal{G}, denoted \mathcal{F} \leq \mathcal{G}, if \mathcal{F} \subseteq \mathcal{G}; equivalently, every set in \mathcal{F} belongs to \mathcal{G}, making \mathcal{G} finer than \mathcal{F}. This reflects the refinement of the collections: finer prefilters contain all sets of coarser ones plus additional, typically smaller sets (due to upward closure). The relation is reflexive and transitive, forming a partial order on the set of all prefilters on X. The order is strict if \mathcal{F} < \mathcal{G}, meaning \mathcal{F} \leq \mathcal{G} but \mathcal{G} \not\leq \mathcal{F}; that is, \mathcal{F} \subsetneq \mathcal{G}. This strict inclusion captures proper refinements where the finer prefilter strictly extends the coarser one by including more sets. For filters (proper prefilters satisfying the additional axioms), the order restricts naturally, preserving the partial order structure. A representative example illustrates this ordering: on the natural numbers \mathbb{N}, the trivial filter \{ \mathbb{N} \} is coarser than the Fréchet filter \mathcal{F} of all cofinite subsets, since \{ \mathbb{N} \} \subseteq \mathcal{F} (as \mathbb{N} is cofinite), but \mathcal{F} includes additional sets such as \mathbb{N} \setminus \{1\}, making \mathcal{F} strictly finer. This relation holds more broadly, where principal filters generated by finite sets are coarser than the Fréchet filter on infinite sets. The partial order exhibits monotone properties that facilitate constructions and comparisons. The intersection of any family of prefilters finer than a given prefilter \mathcal{F} (i.e., each containing \mathcal{F}) yields a prefilter that is again finer than \mathcal{F}, as the intersection still contains \mathcal{F} and satisfies the prefilter axioms. Similarly, for filter generation from bases or prefilters, the order is preserved: if \mathcal{B}_1 \leq \mathcal{B}_2 (where \mathcal{B}_1, \mathcal{B}_2 are bases with \mathcal{B}_1 \subseteq \mathcal{B}_2), then the filter generated by \mathcal{B}_1 is coarser than or equal to the filter generated by \mathcal{B}_2, since the latter includes all supersets of a larger base, resulting in a finer (larger) collection. These properties ensure the order is compatible with standard operations on prefilters.

Subordination and Meshing

In the theory of prefilters on a set X, subordination provides a partial order that generalizes the notions of finer and coarser structures. A prefilter \mathcal{F} is subordinate to a prefilter \mathcal{G}, denoted \mathcal{F} \precsim \mathcal{G}, if for every set A \in \mathcal{F} there exists a set B \in \mathcal{G} such that B \subseteq A. This relation is reflexive and transitive, forming a poset on the collection of all prefilters on X, where \mathcal{F} \precsim \mathcal{G} indicates that \mathcal{G} refines \mathcal{F} by containing "smaller" sets relative to those in \mathcal{F}. The mesh of two prefilters \mathcal{F} and \mathcal{G} on X, denoted \mathcal{F} \wedge \mathcal{G}, is the prefilter generated by the family \{A \cap B \mid A \in \mathcal{F}, B \in \mathcal{G}\}. This construction ensures closure under finite intersections and upward closure under supersets, provided the generating family is nonempty and excludes the empty set. The mesh operation satisfies key properties in the subordination order: \mathcal{F} \wedge \mathcal{G} is the greatest lower bound (infimum) of \mathcal{F} and \mathcal{G}, meaning \mathcal{F} \wedge \mathcal{G} \precsim \mathcal{F}, \mathcal{F} \wedge \mathcal{G} \precsim \mathcal{G}, and for any prefilter \mathcal{H} with \mathcal{H} \precsim \mathcal{F} and \mathcal{H} \precsim \mathcal{G}, it follows that \mathcal{H} \precsim \mathcal{F} \wedge \mathcal{G}. Additionally, if \mathcal{F} \precsim \mathcal{G}, then \mathcal{F} \wedge \mathcal{G} = \mathcal{G}, reflecting that the coarser prefilter absorbs the mesh. A representative example arises in topological spaces, where the neighborhood prefilter \mathcal{N}_x at a point x \in X is subordinate to the principal prefilter \mathcal{P}_x consisting of all subsets containing x. For any A \in \mathcal{N}_x (so x \in A), there exists B = \{x\} \in \mathcal{P}_x with B \subseteq A. This subordination highlights how local structure relates to pointwise adherence, with the mesh \mathcal{N}_x \wedge \mathcal{P}_x = \mathcal{P}_x preserving the coarser principal information.

Trace Operations

In topology, the trace operation allows a prefilter on a set X to be restricted to a subset A \subseteq X, yielding a prefilter on A that captures the "large" sets relative to A. Specifically, for a prefilter \mathcal{F} on X, the trace \mathcal{F}|_A is defined as the collection \{ B \subseteq A \mid \exists C \in \mathcal{F} \text{ such that } C \cap A = B \}. This construction ensures that \mathcal{F}|_A consists precisely of the intersections of \mathcal{F}-sets with A, forming a for the induced prefilter on A. The trace operation preserves the subordination relation between prefilters. If \mathcal{F} \sqsubseteq \mathcal{G} (meaning \mathcal{F} is finer than or equal to \mathcal{G}, so \mathcal{G} \subseteq \mathcal{F}), then \mathcal{F}|_A \sqsubseteq \mathcal{G}|_A, as every intersection defining a set in \mathcal{G}|_A arises from a coarser collection and thus appears in the finer trace. Moreover, if \mathcal{F} is a filter (proper prefilter containing X) and A \in \mathcal{F}, the filter generated by the base \mathcal{F}|_A is a proper filter on A; the condition A \in \mathcal{F} guarantees A \in \mathcal{F}|_A (via X \cap A = A) and excludes the empty set, since any C \in \mathcal{F} must intersect A nontrivially to avoid contradicting the properness of \mathcal{F}. A representative example arises with the Fréchet filter \mathcal{F} on the natural numbers \mathbb{N}, consisting of all cofinite subsets. For the subset A of even numbers, the trace \mathcal{F}|_A consists of all subsets of A with finite complement in A (i.e., cofinite in A), as any cofinite C \subseteq \mathbb{N} intersects A in a cofinite portion of A, and this base generates the Fréchet filter on A. Traces play a key role in relative topology: the subspace topology on A \subseteq X is generated by taking traces of open sets from the topology on X, and analogously, the neighborhood filter of a point x \in A in the relative topology coincides with the trace on A of the neighborhood filter of x in X.

Constructions

Images and Preimages

In topology, given a function f: X \to Y and a prefilter \mathcal{F} on X, the direct image (or image) of \mathcal{F} under f, denoted f(\mathcal{F}), is the collection \{f(A) \mid A \in \mathcal{F}, f(A) \neq \emptyset\}, which forms a prefilter on Y. This construction preserves the finite intersection property: for any A, B \in \mathcal{F}, f(A \cap B) = f(A) \cap f(B) \in f(\mathcal{F}). However, f(\mathcal{F}) is generally not closed under supersets, as a superset of some f(A) need not be expressible as f(B) for any B \in \mathcal{F}; thus, the filter generated by f(\mathcal{F}) is required to obtain a full filter on Y. Dually, for a prefilter \mathcal{G} on Y, the inverse image (or preimage) of \mathcal{G} under f, denoted f^{-1}(\mathcal{G}), is the collection \{f^{-1}(B) \mid B \in \mathcal{G}\}, which forms a prefilter on X provided that B \cap f(X) \neq \emptyset for every B \in \mathcal{G}; this ensures no empty sets appear in the collection. If f is surjective, then f(X) = Y, so the condition holds automatically since sets in \mathcal{G} are nonempty. Like the direct image, f^{-1}(\mathcal{G}) preserves finite intersections: f^{-1}(B_1 \cap B_2) = f^{-1}(B_1) \cap f^{-1}(B_2). The collection f^{-1}(\mathcal{G}) is also directed by inclusion when \mathcal{G} is. A representative example arises with neighborhood filters in topological spaces. Let \mathcal{N}_x be the neighborhood prefilter of a point x \in X in a topological space (X, \tau), consisting of all neighborhoods of x. For a continuous function f: (X, \tau) \to (Y, \sigma), the direct image f(\mathcal{N}_x) consists of sets f(U) where U is a neighborhood of x; since f is continuous, each such f(U) is a neighborhood of f(x) in Y, so f(\mathcal{N}_x) generates the neighborhood filter \mathcal{N}_{f(x)} at f(x). This illustrates how images under continuous mappings relate local structures at points.

Preservation under Mappings

In filter theory, subordination relations are preserved under the formation of images and preimages of filters, with certain conditions ensuring the behavior aligns with the original structure. Specifically, consider a mapping f: X \to Y and filters \mathcal{F} and \mathcal{G} on X such that \mathcal{F} \sqsubseteq \mathcal{G}, meaning \mathcal{F} is coarser than \mathcal{G} (every set in \mathcal{F} contains a set from \mathcal{G}). The image filter f(\mathcal{F}) on Y, generated by the base \{f(A) \mid A \in \mathcal{F}\}, satisfies f(\mathcal{F}) \sqsubseteq f(\mathcal{G}). This preservation holds because every base set f(A) in f(\mathcal{F}) contains f(B) for some B \in \mathcal{G}, as A \supset B by the subordination \mathcal{F} \sqsubseteq \mathcal{G}, ensuring that sets in f(\mathcal{F}) adhere to the coarser relation relative to f(\mathcal{G}). A detailed proof sketch proceeds as follows. The filter f(\mathcal{F}) consists of all subsets of Y that contain some f(A) with A \in \mathcal{F}. For any such base set V \supset f(A) in f(\mathcal{F}), since \mathcal{F} \sqsubseteq \mathcal{G}, there exists B \in \mathcal{G} with B \subset A, implying f(B) \subset f(A) \subset V. Thus, V contains a base set f(B) from f(\mathcal{G}), confirming that every set in f(\mathcal{F}) contains a set from f(\mathcal{G}), so f(\mathcal{F}) \sqsubseteq f(\mathcal{G}). This monotonicity of subordination under images facilitates the study of filter behavior in quotient spaces and continuous extensions. For preimages, the relation holds under surjective mappings. If f: X \to Y is surjective and \mathcal{H} \sqsubseteq \mathcal{K} on Y (with \mathcal{H} coarser than \mathcal{K}), the preimage filter f^{-1}(\mathcal{H}), generated by \{f^{-1}(C) \mid C \in \mathcal{H}\}, satisfies f^{-1}(\mathcal{H}) \sqsubseteq f^{-1}(\mathcal{K}) on X. Surjectivity ensures all preimages are nonempty, preserving the filter base properties directly: for any D \in \mathcal{H}, D \supset E for some E \in \mathcal{K}, so f^{-1}(D) \supset f^{-1}(E), maintaining the coarser subordination. Without surjectivity, strict preservation may fail, as the generated preimage filter omits sets with empty preimages, potentially coarsening the structure beyond the original relation. A counterexample illustrates this failure for non-surjective f. Let X = \{p, q\} with the discrete topology, Y = \{1, 2\} discrete, and f: X \to Y by f(p) = f(q) = 1 (not surjective). On Y, let \mathcal{K} be the principal filter generated by \{1\} (base \{\{1\}\}), and \mathcal{H} the trivial filter \{Y\} (coarser, as \{1, 2\} \supset \{1\}, so \mathcal{H} \sqsubseteq \mathcal{K}). Then f^{-1}(\mathcal{H}) is the trivial filter on X (base \{X\}). However, f^{-1}(\mathcal{K}) is generated by f^{-1}(\{1\}) = X, again the trivial filter. Thus, f^{-1}(\mathcal{H}) = f^{-1}(\mathcal{K}), losing the strict subordination despite \mathcal{H} \sqsubseteq \mathcal{K} properly. Continuous functions further leverage this preservation to maintain convergence properties through subordination of images. If f: X \to Y is continuous at x \in X and a filter \mathcal{F} on X converges to x (i.e., the neighborhood filter \mathcal{N}_x \sqsubseteq \mathcal{F}), then f(\mathcal{F}) converges to f(x) in Y, as continuity implies \mathcal{N}_{f(x)} \sqsubseteq f(\mathcal{N}_x) \sqsubseteq f(\mathcal{F}) by the subordination preservation theorem. This ensures that images of convergent filters remain convergent, underpinning definitions of continuity in terms of filters.

Products of Prefilters

In topology, given a family of prefilters \mathcal{F}_i on sets X_i for i \in I, the product prefilter \prod_{i \in I} \mathcal{F}_i on the product space \prod_{i \in I} X_i is defined as the collection of all sets of the form \prod_{i \in I} A_i, where A_i \in \mathcal{F}_i for every i \in I. This construction forms a prefilter on the product space, serving as a base for the filter generated by taking all supersets closed under finite intersections. The product prefilter corresponds to the structure underlying the box topology on \prod_{i \in I} X_i, where basic open sets are finite intersections of full cartesian products of open sets from each factor. In contrast, the filter for the product topology (Tychonoff topology) is generated by cylinder sets, which fix sets from finitely many \mathcal{F}_i and take the full product over the remaining coordinates. For finite index sets I, the box and product topologies coincide, but for infinite I, the box topology is strictly finer. The product of filters—that is, the filter generated by the product prefilter—is itself a filter, satisfying the axioms of upward closure and closure under finite intersections. Moreover, subordination of prefilters is preserved under products: if \mathcal{F}_i \subset \mathcal{G}_i for each i \in I (meaning \mathcal{G}_i is coarser than \mathcal{F}_i), then \prod_{i \in I} \mathcal{F}_i \subset \prod_{i \in I} \mathcal{G}_i. This preservation ensures that the product construction respects the partial order on filters. A representative example arises in neighborhood systems: if each \mathcal{N}_{x_i} is the neighborhood prefilter at a point x_i \in X_i in a topological space (X_i, \tau_i), then the product \prod_{i \in I} \mathcal{N}_{x_i} generates the neighborhood filter at the point (x_i)_{i \in I} in the box topology on \prod_{i \in I} X_i, consisting of sets containing a neighborhood of the form \prod_{i \in I} U_i around that point where U_i is a neighborhood of x_i for each i.

Convergence

Filter Limits

In a topological space (X, \tau), a filter \mathcal{F} on X is said to converge to a point x \in X, denoted \mathcal{F} \to x, if every open neighborhood of x belongs to \mathcal{F}. This means that \mathcal{F} contains all sets that are "large" with respect to the local structure at x, capturing the idea that the filter "approaches" x in the topology. Equivalently, the neighborhood filter \mathcal{N}_x of x is coarser than \mathcal{F}, ensuring that convergence is defined purely in terms of the filter's membership properties without reference to sequences or nets. A related but weaker notion is adherence: a point x \in X adheres to the filter \mathcal{F}, denoted \mathcal{F} \vdash x or x \in \overline{\mathcal{F}}, if every open neighborhood of x intersects every member of \mathcal{F}. This condition implies that x is a possible "accumulation" point for the sets in \mathcal{F}, but does not require the neighborhoods themselves to be in \mathcal{F}. Adherence generalizes the closure operator, as the adherence of \mathcal{F} is the intersection of the closures of all sets in \mathcal{F}. Unlike convergence, multiple points may adhere to a given filter. Several key properties distinguish filter limits in different topological settings. In a Hausdorff space, limits of filters are unique: if \mathcal{F} \to x and \mathcal{F} \to y, then x = y, because disjoint neighborhoods of distinct points cannot both belong to the same filter. For ultrafilters specifically, which are maximal filters, convergence occurs to at most one point even in more general spaces, reflecting their "decisive" nature in selecting limits; this property underpins characterizations of compactness, where every ultrafilter converges to some point. Filters may fail to have limits in non-Hausdorff spaces, allowing non-unique convergence, but adherence always identifies the possible limit candidates. A simple example illustrates these concepts: consider the constant filter \mathcal{F}_x generated by the constant sequence (x, x, x, \dots) on X, consisting of all subsets of X containing x. This is a principal ultrafilter, and it converges to x because every neighborhood of x contains x and thus belongs to \mathcal{F}_x. Moreover, x adheres to \mathcal{F}_x, and no other point does in a T_1 space, highlighting how constant filters encode precise pointwise convergence.

Cluster Points

In a topological space X, a point x \in X is a cluster point of a filter \mathcal{F} on X if \mathcal{F} adheres to x, denoted \mathcal{F} \vdash x. This means that every open neighborhood of x intersects every member of \mathcal{F}, or equivalently, x belongs to the closure of every set in \mathcal{F}. Convergence of a filter to a point implies adherence at that point, so every limit point of \mathcal{F} is a cluster point of \mathcal{F}. However, the converse does not hold in general: a filter may adhere to a point without converging to it. This distinction arises prominently in non-Hausdorff spaces, where a converging filter may adhere to multiple points simultaneously. For ultrafilters, the situation simplifies. In a compact space, every ultrafilter \mathcal{U} on X converges to at least one point, and thus has at least one cluster point; moreover, the cluster points of \mathcal{U} coincide exactly with its limit points. In compact Hausdorff spaces, this yields exactly one cluster point. As an illustrative example in \mathbb{R} with the standard topology, consider the filter \mathcal{F} generated by the base \{B_n \mid n \in \mathbb{N}\} where B_n = (-1/n, 1/n) \cup (n, \infty). This filter adheres to 0, since every neighborhood of 0 intersects every B_n via the interval (-1/n, 1/n), making 0 a cluster point. However, \mathcal{F} does not converge to 0, because no B_n is contained in an arbitrary small neighborhood (- \epsilon, \epsilon) of 0, as each B_n extends to infinity. Furthermore, 0 is the only cluster point, since for any y \neq 0, a sufficiently small neighborhood of y avoids the sets (-1/n, 1/n) and misses B_n for sufficiently large n.

Relationships Among Limits

A fundamental relationship in the theory of filters concerns the connection between convergence and cluster points. In a , a filter F converges to a point x \in X if and only if x is the unique cluster point of F, where x is a cluster point of F if there exists a filter G finer than F (i.e., G \supseteq F) such that G converges to x. This equivalence highlights that convergence implies not only adherence but also the exclusion of other accumulation points for the filter. This relationship extends to properties of refinement. Specifically, if F converges to x and G is finer than F (i.e., every set in F belongs to G), then G also converges to x. The contrapositive ensures that non-convergent filters cannot be refined to achieve convergence at that point without altering the cluster structure. Adherence to F at x means x is a cluster point, but limits require uniqueness among such points. In Hausdorff spaces, where every ultrafilter converges to at most one point, cluster points of a filter admit a precise characterization via ultrafilters. Every cluster point of F is the limit of some ultrafilter refinement of F, and these limits are unique for such refinements. As an illustrative example, consider the neighborhood filter \mathcal{N}_x at a point x in any topological space, which converges to x by definition. Any refinement of \mathcal{N}_x, such as the principal filter generated by a smaller neighborhood V of x, also converges to x, preserving the unique cluster point at x. This demonstrates the stability of convergence under refinement in practice.

Functional Limits

Limits via Prefilter Convergence

In topological spaces X and Y, consider a function f: X \to Y and a prefilter \mathcal{F} on X. The limit of f along \mathcal{F} is a point y \in Y, denoted \lim_{\mathcal{F}} f = y, if the pushforward prefilter f(\mathcal{F}) = \{f(A) \mid A \in \mathcal{F}\} converges to y in Y. Convergence of f(\mathcal{F}) to y means that for every neighborhood V of y, there exists A \in \mathcal{F} such that f(A) \subseteq V, or equivalently, the filter generated by f(\mathcal{F}) is finer than the neighborhood filter of y. This generalizes classical limits by allowing \mathcal{F} to capture directional or generalized approaches to points in non-metric spaces. Sequential limits form a special case of this construction. For a sequence (x_n)_{n \in \mathbb{N}} in X, the associated prefilter \mathcal{F} is the eventuality prefilter generated by the sets \{x_n \mid n \geq k\} for k \in \mathbb{N}, consisting of all subsets of X containing a tail of the sequence. Then, \lim_{\mathcal{F}} f = y if and only if f(x_n) \to y in the usual sequential sense, provided the topology on Y admits sequential characterization of convergence. This correspondence highlights how filters extend sequential notions to arbitrary topological spaces where sequences alone may not suffice. A key property relates filter convergence in the domain to function limits via continuity. If \mathcal{F} \to x in X and f is continuous at x, then \lim_{\mathcal{F}} f = f(x). This follows from the preservation of neighborhood filters under continuous maps: the image of the neighborhood filter of x under f refines the neighborhood filter of f(x), ensuring f(\mathcal{F}) converges to f(x). In non-Hausdorff spaces, multiple limits may exist, but uniqueness holds in Y. For an illustrative example, consider X = Y = \mathbb{R} with the standard topology and f(x) = \sin(1/x) for x \neq 0, extended arbitrarily at 0. Let \mathcal{F} be the prefilter on \mathbb{R} generated by the shrinking intervals ( -1/n, 1/n ) for n \in \mathbb{N}, which converges to 0. Then \lim_{\mathcal{F}} f does not exist, as f(\mathcal{F}) oscillates and adheres to both 1 and -1, reflecting the function's behavior along paths approaching 0 without settling to a unique value. This demonstrates how prefilter limits detect oscillatory or indeterminate behaviors in classical examples.

Continuity and Uniform Convergence

In topological spaces, a function f: X \to Y is continuous at a point x \in X if and only if, for every filter \mathcal{F} on X converging to x, the image filter f(\mathcal{F}) converges to f(x) in Y. This characterization generalizes the sequential notion of continuity, allowing it to apply in spaces where sequences may not suffice to describe limits, such as non-first-countable topologies. In uniform spaces, uniform continuity extends this idea by incorporating the uniform structure, typically defined via entourages. A function f: (X, \mathcal{U}_X) \to (Y, \mathcal{U}_Y) between uniform spaces is uniformly continuous if, for every entourage E \in \mathcal{U}_Y, the preimage (f \times f)^{-1}(E) belongs to \mathcal{U}_X. Equivalently, the filter generated by the image of \mathcal{U}_X under f \times f is finer than the entourage filter \mathcal{U}_Y. This ensures that the function preserves the "uniform closeness" defined by the entourages, distinguishing it from mere topological continuity. For sequences of functions, uniform convergence along a Cauchy filter provides a concrete illustration. In a uniform space, if a net of continuous functions f_\alpha: X \to Y converges uniformly to f along a Cauchy filter \mathcal{C} on the index set (meaning f_\alpha(\mathcal{C}) converges to f in the uniform structure on the function space), then f is continuous, inheriting the joint continuity properties from the approximants. A key distinction arises between pointwise and uniform filter convergence for series or sequences of functions. Pointwise convergence requires that, for each fixed point x \in X, the filter of tails converges to the limit value at x via the evaluation filter at that point; in contrast, uniform convergence demands that the filter of tails converges in the uniform structure on the space of functions, ensuring the convergence is "global" and independent of the point. This uniform version implies pointwise convergence but not vice versa, as seen in examples like the sequence of functions f_n(x) = n x e^{-n x} on \mathbb{R}, which converges pointwise to 0 but not uniformly, since the maximum value $1/(2e) does not tend to 0.

Nets and Filters

Constructing Prefilters from Nets

In topological spaces, nets provide a means to generalize sequences for describing convergence, and they naturally induce prefilters on the underlying set. Given a net (x_\alpha)_{\alpha \in A} in a set X, where A is a directed set, the associated prefilter \mathcal{F} on X is generated by the base consisting of the images of the tails of the net. Specifically, for each \alpha \in A, define the tail A_\alpha = \{\beta \in A \mid \beta \geq \alpha\}, and let B_\alpha = x_\alpha(A_\alpha) = \{x_\beta \mid \beta \in A_\alpha\}. The family \{B_\alpha \mid \alpha \in A\} forms a base for \mathcal{F}, as these sets are directed under inclusion and their finite intersections remain within the family up to subordination. This construction ensures that \mathcal{F} captures the "eventual" behavior of the net, abstracting away the specific indexing while preserving the directional order. A key property is that, in a topological space (X, \tau), the net (x_\alpha) converges to a point x \in X if and only if the prefilter \mathcal{F} converges to x, meaning every neighborhood of x belongs to \mathcal{F}. This equivalence highlights the duality between nets and prefilters in defining limits, allowing proofs involving one to transfer to the other. For a concrete illustration, consider a constant net where x_\alpha = x for all \alpha \in A and some fixed x \in X. Here, each tail image B_\alpha = \{x\}, so the base \{B_\alpha\} consists solely of the singleton \{x\}, generating the principal prefilter \mathcal{F}_x with base \{\{x\}\}. This prefilter converges precisely to x in any topology containing \{x\} as a closed set. More broadly, this generation process demonstrates how prefilters generalize by replacing the concrete directed set A with an abstract family of subsets, enabling the study of convergence without reliance on a particular indexing structure. This abstraction is foundational in modern , facilitating uniform treatments of limits in non-first-countable spaces.

Constructing Nets from Prefilters

In topology, given a prefilter \mathcal{F} on a set X, one can construct a net (x_\alpha)_{\alpha \in D} whose derived filter (image filter) is equivalent to \mathcal{F}. To do so, regard \mathcal{F} as a directed set D = \mathcal{F} ordered by reverse inclusion, where A \leq B if and only if A \supseteq B. Using the axiom of choice, select a point x_A \in A for each A \in \mathcal{F}, and define the net by x(A) = x_A. The tails of this net correspond to the sets in \mathcal{F}, ensuring that the derived filter consists precisely of the sets S \subseteq X such that some tail is contained in S, which yields equivalence to \mathcal{F}. This construction may be refined by applying Zorn's lemma to obtain a net indexed by a maximal directed set. Consider the partially ordered set of all directed subsets of X that are cofinal with respect to \mathcal{F} (in the sense of reverse inclusion), partially ordered by inclusion; Zorn's lemma guarantees a maximal element, providing a maximal directed set D' for the index. Choosing points via the axiom of choice on D' yields a net whose derived filter remains equivalent to \mathcal{F}, as the maximality ensures no further refinement alters the tail structure. A particular instance of this construction is the universal net associated to \mathcal{F}, where the indices are the sets in \mathcal{F} themselves, ordered by reverse inclusion, and the net selects an arbitrary point from each set. This net is "universal" in the sense that its derived filter is exactly \mathcal{F}, and any subnet preserves the equivalence due to the cofinal nature of the tails. Equivalence of derived filters is maintained under this process, as subordination relations between prefilters correspond directly to subnet inclusions. A key property is that every prefilter on a set is the derived filter of some net; the above constructions establish this via the directed structure of prefilters and the axiom of choice (or equivalently, for maximal extensions). Conversely, as noted in the construction of prefilters from nets, this duality ensures bidirectional embedding without loss of convergence information. For a concrete example, consider the Fréchet prefilter \mathcal{F} on the natural numbers \mathbb{N}, consisting of all cofinite subsets (subsets with finite complement). The identity net (x_n)_{n \in \mathbb{N}} defined by x_n = n has tails \{n, n+1, \dots \} that are cofinite, generating precisely the Fréchet prefilter as its derived filter.

Subnets and Subordinate Filters

In the theory of nets in topological spaces, a subnet provides a generalization of the subsequence concept from sequences to more general directed index sets. Specifically, given a net (x_\alpha)_{\alpha \in A} in a set X, where A is a directed set, a net (x_\beta)_{\beta \in B} in X, with B directed, is a subnet if there exists an increasing map \phi: B \to A such that \phi(B) is cofinal in A and x_\beta = x_{\phi(\beta)} for all \beta \in B. This cofinality ensures that the subnet "eventually" covers the original net's progression, analogous to how a subsequence selects infinitely many terms from a sequence. A key property of subnets is their preservation of convergence: if the original net (x_\alpha) converges to a point x \in X in a topological space, then every subnet (x_\beta) also converges to x. This mirrors the behavior of subsequences in metric spaces, where limits are preserved under extraction. Subnets thus serve as a tool for analyzing cluster points and compactness, for instance, by guaranteeing the existence of convergent subnets in compact spaces. In the dual framework of filters, the analogous structure is a subordinate filter. Given filters \mathcal{F} and \mathcal{G} on X, \mathcal{G} is subordinate to \mathcal{F} (denoted \mathcal{G} \precsim \mathcal{F}) if \mathcal{F} \subseteq \mathcal{G}, meaning \mathcal{G} is finer than \mathcal{F} and refines it by including all sets from \mathcal{F} along with additional smaller sets. For a net inducing the filter \mathcal{F}, a subordinate filter \mathcal{G} corresponds to the induced filter of some subnet of the original net, preserving convergence in the sense that if \mathcal{F} converges to x, then so does \mathcal{G}. This duality highlights how subnets and subordinate filters both capture "refinements" that retain topological limits, much like subsequences and their induced tails in sequences. However, the correspondence between subordinate filters and subnets is not perfect without additional set-theoretic assumptions. Not every subordinate filter arises as the induced filter of a subnet of a given net generating the coarser filter; constructing such a subnet generally requires the axiom of choice to select appropriate cofinal mappings aligned with the finer structure. This non-equivalence underscores a subtle distinction in the net-filter duality, where filters offer a more set-inclusion-based refinement without explicit indexing.

Topologies via Filters

Neighborhood Filters and Bases

In a topological space (X, \tau), the neighborhood filter N(x) at a point x \in X is the filter generated by the collection of all open neighborhoods of x, consisting of all subsets of X that contain an open set containing x. This filter satisfies the standard filter axioms: it is nonempty, closed under finite intersections, and upward closed with respect to supersets. In particular, N(x) converges to x in the sense of filter convergence, as every member of N(x) contains x and the filter is generated by sets adhering to x. A fundamental system (or base) for the neighborhood filter N(x) is a filter base \mathcal{B}(x) such that the filter it generates coincides with N(x), meaning every neighborhood of x contains some element of \mathcal{B}(x). Such bases provide a local characterization of the topology at x; for instance, in , countable local bases exist for each N(x). The neighborhood filter N(x) plays a central role in defining convergence, where a filter \mathcal{F} on X converges to x if and only if N(x) \subseteq \mathcal{F}. A concrete example arises in metric spaces (X, d), where the open balls B(x, r) = \{ y \in X \mid d(x, y) < r \} for r > 0 form a for N(x), as every open neighborhood of x contains some such ball. This base is countable if the metric allows enumeration of positive radii, illustrating how structures yield explicit generators for neighborhood filters. A family \mathcal{B} \subseteq \tau serves as a for the \tau if and only if, for every x \in X, the trace \mathcal{B}(x) = \{ B \in \mathcal{B} \mid x \in B \} is a for N(x). This equivalence highlights the pointwise nature of neighborhood filters in structuring the global .

Defining Topologies

A topology on a set X can be generated from a family of filters \{ \mathcal{F}_x \mid x \in X \}, where each \mathcal{F}_x is a filter on the power set of X satisfying x \in \bigcap \mathcal{F}_x (i.e., x belongs to every member of \mathcal{F}_x). The open sets in this topology are precisely the subsets U \subseteq X such that for every x \in U, there exists V \in \mathcal{F}_x with V \subseteq U. This definition ensures that the resulting space is topological, with the filters \mathcal{F}_x serving as the neighborhood filters at each point x, and it is the unique such topology compatible with these neighborhood systems. More generally, topologies can be defined via convergence structures specified by filters. Given a collection of pairs ( \mathcal{F}, x ), where \mathcal{F} is a filter on X designated to converge to x \in X, the associated convergence topology is the finest topology on X (i.e., the one with the most open sets) in which all these specified convergences hold. In this topology, a subset U \subseteq X is open if and only if, for every x \in U and every filter \mathcal{F} converging to x, U \in \mathcal{F}. This construction aligns filter convergence with the standard notion: \mathcal{F} converges to x if every open neighborhood of x belongs to \mathcal{F}. A key example arises in the context of initial topologies induced by mappings. Consider a set X and a family of continuous maps \{ f_i : X \to (Y_i, \tau_i) \mid i \in I \}, where each (Y_i, \tau_i) is a . The initial topology on X is the coarsest topology making all f_i continuous, generated as the topology with subbasis \{ f_i^{-1}(O_i) \mid i \in I, O_i \in \tau_i \}. Equivalently, in terms of filters, the neighborhood filter \mathcal{N}_x at each x \in X is the finest filter such that f_i(\mathcal{N}_x) \subseteq \mathcal{N}_{f_i(x)} for all i \in I, or more concretely, the filter generated by sets of the form f_i^{-1}(N_i) where N_i is a neighborhood of f_i(x) in Y_i. This filter-based description determines the topology uniquely when the \mathcal{N}_x are the neighborhood filters. The neighborhood filters thus fully determine the topology: two topologies on X coincide if and only if their neighborhood filters agree at every point. This equivalence extends to convergence, as filters capture the topological structure precisely when they serve as the pointwise neighborhood systems.

Topological Properties

In topological spaces, the Hausdorff separation axiom can be characterized using filters as follows: a space X is Hausdorff if and only if every convergent filter on X converges to at most one point. Compactness admits a filter-based : a X is compact every on X has at least one in X. This condition is equivalent to the statement that every ultrafilter on X converges to some point in X, highlighting the role of maximal filters in capturing the "fullness" of the space. In such spaces, no can "escape" without adhering to some point, preventing the existence of covers without finite subcovers. A space X is first-countable if, for every point x \in X, the \mathcal{N}(x) admits a countable base. This countable base consists of a of neighborhoods \{B_n\}_{n \in \mathbb{N}} such that every neighborhood of x contains some B_n, allowing and to be determined via sequences rather than more general filters or nets. In a compact , every ultrafilter converges to its unique limit point, as the compactness ensures a limit point exists and the Hausdorff property guarantees uniqueness. For instance, in the unit interval [0,1] with the standard , which is compact and Hausdorff, any ultrafilter on [0,1] must converge to a single point in [0,1], reflecting the space's completeness in terms of filter limits.

Applications

Cauchy Prefilters in Uniformities

A is a set X equipped with a filter \mathcal{U} on X \times X, called the uniformity, consisting of subsets known as entourages that satisfy certain axioms: the uniformity contains the diagonal \Delta_X = \{(x,x) \mid x \in X\}, is symmetric (closed under inversion), and satisfies a triangle inequality in the sense that for every entourage E \in \mathcal{U}, there exists E' \in \mathcal{U} such that E' \circ E' \subseteq E, where \circ denotes the composition of relations. The uniformity \mathcal{U} generates a on X by defining, for each point x \in X, a neighborhood basis consisting of the slices E = \{y \in X \mid (x,y) \in E\} for E \in \mathcal{U}. In this context, a prefilter \mathcal{F} on X (a directed family of subsets closed under finite intersections and generating a filter) is called a Cauchy prefilter if for every E \in \mathcal{U}, there exists A \in \mathcal{F} such that A \times A \subseteq E. This condition ensures that the elements of \mathcal{F} become arbitrarily "small" with respect to the uniformity, generalizing the notion of Cauchy sequences in metric spaces. A uniformity \mathcal{U} is complete if every Cauchy (the filter generated by a Cauchy prefilter) converges in the induced to some point in X. For example, in a (X, d), the uniformity is generated by the entourages E_\epsilon = \{(x,y) \in X \times X \mid d(x,y) < \epsilon\} for \epsilon > 0. A (x_n) in X is Cauchy if for every \epsilon > 0, there exists N such that d(x_m, x_n) < \epsilon for all m,n \geq N; the prefilter generated by the tails \{ \{x_n \mid n \geq k\} \mid k \in \mathbb{N} \} is then a Cauchy prefilter in this uniformity. Filter convergence in the topology induced by the uniformity aligns with the Cauchy condition in complete spaces, where every Cauchy prefilter adheres to a unique limit point.

Topology on Filter Spaces

The set Pref(X) of all prefilters on a set X can be endowed with a natural topology, known as the filter topology or the topology of subordination. This topology has a base consisting of the sets U_A = \{ \mathcal{F} \in \Pref(X) \mid A \in \mathcal{F} \} for each non-empty subset A \subseteq X. These basic open sets capture the idea of prefilters "containing" a fixed set A, and the topology is the coarsest one making the evaluation maps \mathcal{F} \mapsto \mathbf{1}_A(\mathcal{F}) (where \mathbf{1}_A(\mathcal{F}) = 1 if A \in \mathcal{F} and 0 otherwise) continuous for all A, viewing the codomain as the discrete space {0,1}. In this topology, convergence of nets of prefilters is defined as follows: a net (\mathcal{F}_\lambda)_{\lambda \in \Lambda} in Pref(X) converges to a prefilter \mathcal{F} \in \Pref(X) if, for every A \in \mathcal{F}, there exists \lambda_0 \in \Lambda such that A \in \mathcal{F}_\lambda for all \lambda \geq \lambda_0. This notion aligns with the subordination relation between prefilters, where \mathcal{F}_\lambda becomes "coarser" than \mathcal{F} eventually, in the sense of inclusion reversed. The space Pref(X) with this topology is always compact and Hausdorff, regardless of any additional structure on X. Compactness follows from identifying Pref(X) as a closed subspace of the product space \prod_{A \subseteq X, A \neq \emptyset} \{0,1\} equipped with the , where each coordinate corresponds to membership of A in the prefilter; since filters and prefilters satisfy closure under supersets and finite intersections (with the empty set excluded for proper ones), this embedding preserves compactness via . The Hausdorff property holds because if \mathcal{F} \neq \mathcal{G}, then either there exists B \in \mathcal{F} \setminus \mathcal{G} (separating via U_B) or C \in \mathcal{G} \setminus \mathcal{F} (separating via U_C). When X is finite, Pref(X) reduces to a finite , hence compact Hausdorff; for infinite X, the cardinality of Pref(X) is $2^{2^{|X|}}, underscoring its largeness while maintaining these . A representative example involves principal prefilters generated by singletons (point filters). For x \in X, the principal prefilter \mathfrak{m}(x) = \{ B \subseteq X \mid x \in B \}. Consider a (x_\lambda)_{\lambda \in \Lambda} in X, and the corresponding net of point filters \mathfrak{m}(x_\lambda). This net converges to \mathfrak{m}(x) in the topology , for every A \ni x, eventually x_\lambda \in A for all \lambda \geq \lambda_0, which holds precisely when the net (x_\lambda) is eventually constantly equal to x. If X carries the discrete topology, the X \to \Pref(X) given by x \mapsto \mathfrak{m}(x) is continuous and identifies X with a dense of the point filters within Pref(X).

Extensions to Compactifications

The Stone-Čech compactification of a X, denoted \beta X, can be constructed using ultrafilters on the underlying set of X when X is equipped with the discrete . Specifically, \beta X is the set of all ultrafilters on X, endowed with the generated by the basis consisting of sets U_A = \{\mathcal{U} \in \beta X : A \in \mathcal{U}\} for subsets A \subseteq X. The points of X embed densely into \beta X via principal ultrafilters, where each x \in X corresponds to the ultrafilter \{B \subseteq X : x \in B\}, while the remaining points in \beta X \setminus X are (non-principal) ultrafilters. This construction yields a compact containing X as a dense , satisfying the universal property that every from X to a compact Y extends uniquely to a from \beta X to Y. Convergence in \beta X is characterized in terms of filters: a filter \mathcal{F} on X converges to a point p \in \beta X if and only if p lies in the closure of \mathcal{F} in \beta X, which is equivalent to \mathcal{F} \subseteq p as collections of subsets. In this setting, every filter \mathcal{F} on X extends to an ultrafilter on X, and the corresponding point in \beta X serves as a limit point for \mathcal{F}. This extension property underscores the role of filters in compactifying spaces, as \beta X universally embeds X while preserving and extending filter-based convergence. A prominent example arises with X = \mathbb{N}, the natural numbers under the discrete topology, where \beta \mathbb{N} consists of all ultrafilters on \mathbb{N}. Free ultrafilters in \beta \mathbb{N} \setminus \mathbb{N} enable the definition of limits along these ultrafilters, which capture various asymptotic behaviors of s, such as different growth rates or densities at infinity. For instance, evaluating a bounded a_n at such an ultrafilter p yields \lim_{n \to p} a_n, providing a generalized that distinguishes, say, logarithmic from growth in the "direction" defined by p. In modern applications, ultrafilters from \beta \mathbb{N} play a key role in non-standard analysis, where they construct hyperreal numbers via ultrapowers, allowing infinitesimals and infinite quantities that model intuitive notions of growth and approximation in analysis and beyond. This connection highlights how the filter-based compactification \beta X facilitates rigorous treatments of limits and in non-archimedean settings, extending classical to non-standard models.

References

  1. [1]
    4.3 Filters - Math 581: Topology 1
    4.3 Filters. 4.3.1 Definition of a filter. Let X be a set. A filter on X is a family F of subsets of X such that: 1. F≠∅ and ∅∉F;. 2. if F∈F and F⊆F′⊆X, ...
  2. [2]
    [PDF] filter convergence and tychonoff's theorem
    A filter defines a notion of largeness for subsets of S: the large sets are the members of S, and the small sets are the sets whose complements are large. An ...
  3. [3]
    [PDF] The emergence of the concept of filter in topological categories - arXiv
    Jul 19, 2017 · Filters on a set have been introduced in 1937 by Henri Cartan (see [2, 3]) to replace sequences in the study of a general topological space (if ...
  4. [4]
    [PDF] The Use Of Filters In Topology - ucf stars
    The study of filters is a very natural way to describe convergence in general topological space. Filters were introduced in 1937 by Cartan (1937 a,b). Bourbaki ...
  5. [5]
    [PDF] Filters in Analysis and Topology. - Home | Department of Mathematics
    Feb 7, 2019 · Tychonoff's theorem in point set topology states that an arbitrary product of compact spaces is compact in the product topology.
  6. [6]
    General topology : Bourbaki, Nicolas : Free Download, Borrow, and ...
    Aug 8, 2022 · General topology. by: Bourbaki, Nicolas ... 1.0G. 2 volumes. Translation of: Topologie générale. Includes index [1] Chapters 1-4 -- [2] Chapters 5 ...Missing: 1940 PDF
  7. [7]
    A General Theory of Limits - jstor
    MOORE AND H. L. SMITH. Introduction. I. As to Notions.-In this paper we investigate a simple general limit.
  8. [8]
    H. Cartan, “Théorie des Filtres,” Comptes Rendus de l'Académie des ...
    H. Cartan, “Théorie des Filtres,” Comptes Rendus de l'Académie des Sciences, Paris, Vol. 205, 1937, pp. 595-598.
  9. [9]
    [PDF] A Tribute to Henri Cartan
    Mathematical Work Henri Cartan worked on many subjects but there was one to which he was particularly attached, and that was the theory of functions of several ...Missing: Smith | Show results with:Smith
  10. [10]
    Bourbaki 1 - MacTutor History of Mathematics
    ... Bourbaki published. The second Bourbaki publication came in 1940 when Chapters 1 and 2 of Book III appeared:- The first chapter entitled "Topological ...
  11. [11]
    Proofs in expository writing — Some examples from Bourbaki's early ...
    Cartan, H. (1937a). Théorie des filtres.Comptes rendus de l'Académie des Sciences de Paris,205, 595–596. Google Scholar. Cartan, H. (1937b). Filtres et ...<|separator|>
  12. [12]
    [PDF] Sequences and nets in topology - arXiv
    Jun 23, 2010 · Convergence of sequences works fine when the space is first countable, because a countable basis at a point allows us to approach that point ...
  13. [13]
    Full text of "General Topology" - Internet Archive
    This book is designed to develop the fundamental concepts of general topology which are the basic tools of working mathematicians in a variety of fields.
  14. [14]
    [PDF] Filters and ultrafilters
    The Fréchet filter on an infinite set is not an ultrafilter, and the neighbourhood filter of a point x in a topological space X is an ultrafilter iff x is an ...
  15. [15]
    [PDF] FILTERS AND ULTRAFILTERS 1. Filters Given a set X, a filter is a ...
    The canonical example of a free filter is the set of cofinite subsets of X. This is called the Frechet filter on X. Here are some important propositions:.
  16. [16]
    [PDF] The meaning of infinity in calculus and computer algebra systems
    We will represent ∞ and its cousins indeterminate and interval by filters over some underlying topological space. In calculus textbooks this is the space of ...
  17. [17]
    [PDF] ultrafilters, compactness, and the stone-ˇcech compactification
    the natural numbers with the discrete topology. The results in this section all have analogues ...
  18. [18]
    Comparison of countability conditions within three fundamental ...
    Jun 1, 2020 · We call kernel of F ∈ F X the set ker ⁡ F : = ⋂ F ∈ F F ... free filter. But if G n ∈ F ∘ there is F n ∈ F with F n ∩ ( ker ...
  19. [19]
    [PDF] convergence.pdf - mimuw
    Then there is the rival theory of filters, discovered by Henri. Cartan in 1937 amidst a Séminaire Bourbaki. Cartan successfully convinced his fel- low ...
  20. [20]
    [PDF] GeneralTopologyWillard.pdf - Rexresearch1.com
    ... coarser) than τ. 2. , or τ. 2 is stronger (larger, finer) than τ. 1 iff τ. 1 ⊂ τ ... Frechet filter on R has no cluster points. c) Let ℱ be the filter on R ...
  21. [21]
  22. [22]
  23. [23]
    [PDF] Notes on Topology - UC Berkeley math
    Dec 3, 2010 · The following proposition shows that the filter of neighborhoods of a point p in a topological space is the infimum of the elementary filters of ...
  24. [24]
    [PDF] Theory of Distributions
    ... kernel of f and the range of f> ..................... 142. 7.4.2. Some ... free filter if the intersection of all of its members is empty, whereas F is ...
  25. [25]
    None
    ### Summary of Filters in Topology from Bourbaki's General Topology (Chapters 1-4)
  26. [26]
  27. [27]
    General Topology: Chapters 1-4 - N. Bourbaki - Google Books
    Aug 3, 1998 · 5 Induced filters. 61. 6 Direct image and inverse image of a filter base. 62. 7 Product of filters. 63. 8 Elementary filters. 64. 9 Germs with ...
  28. [28]
    General Topology - John L. Kelley - Google Books
    Aimed at graduate math students, this classic work is a systematic exposition of general topology and is intended to be a reference and a text.Missing: filter kernel
  29. [29]
    [PDF] Chapter 9 Convergence
    Both nets and filters are sufficient to describe the topology in any space, so we should be able to use them to describe continuous functions. Theorem 5.12 ...<|control11|><|separator|>
  30. [30]
    [PDF] MAST30026 Resources, Arun Ram, July 17, 2022
    Jul 17, 2022 · The space (X,T ) is filter compact if every filter has a cluster point. • The space (X,T ) is ultrafilter compact if every ultrafilter has a ...
  31. [31]
    [PDF] Topological Vector Spaces
    Jan 2, 2015 · Given a filter F in a topological space X, we say that it converges to a point x ∈ X if every neighborhood of x belongs to F, in other words if ...
  32. [32]
    [PDF] Ultrafilters and Topology - Universiteit Leiden
    Jun 13, 2014 · The Eilenberg-Moore category of the ultrafilter monad is equivalent to the category of compact Hausdorff spaces. We will start with the ...Missing: cluster | Show results with:cluster
  33. [33]
    General Topology: Chapters 1–4 - SpringerLink
    Free delivery 14-day returnsThis is the softcover reprint of the English translation of 1971 (available from Springer since 1989) of the first 4 chapters of Bourbaki's Topologie générale.
  34. [34]
    [PDF] Filters in Analysis and Topology - David R. MacIver
    Jul 1, 2004 · Proof: We shall show the equivalence with the closed set definition of compactness. First suppose X is compact. Let F be a filter. Then Z = {U : ...
  35. [35]
    [PDF] Notes on Uniform Structures - Berkeley Math
    Dec 13, 2013 · Definition 2.2 A set X equipped with a uniform structure U is called a uniform space and the filter U is often referred to as its uniformity.
  36. [36]
    General topology : Kelley, John L - Internet Archive
    Jul 5, 2012 · Images An illustration of a heart shape Donate An illustration of ... General topology. by: Kelley, John L. Publication date: 1955. Topics ...
  37. [37]
    [PDF] The Connection between Nets and Filters
    J.L. Kelley, "General Topology", Van Nostrand (1955). M.F. Smiley, "Filters and Equivalent Nets", American Math. Monthly, 64 (1957), 336-338. University ...
  38. [38]
    [PDF] Nets and filters (are better than sequences)
    In this language, the proposition above says that every closed subset of a topological space is sequentially closed, though the converse is not true. Next, we ...<|control11|><|separator|>
  39. [39]
    [PDF] Convergence in Topological Spaces. (1970)
    J. L. Kelley, General Topology, D. VanNostrand Co., Inc.,. Princeton, 1955. 6. E. J. McShane, "Partial Orderings and Moore-Smith Limits",.
  40. [40]
    [PDF] Topology - I - UChicago Math
    Note that in general, a sequence can converge to more than one point, but this is not possible in a Hausdorff space such as a metric space. Exercise 5 Show that ...<|control11|><|separator|>
  41. [41]
    [PDF] Nets and Filters
    Since a topology is defined by its open sets, and we've just shown that open sets can be defined by convergent nets, we can deduce the following: 14. Page 15 ...
  42. [42]
    Ultrafilters, nonstandard analysis, and epsilon management
    Jun 25, 2007 · This sequence might not converge in any reasonable topology, but it often lies in a sequentially compact set in some weak topology (e.g. by ...