Fact-checked by Grok 2 weeks ago

General topology

General topology, also known as point-set topology, is a branch of concerned with the study of topological spaces, which are sets equipped with a called a consisting of a collection of subsets known as open sets that satisfy specific axioms: the and the whole space are open, arbitrary unions of open sets are open, and finite intersections of open sets are open. This framework generalizes the notions of , limits, and neighborhoods from metric spaces to more settings, allowing the of properties preserved under continuous deformations. Unlike algebraic topology, which uses algebraic tools like to study global properties such as groups, general topology relies on set-theoretic methods to examine local properties and separation axioms. The development of general topology emerged in the early 20th century as a unification of ideas from , , and , with foundational contributions from , who formalized the concept of in his 1914 book Grundzüge der Mengenlehre. Earlier precursors include 19th-century work on continuous functions and , such as by and the Heine-Borel theorem attributed to . The axiomatic approach was further developed in the 1920s, with key contributions from mathematicians like Maurice Fréchet and . By the mid-20th century, general topology had become a cornerstone of modern mathematics, influencing fields from to . Central concepts in general topology include continuous functions, which preserve the topological structure by mapping open sets to open sets; compactness, a property ensuring every open cover has a finite subcover, generalizing boundedness in spaces; and connectedness, which describes spaces that cannot be divided into disjoint nonempty open subsets. Separation axioms, such as T1 (points are closed) and Hausdorff (, distinct points have disjoint neighborhoods), classify topological spaces by their ability to distinguish points. Bases and subbases provide ways to generate topologies, while and interior operators extend the structure to closed sets and other derived notions. These tools enable the study of , , and metrizability, with applications in understanding the real line, function spaces, and abstract geometric objects.

History

Early development

The foundations of general topology trace back to late 19th-century advancements in analysis and geometry, where mathematicians began addressing properties invariant under continuous transformations. Bernhard Riemann's work on Riemann surfaces, introduced in his 1851 doctoral dissertation, provided a geometric model for multi-valued functions in , incorporating notions of , branching, and that highlighted qualitative features beyond metric considerations. These ideas influenced the study of surfaces as abstract objects, foreshadowing topological invariants. Parallel developments in by during the 1870s and 1890s contributed essential tools for point-set topology. In his 1872 paper published in Mathematische Annalen, Cantor defined the derived set as the collection of limit points of a given set and characterized closed subsets of line as those containing all their limit points, laying the groundwork for concepts of accumulation, , and in terms of sets rather than functions. Cantor's subsequent explorations of perfect sets and the in the 1880s further emphasized dense, nowhere-dense structures, bridging with emerging topological ideas. A significant intuitive leap occurred in 1895 with Henri Poincaré's seminal paper "Analysis Situs," published in the Journal de l'École Polytechnique, which focused on the "analysis of position" or qualitative geometry of manifolds. Poincaré introduced groups and Betti numbers to classify surfaces and curves up to , emphasizing invariants like and that resist deformation, thus motivating as a distinct discipline from metric geometry. This work influenced the shift toward abstract, deformation-invariant properties. In 1906, Maurice Fréchet advanced abstraction in his thesis "Sur quelques points du calcul fonctionnel," published in Rendiconti del Circolo Matematico di , by defining spaces as sets equipped with a distance satisfying basic axioms, independent of any embedding in . This framework unified prior work on spaces and , serving as a direct precursor to non-metric topologies by generalizing notions of nearness. Felix Hausdorff's 1914 monograph Grundzüge der Mengenlehre synthesized these threads into a rigorous . Hausdorff formalized topological spaces using neighborhood filters with four axioms—symmetry, inclusion, minimality, and intersection—while introducing separation axioms like T1 and to distinguish points, enabling the study of abstract spaces without reliance on metrics or order. This text established general topology as a foundational branch of mathematics, paving the way for axiomatic refinements.

Key contributors and milestones

In the 1920s, the Polish school of mathematics, centered in and Lwów, advanced general topology through axiomatic approaches, with Kuratowski playing a pivotal role by introducing the closure operator axioms in 1922, utilizing to define topological structures independently of points. This work complemented earlier efforts and solidified the school's emphasis on point-set topology, as seen in foundational publications like Fundamenta Mathematicae starting in 1920. Concurrently, the Soviet school, led by and Pavel Urysohn, developed the neighborhood system axioms for topological spaces in their 1924 collaboration, providing an intrinsic characterization of and without reliance on metrics. Their joint paper "Zur Theorie der topologischen Räume" formalized these ideas, establishing a framework that influenced subsequent axiomatizations. A landmark achievement was Urysohn's metrization theorem, announced in 1924 and published posthumously in 1925, stating that a second-countable, regular is metrizable. The proof outline involves first applying to construct continuous functions separating points from closed sets in the normal space, then embedding the space into a product of intervals via these functions to yield a countable basis, and finally deriving a from the . In 1930, Andrey Tychonoff proved that the product of any collection of compact topological spaces is compact in the product topology, a result that extended finite-product compactness to arbitrary families and relied on the axiom of choice. This theorem, detailed in his paper "Über die topologische Erweiterung von Räumen," became foundational for infinite-dimensional topology. During the 1930s, Andrey Kolmogorov contributed to the structure of Borel sets by investigating the hierarchy of Borel classes and posing problems on its length, as in his 1935 query in Fundamenta Mathematicae. Roman Sikorski advanced related work on Boolean algebras representing Borel structures, linking set-theoretic operations to topological measurability in Polish topological contexts. Post-World War II developments included André Weil's 1937 introduction of structures, providing a beyond metrics for notions like and . Richard Arens' 1946 work developed topologies for spaces of transformations using these uniform ideas, enabling the study of in spaces. In the 1940s, Norman Steenrod's work on fiber bundles, culminating in his 1951 monograph The Topology of Fibre Bundles, provided axiomatic foundations that bridged general and , influencing classifications and applications.

Topological spaces

Definition of a topology

In general topology, a topological space is formally defined as a pair (X, \tau), where X is a set and \tau is a collection of subsets of X satisfying certain axioms; this axiomatic framework, equivalent to the closure operator axioms introduced by Kazimierz Kuratowski in 1922, provides the foundation for abstracting notions of continuity and proximity without relying on a specific metric. The collection \tau is called a topology on X, and its elements are termed open sets. This definition assumes familiarity with basic set-theoretic operations, such as unions and intersections of families of sets. The axioms for \tau to be a topology on X are as follows:
  1. The empty set \emptyset and the whole set X belong to \tau.
  2. The union of any (possibly infinite) collection of sets in \tau is again in \tau.
  3. The intersection of any finite collection of sets in \tau is again in \tau.
These properties ensure that \tau captures the intuitive behavior of "open" regions in familiar spaces like the real line with the standard topology. A subset C \subseteq X is defined to be closed if its complement X \setminus C is open (i.e., belongs to \tau); equivalently, the closed sets form the collection of complements of open sets, and they satisfy the dual properties: containing X and \emptyset, closed under finite unions, and closed under arbitrary intersections. This duality between open and closed sets arises directly from the axioms and allows for flexible formulations of topological concepts. From the \tau, several derived operators can be defined on subsets of X. The interior of a set A \subseteq X, denoted \operatorname{int}(A), is the largest open set contained in A, or equivalently, the of all open sets subsets of A. The closure of A, denoted \operatorname{cl}(A) or \overline{A}, is the smallest closed set containing A, or equivalently, \operatorname{cl}(A) = X \setminus \operatorname{int}(X \setminus A). The boundary of A is \operatorname{bd}(A) = \operatorname{cl}(A) \setminus \operatorname{int}(A), and the exterior of A is \operatorname{int}(X \setminus A). These operators satisfy properties like \operatorname{int}(A) \subseteq A \subseteq \operatorname{cl}(A) and idempotence (\operatorname{cl}(\operatorname{cl}(A)) = \operatorname{cl}(A)), mirroring the closure axioms. Trivial examples illustrate the definition's scope. For the empty set as the underlying space, X = \emptyset and \tau = \{\emptyset\} forms a topological space, satisfying the axioms vacuously. Similarly, for a singleton space X = \{p\}, the topology \tau = \{\emptyset, \{p\}\} is valid, with both sets open (and closed). These cases highlight the generality of the definition, applicable even to degenerate spaces. This structure underpins the definition of continuous functions between topological spaces, where a map f: (X, \tau) \to (Y, \sigma) is continuous if the preimage of every open set in \sigma is open in \tau.

Bases and subbases

A (or basis) for a \tau on a set X is a subcollection \mathcal{B} \subseteq \tau of s such that every open set in \tau can be expressed as an arbitrary union of elements from \mathcal{B}. Moreover, the union of all sets in \mathcal{B} must equal X. To determine whether a given collection \mathcal{B} of subsets of X forms a basis for some topology on X, it must satisfy two conditions: first, the union of elements in \mathcal{B} covers X; second, for any two elements B_1, B_2 \in \mathcal{B} and any point x \in B_1 \cap B_2, there exists an element B_3 \in \mathcal{B} such that x \in B_3 \subseteq B_1 \cap B_2. If \mathcal{B} meets these criteria, the topology generated by \mathcal{B} consists of all arbitrary unions of elements from \mathcal{B}, and \mathcal{B} serves as a basis for this . Equivalently, for a collection \mathcal{B} to be a basis for an existing \tau, it must hold that for every U \in \tau and every x \in U, there exists B \in \mathcal{B} with x \in B \subseteq U. A subbasis (or subbase) for a topology on X is a collection \mathcal{S} of subsets of X whose union covers X. The topology \sigma(\mathcal{S}) generated by \mathcal{S} is formed by first taking all finite intersections of elements from \mathcal{S} (which yields a basis), and then taking all arbitrary unions of those finite intersections. Formally, \sigma(\mathcal{S}) = \left\{ \bigcup_{i \in I} \left( \bigcap_{j=1}^{n_i} S_{i j} \right) \;\middle|\; I \text{ any index set}, \; n_i \in \mathbb{N}, \; S_{i j} \in \mathcal{S} \right\}, where the empty union is the empty set and the full space X arises from appropriate choices. In the standard topology on the Euclidean space \mathbb{R}^n, the collection of all open balls forms a basis. For a simple illustration, consider X = \{a, b, c, d\} with subbasis \mathcal{S} = \{\{a, b, c\}, \{b, c, d\}\}; the generated topology is \{\emptyset, \{b, c\}, \{a, b, c\}, \{b, c, d\}, X\}.

Subspaces and quotient topologies

In general topology, the provides a natural way to endow of with its own , inheriting the structure from the ambient space. Let (X, \tau) be and Y \subseteq X . The \tau_Y on Y is defined as \tau_Y = \{ U \cap Y \mid U \in \tau \}. This construction ensures that the inclusion map i: Y \hookrightarrow X is continuous, and open (respectively, closed) sets in the subspace are precisely the intersections of open (closed) sets in X with Y. Thus, subspaces inherit key topological properties such as openness and closedness relative to the ambient space, allowing for the study of induced structures on subsets without altering the original . The also interacts naturally with of the original : if \mathcal{B} is a base for \tau, then \{ B \cap Y \mid B \in \mathcal{B} \} forms a base for \tau_Y. In contrast, the constructs a topology on a set by identifying points via an equivalence relation or surjective map, often used to model gluing or collapsing in spaces. Let q: X \to Y be a surjective map from a topological space (X, \tau_X) to a set Y. The quotient topology \tau_Y on Y is the finest topology such that q is continuous, defined by \tau_Y = \{ V \subseteq Y \mid q^{-1}(V) \in \tau_X \}. Equivalently, for an equivalence relation \sim on X, the quotient space X / \sim carries the topology where a set U \subseteq X / \sim is open if the preimage under the canonical projection \pi: X \to X / \sim is open in X. This topology ensures that saturated open sets in X (those unions of equivalence classes) map to open sets in Y, and similarly for closed sets, preserving relevant openness and closedness properties under the identification. A classic example is the real projective line \mathbb{RP}^1, obtained as the quotient of the circle S^1 under the antipodal identification z \sim -z, where opposite points are glued together; this space is homeomorphic to S^1 itself but illustrates the quotient construction without relying on coordinate charts.

Examples of topological spaces

Discrete and indiscrete topologies

The discrete topology on a nonempty set X is defined as the collection of all subsets of X as open sets, making it the power set \mathcal{P}(X). This topology is the finest (largest) possible on X, containing every open set from any other topology on the same set. In the discrete topology, every singleton \{x\} for x \in X is open, and consequently, every subset of X is both open and closed. Any function f: (X, \tau_d) \to (Y, \tau_Y), where \tau_d is the discrete topology on X and \tau_Y is any topology on Y, is continuous, since the preimage f^{-1}(V) of any open V \in \tau_Y is a subset of X, hence open in \tau_d. The discrete topology satisfies the Hausdorff separation axiom, as distinct points x, y \in X can be separated by the open singletons \{x\} and \{y\}. Moreover, a is compact if and only if X is finite, because an infinite discrete space admits an open cover by singletons with no finite subcover. The indiscrete topology (also called the ) on a set X consists solely of the \emptyset and X itself as open sets, making it the coarsest (smallest) topology on X. In this topology, the only closed sets are also \emptyset and X, so no proper nonempty , including singletons, is closed unless |X| \leq 1. Any f: (Y, \tau_Y) \to (X, \tau_i), where \tau_i is the indiscrete topology on X and \tau_Y is any topology on Y, is continuous, because the preimage f^{-1}(U) of any open U \in \tau_i (either \emptyset or X) is either \emptyset or Y, both open in \tau_Y. However, a f: (X, \tau_i) \to (Y, \tau_Y) is continuous only if it is constant when Y is nontrivial (e.g., with at least two points separable by open sets), as nonconstant maps would map some open V \in \tau_Y to a preimage that is neither \emptyset nor X. The discrete topology arises naturally on finite sets in many mathematical contexts, such as when considering sets without additional structure, ensuring all subsets are distinguishable topologically. The indiscrete topology models trivial situations where no proper distinctions are made, such as the whole in certain constructions or as a baseline for comparing coarser topologies.

Cofinite and cocountable topologies

The cofinite topology on a set X is the topology whose open sets consist of the and all subsets of X whose complements are finite. Equivalently, the closed sets in this topology are the finite subsets of X and X itself. This topology coincides with the discrete topology when X is finite. For an X equipped with the , the space is T_1, as singletons are finite and hence closed. However, it is not Hausdorff, since any two nonempty have nonempty intersection: their complements are finite, so their union is finite, and thus the intersection is cofinite and nonempty given that X is infinite. The space is compact, as any open cover admits a finite subcover: select one open set U from the cover, whose complement is finite; the remaining points in the complement can then be covered by finitely many additional open sets from the cover. It is also connected: if X = U \cup V with U and V nonempty, disjoint, and open, then the complements X \setminus U = V and X \setminus V = U are both finite, implying X is finite, a contradiction. When X is uncountable with the , the space is not second-countable: a countable basis \mathcal{B} would yield a countable union of the finite complements X \setminus B for B \in \mathcal{B}, leaving points outside this union uncovered by any basis element in a way that generates all opens. If X is countable and infinite, such as the natural numbers, the cofinite topology is not Hausdorff, and sequences exhibit nonunique ; for instance, a of distinct natural numbers converges to every point in the space, as any neighborhood of a limit point excludes only finitely many elements and thus contains infinitely many terms. The cocountable topology is defined on an X as the topology whose open sets are the and all subsets whose complements are . Equivalently, the closed sets are the countable subsets of X and X itself. On a , this reduces to the discrete topology. For uncountable X with the cocountable topology, the space is T_1, since singletons are countable and hence closed. It is not Hausdorff, as any two nonempty open sets intersect: their complements are countable, so their is countable, leaving the intersection uncountable and nonempty. The space is connected—in fact, hyperconnected—because every pair of nonempty open sets has nonempty . Unlike the cofinite case, it is not compact: consider a countable infinite subset \{y_n\}_{n=1}^\infty \subset X; the open sets V_n = X \setminus \{y_1, \dots, y_n\} form a countable open cover of X with no finite subcover, as any finite V_1 \cup \cdots \cup V_m = V_m misses y_{m+1}. It is also not second-countable when X is uncountable, by an argument analogous to the cofinite case but involving countable rather than finite complements.

Topologies on real and complex numbers

The standard topology on the real numbers \mathbb{R} is the order topology induced by the usual linear order \leq, where a subbasis consists of the open rays (-\infty, a) and (a, \infty) for all a \in \mathbb{R}. A basis for this topology is given by the open intervals (a, b) = \{x \in \mathbb{R} \mid a < x < b\} for a < b. This topology makes \mathbb{R} a connected space, as it cannot be expressed as a union of two nonempty disjoint open sets, but \mathbb{R} is not compact, since the open cover \{(n, n+2) \mid n \in \mathbb{Z}\} has no finite subcover. Another notable topology on \mathbb{R} is the lower limit topology, also called the Sorgenfrey topology, generated by the basis of half-open intervals [a, b) = \{x \in \mathbb{R} \mid a \leq x < b\} for a < b. This topology is finer than the standard topology, meaning every standard open set is Sorgenfrey-open, but it includes additional open sets. The Sorgenfrey line is hereditarily Lindelöf, as every subspace has the Lindelöf property (every open cover has a countable subcover), yet it is not second-countable, since any countable collection of basis elements cannot cover all singletons in an uncountable disjoint family. The complex numbers \mathbb{C} carry the standard topology by identifying \mathbb{C} with \mathbb{R}^2 via the map z = x + iy \mapsto (x, y), endowing it with the product topology (or equivalently, the Euclidean topology on \mathbb{R}^2). Open sets are unions of open disks B_r(\zeta) = \{z \in \mathbb{C} \mid |z - \zeta| < r\} for \zeta \in \mathbb{C} and r > 0. Like \mathbb{R}, \mathbb{C} is connected but not compact in this topology. A non-Hausdorff topology on \mathbb{C}, viewed as the affine line \mathbb{A}^1(\mathbb{C}), is the , where closed sets are the whole space \mathbb{C} or finite subsets (zeros of nonzero polynomials in \mathbb{C}). Since any two nonempty open sets (complements of finite sets) intersect, the space fails to separate distinct points with disjoint open neighborhoods. As a non-Hausdorff variant on a line-like space, consider the double-pointed line (or line with two origins), constructed by taking two copies of \mathbb{R} and identifying all points except the origins, resulting in a space homeomorphic to \mathbb{R} away from the two distinct origin points o_1 and o_2. Basic open sets around points other than the origins are standard intervals, but neighborhoods of o_1 and o_2 cannot be disjoint while containing each, violating the Hausdorff axiom, though the space is otherwise locally Euclidean.

Continuous functions

Primary definitions

In general topology, a function f: X \to Y between topological spaces (X, \mathcal{T}_X) and (Y, \mathcal{T}_Y) is defined to be continuous if the preimage f^{-1}(V) of every open set V \in \mathcal{T}_Y is an open set in X. This open-set definition generalizes the intuitive notion of continuity from metric spaces, where it corresponds to the \epsilon-\delta condition, but applies to arbitrary topological spaces without relying on distances. An equivalent characterization of continuity is that the preimage f^{-1}(C) of every closed set C in Y is closed in X. Another equivalent condition is that the graph of f, defined as \Gamma_f = \{(x, f(x)) \mid x \in X\} \subseteq X \times Y equipped with the , is a closed subset of X \times Y. Related to are the concepts of . A function f: X \to Y is an open map if the f(U) of every open set U \in \mathcal{T}_X is open in Y, and it is a closed map if the f(C) of every closed set C in X is closed in Y. While continuous functions need not be open or closed, these properties highlight how f interacts with the topological structures of X and Y. Continuous functions preserve certain closure properties: for any subset A \subseteq X, the image of the closure satisfies f(\overline{A}) \subseteq \overline{f(A)}, where \overline{A} denotes the of A in X and \overline{f(A)} the of f(A) in Y. This inclusion reflects how continuity ensures that limits and accumulations in the map into accumulations in the .

Alternative characterizations

In general topological spaces, continuity of a function f: X \to Y at a point x \in X can be equivalently characterized using neighborhoods: for every neighborhood V of f(x) in Y, there exists a neighborhood U of x in X such that f(U) \subseteq V. This formulation emphasizes local preservation of nearness and generalizes the epsilon-delta condition from metric spaces without relying directly on inverse images of open sets. Another characterization employs sequences, particularly in spaces with sufficient countability. A function f: X \to Y is sequentially continuous at x \in X if, whenever a \{x_n\} in X converges to x, the \{f(x_n)\} in Y converges to f(x). In any , continuity implies sequential continuity, but the converse holds if and only if X is first-countable, meaning every point has a countable local basis. For example, in metric spaces, which are first-countable, these notions coincide, allowing sequential limits to fully capture topological . To extend this to arbitrary topological spaces, nets provide a generalization of sequences. A net in X is a function from a directed set to X, and it converges to x \in X if every neighborhood of x eventually contains all net values beyond some index. A function f: X \to Y is continuous at x if and only if, for every net \{x_\alpha\} in X converging to x, the net \{f(x_\alpha)\} in Y converges to f(x). This characterization works universally, as nets detect the topology in non-first-countable spaces where sequences may fail, such as the product topology on uncountable products. Filters offer yet another equivalent perspective, generalizing both sequences and nets through ultrafilter-like structures. A filter on X converges to x \in X if every neighborhood of x belongs to the . The f: X \to Y is continuous at x , for every \mathcal{F} on X converging to x, the image f(\mathcal{F}) = \{f(A) \mid A \in \mathcal{F}\} converges to f(x) in Y. This approach is particularly useful in spaces where needs to be defined without ordering, and it aligns with the neighborhood at x, which always converges to x. Finally, continuity can be expressed using closure operators. The closure \mathrm{cl}_X(A) of a A \subseteq X is the smallest containing A. The f: X \to Y is if and only if, for every A \subseteq X, f(\mathrm{cl}_X(A)) \subseteq \mathrm{cl}_Y(f(A)). This condition reflects how continuous maps preserve limits of sets, ensuring that images of adherent points remain adherent, and it holds globally without reference to points or open sets.

Homeomorphisms

A homeomorphism between two topological spaces X and Y is a bijective continuous map f: X \to Y whose inverse f^{-1}: Y \to X is also continuous. This condition ensures that f preserves the topological structure of open sets, meaning f maps open sets in X to open sets in Y, and equivalently for f^{-1}. Two spaces are said to be homeomorphic, denoted X \cong Y, if such a map exists, establishing topological equivalence where the spaces are indistinguishable by any topological property. Examples of homeomorphisms abound in familiar spaces. On smooth manifolds, a diffeomorphism—a bijective smooth map with a smooth inverse—is necessarily a homeomorphism, as smoothness implies continuity. A concrete instance is the homeomorphism between the real line \mathbb{R} and the open interval (-\pi/2, \pi/2), given by the tangent function \tan: (-\pi/2, \pi/2) \to \mathbb{R}, which is bijective, continuous, and has a continuous inverse \arctan: \mathbb{R} \to (-\pi/2, \pi/2). This extends to any open interval (a, b) via a scaled and translated version, such as f(x) = \tan\left(\pi \frac{x - (a+b)/2}{b-a}\right), confirming that all bounded open intervals on \mathbb{R} are homeomorphic to the entire line. Homeomorphisms preserve topological invariants—properties invariant under such maps—allowing classification of spaces up to topological equivalence. Key preserved properties include compactness, as the continuous image of a compact set is compact and bijectivity ensures the inverse behaves similarly; connectedness, where the image of a connected remains connected; and separation axioms like Hausdorffness. In contrast, properties beyond topology, such as differentiability or distances, are not preserved; for instance, the tangent map above is not differentiable at the endpoints in an extended sense, despite being a . Homeomorphisms facilitate construction via gluing: if compatible open covers of spaces yield homeomorphic local pieces, the global glued space inherits the homeomorphism. Specifically, the pasting lemma ensures that continuous bijections on overlapping open sets combine to a homeomorphism on the union, as seen in gluing charts for manifolds like the Grassmannian, where local Euclidean pieces are homeomorphic under compatible transitions. Thus, if two spaces are homeomorphic and glued along homeomorphic subsets via compatible maps, the resulting spaces are homeomorphic.

Core topological properties

Compactness

In general topology, a topological space X is defined to be compact if every open cover of X has a finite subcover. This property captures a form of "finiteness" in infinite spaces, generalizing the behavior of finite sets where any cover trivially admits a finite subcover. is a topological , preserved under homeomorphisms, and serves as a foundational concept for many theorems in and . In metric spaces, compactness implies that the space is both closed and bounded, though the converse does not hold in general. Specifically, the Heine-Borel theorem establishes that for subsets of \mathbb{R}^n with the standard topology, a set is it is closed and bounded. However, in arbitrary Hausdorff spaces or non-metric topologies, closed and bounded sets need not be compact; counterexamples include certain infinite-dimensional normed spaces where bounded closed balls fail to be compact. Compact sets in Hausdorff spaces are always closed, as the complement of a compact set is open. A key property of compact spaces is that continuous images of compact spaces are compact. If f: X \to Y is a and X is compact, then f(X) is compact in Y. This preservation under continuous maps underscores compactness's role in ensuring extremal values, such as the for continuous functions on compact subsets of \mathbb{R}^n. Sequential compactness provides an equivalent characterization in : a is compact every in the space has a convergent . In general topological spaces, sequential compactness (every sequence has a convergent subsequence) implies compactness but not conversely, as there exist compact spaces without this sequential property, such as certain uncountable products. Local compactness is a related but weaker notion: a topological space is locally compact if every point has a neighborhood basis consisting of compact sets. Every compact space is locally compact, but the converse fails; for example, the real line \mathbb{R} is locally compact but not compact overall. Locally compact Hausdorff spaces admit useful one-point compactifications, extending to non-compact cases like \mathbb{R}^n. Tychonoff's theorem states that the product of any collection of compact topological spaces, equipped with the product topology, is compact. This result, relying on the axiom of choice, enables the compactness of infinite products like the Hilbert cube [0,1]^\mathbb{N}, with applications in functional analysis.

Connectedness

A topological space X is said to be connected if it cannot be expressed as the union of two nonempty disjoint open subsets whose union is X. Equivalently, X is connected if the only clopen subsets (subsets that are both open and closed) of X are the and X itself. This property captures the intuitive notion of a space being "in one piece," preventing it from being split into separated parts by the topology. The connected components of a topological space X are the maximal connected subsets of X. For each point x \in X, the containing x, denoted C(x), is the largest connected subset of X that includes x, and it can be constructed as the intersection of all connected subsets containing x. These connected components form a of X, meaning every point belongs to exactly one component, and distinct components are disjoint. A space X is locally connected at a point x if every neighborhood of x contains a connected neighborhood of x, and X is locally connected if it is locally connected at every point; in such spaces, the connected components are open subsets. A stronger notion is path-connectedness: a topological space X is path-connected if, for any two points x, y \in X, there exists a continuous function \gamma: [0,1] \to X such that \gamma(0) = x and \gamma(1) = y. Path-connectedness implies connectedness, since if X = U \cup V were a separation into nonempty disjoint open sets, a path from a point in U to a point in V would have to jump discontinuously between them, contradicting continuity. However, the converse does not hold; the topologist's sine curve provides a classic counterexample. Define S = \{(x, \sin(1/x)) \mid 0 < x \leq 1\} \cup \{(0, y) \mid -1 \leq y \leq 1\} with the subspace topology from \mathbb{R}^2. This space S is connected because its only connected subsets are either contained in the vertical segment at x=0 or include points from the sine curve, preventing a separation, but it is not path-connected since no continuous path can connect a point on the vertical segment (away from the origin) to a point on the oscillating curve without violating the uniform continuity bound on paths. Examples of disconnected spaces abound. The set of rational numbers \mathbb{Q}, endowed with the subspace topology from \mathbb{R}, is disconnected; for instance, it can be separated into \mathbb{Q} \cap (-\infty, \sqrt{2}) and \mathbb{Q} \cap (\sqrt{2}, \infty), both nonempty and open in \mathbb{Q}. In fact, \mathbb{Q} is totally disconnected, meaning its only connected subsets are singletons, as any two distinct rationals can be separated by open intervals avoiding irrationals between them. Discrete spaces, where every subset is open, are also totally disconnected, with each singleton as a component. Connectedness is preserved under continuous maps: the continuous image of a connected space is connected. For example, the real line \mathbb{R} with the standard topology is connected, as any separation would contradict the intermediate value theorem for continuous functions on intervals. Variants like the topologist's sine curve illustrate spaces that are connected but lack path-connectedness, highlighting the subtlety between these notions.

Separation and countability

Separation axioms

Separation axioms form a hierarchy of conditions on topological spaces that quantify the extent to which distinct points or closed sets can be distinguished using open neighborhoods. These properties, developed primarily in the early 20th century, are essential for ensuring well-behaved convergence, continuity, and embedding into metric spaces. The axioms range from weak separation, where points are minimally distinguishable, to stronger ones that allow separation of closed sets, facilitating applications in analysis and geometry. The T0 axiom, also called the Kolmogorov axiom after Andrey Kolmogorov's 1937 work on topological invariants, requires that for any two distinct points x, y in a space X, there exists an open set containing one but not the other. This ensures points are topologically distinguishable in at least one direction, though not symmetrically. Examples include the Zariski topology on algebraic varieties, where non-T0 spaces can arise but are often quotiented to T0 versions for uniqueness. A T1 space, named after Maurice Fréchet's 1906 thesis on functional calculus, strengthens T0 by requiring that singletons are closed sets, or equivalently, that every pair of distinct points has open neighborhoods excluding the other. In T1 spaces, finite sets are closed, and limits of sequences, if they exist, are unique. The cofinite topology on an infinite set exemplifies a T1 space: open sets are those with finite complements, making singletons closed but failing stronger separation. The T2 axiom, or Hausdorff axiom from Felix Hausdorff's 1914 foundational text on set theory, demands that distinct points possess disjoint open neighborhoods. This symmetric separation implies unique sequential limits and is standard in most practical topologies, such as Euclidean spaces. Non-Hausdorff examples like the cofinite topology highlight pathologies, such as non-unique limits, underscoring T2's importance for analysis. T3 spaces, often defined as T1 plus regular (where points and disjoint closed sets have disjoint open neighborhoods), allow separation of points from closed sets not containing them. Regularity ensures closed sets behave well under continuous functions. Some conventions merge T3 with T0 instead of T1, but the T1 version is common for Hausdorff-like progression. The strongest common axiom, T4 or normal, combines T1 with the ability to separate any two disjoint closed sets by disjoint open sets. Normality enables the existence of continuous functions separating closed sets, as in . A key implication is : a second-countable T3 space is metrizable, embedding it into a metric space while preserving topology. This bridges abstract topology to metric analysis, with examples like manifolds benefiting from such structure. Weaker axioms like T0 appear in algebraic geometry, while stronger ones like T4 dominate in differential topology; enhancements with countability axioms yield even finer classifications.

Countability axioms

In general topology, countability axioms impose restrictions on the "size" of the topology by requiring certain countable structures, such as bases or dense subsets, which facilitate the use of sequences and countable covers in proofs of continuity and compactness-like properties. These axioms are particularly useful in distinguishing metrizable spaces and ensuring that abstract topological concepts behave similarly to those in familiar Euclidean spaces. A topological space X is first-countable if, for each point x \in X, there exists a countable local basis \{B_n(x)\}_{n \in \mathbb{N}} consisting of neighborhoods of x such that every neighborhood of x contains some B_n(x). In such spaces, sequences suffice to characterize continuity and limits: a function f: X \to Y is continuous at x if and only if, for every sequence \{x_n\} in X converging to x, the sequence \{f(x_n)\} converges to f(x) in Y. First-countability holds in all metrizable spaces, as the balls of rational radii around each point form a countable local basis. A topological space X is second-countable if it admits a countable basis \mathcal{B} = \{U_n\}_{n \in \mathbb{N}} for its topology, meaning every open set in X is a union of elements from \mathcal{B}. Every second-countable space is first-countable, as the collection of basis elements containing a fixed point x provides a countable local basis at x. The real line \mathbb{R} with the standard topology is second-countable, with the open intervals having rational endpoints forming a countable basis. Subspaces and countable products of second-countable spaces remain second-countable. A topological space X is separable if it contains a countable dense subset D \subseteq X, meaning the closure of D is all of X. Second-countable spaces are necessarily separable: given a countable basis \mathcal{B}, select one point from each non-empty basis element to form a countable dense set (using the axiom of countable choice). However, separability does not imply second-countability in general; for example, the lower limit topology on \mathbb{R} (also known as the ) is separable but not second-countable. An uncountable discrete space, where every subset is open, cannot be separable, as any dense subset would need to intersect every singleton and thus be uncountable. A topological space X is Lindelöf if every open cover of X admits a countable subcover. Second-countable spaces are Lindelöf, since any open cover can be refined to the countable basis, and the basis elements used in the unions suffice as a countable subcover. The product of two is not Lindelöf, despite each factor being Lindelöf, illustrating that the property is not preserved under arbitrary products. First-countable spaces need not be Lindelöf, but in , separability, Lindelöf, and second-countability are equivalent. The implications among these axioms form a hierarchy: second-countable implies first-countable (and thus sequential, where the topology is determined by sequential convergence), which in turn implies properties like the sequential characterization of closed sets, but none of these fully imply separability or Lindelöf without additional assumptions. These countability conditions interact with separation axioms by ensuring that points can be distinguished using countable structures, though they primarily address covering and density rather than point separation. A topological space X satisfies the countable chain condition (ccc) if every collection of pairwise disjoint non-empty open sets in X is at most countable. Separable spaces satisfy the ccc, as a countable dense set intersects every non-empty open set, limiting the size of disjoint families. However, the ccc does not imply separability; for instance, the Cantor cube \{0,1\}^\kappa for uncountable \kappa with the product topology has the ccc but is not separable. The ccc is a cardinal restriction on the topology, often used in studying compactness and covering properties in non-metrizable spaces.

Metric and uniform structures

Metric spaces

A metric space is a pair (X, d), where X is a set and d: X \times X \to [0, \infty) is a function, called a metric, satisfying the following axioms for all x, y, z \in X:
  • d(x, y) = 0 if and only if x = y (identity of indiscernibles),
  • d(x, y) = d(y, x) (symmetry),
  • d(x, z) \leq d(x, y) + d(y, z) (triangle inequality).
This structure was first axiomatized by in his 1906 doctoral thesis to unify notions of convergence in function spaces. The metric d induces a topology on X by declaring sets of the form B(x, r) = \{ y \in X \mid d(x, y) < r \} for x \in X and r > 0 to be open balls, which form a basis for the . In this , a set U \subseteq X is open if for every x \in U there exists r > 0 such that B(x, r) \subseteq U. Two metrics d and d' on X are equivalent if they induce the same , meaning the collections of open balls generate the same open sets; equivalence holds if for every x \in X and r > 0, there exist r', s, s' > 0 such that B'(x, r') \subseteq B(x, r) \subseteq B'(x, s) and B(x, s) \subseteq B'(x, s'). Every is Hausdorff: if x \neq y, then d(x, y) > 0, so the open balls B(x, d(x, y)/2) and B(y, d(x, y)/2) are disjoint and contain x and y, respectively. Moreover, are first-countable: at each point x, the countable collection \{ B(x, 1/n) \mid n \in \mathbb{N} \} forms a local basis. A sequence (x_n) in a (X, d) is Cauchy if for every \epsilon > 0 there exists N \in \mathbb{N} such that d(x_m, x_n) < \epsilon for all m, n > N. The space is complete if every converges to some point in X. Completeness is not a , as it depends on the specific metric rather than the induced topology; for example, the rational numbers \mathbb{Q} with the standard metric d(p, q) = |p - q| form an incomplete homeomorphic to a dense of the \mathbb{R} with the same metric. The Urysohn metrization theorem states that a is metrizable if and only if it is , Hausdorff, and second-countable. To sketch the proof of sufficiency, let \{ U_n \mid n \in \mathbb{N} \} be a countable basis for the second-countable Hausdorff X. Using regularity, continuous functions f_{ij}: X \to [0,1] (Urysohn functions) can be constructed such that f_{ij}^{-1}(0) \subseteq U_i and f_{ij}^{-1}(1) \supseteq X \setminus U_j for pairs where \overline{U_i} \subseteq U_j. The metric d(x, y) = \sum_{k=1}^\infty 2^{-k} |f_k(x) - f_k(y)|, where \{f_k\} enumerates the f_{ij}, induces the original topology on X. Common examples include the Euclidean metric on \mathbb{R}^n, defined by d(x, y) = \sqrt{\sum_{i=1}^n (x_i - y_i)^2}, which induces the standard topology; the taxicab (Manhattan) metric d(x, y) = \sum_{i=1}^n |x_i - y_i|, equivalent to the Euclidean metric and thus inducing the same topology; and the discrete metric d(x, y) = 1 if x \neq y and $0 otherwise, which induces the discrete topology where every subset is open.

Uniform spaces

Uniform spaces provide a generalization of metric spaces that captures the notion of and without requiring a numerical function, allowing for more abstract structures in . Introduced by in 1937, they consist of a set equipped with a collection of relations called entourages that axiomatize nearness in a way compatible with the . This framework is particularly useful for studying properties like Cauchy sequences and completions in non-metric settings, such as topological groups or product spaces. A uniform structure on a set X is a \mathcal{U} on the product set X \times X, where the elements of \mathcal{U} are subsets called , satisfying three axioms: reflexivity, ensuring that the diagonal \Delta_X = \{(x,x) \mid x \in X\} is contained in every entourage; symmetry, requiring that if E \in \mathcal{U}, then its opposite E^{op} = \{(y,x) \mid (x,y) \in E\} is also in \mathcal{U}; and the , stating that for every E \in \mathcal{U}, there exists E' \in \mathcal{U} such that the composition E' \circ E' \subseteq E, where E_1 \circ E_2 = \{(x,z) \mid \exists y \in X \text{ s.t. } (x,y) \in E_1, (y,z) \in E_2\}. The pair (X, \mathcal{U}) is then called a . The uniformity \mathcal{U} induces a on X, known as the uniform topology, where a set U \subseteq X is open if for every x \in U, there exists an entourage E \in \mathcal{U} such that the slice E_x = \{y \in X \mid (x,y) \in E\} \subseteq U. Equivalently, a point x belongs to the of a A \subseteq X every entourage E \in \mathcal{U} satisfies E_x \cap A \neq \emptyset. Uniform continuity between uniform spaces (X, \mathcal{U}) and (Y, \mathcal{V}) is defined for a f: X \to Y such that for every E \in \mathcal{V}, the preimage (f \times f)^{-1}(E) \in \mathcal{U}. This notion extends the version, where small distances in the domain imply small distances in the uniformly across the space, without dependence on specific points. A \mathcal{F} on X is Cauchy with respect to \mathcal{U} if for every E \in \mathcal{U}, there exists F_0 \in \mathcal{F} such that F_0 \times F_0 \subseteq E; the is complete if every Cauchy converges in the induced . Every d on X induces a structure \mathcal{U}_d with a basis of entourages E_\epsilon = \{(x,y) \in X \times X \mid d(x,y) < \epsilon\} for \epsilon > 0, which satisfies the uniformity axioms. Conversely, uniform spaces abstract this construction, as every uniformity admits a compatible family of pseudometrics generating it. Examples include the indiscrete uniformity on any set X, generated by all subsets of X \times X containing \Delta_X, which induces the indiscrete ; and the product uniformity on \prod_{i \in I} X_i, where the entourages are those whose projections belong to the respective \mathcal{U}_i for all but finitely many i, inducing the .

Advanced theorems and concepts

Baire category theorem

The is a fundamental result in general topology that highlights the "largeness" of certain topological spaces in terms of category. In its primary form, it asserts that a is a , meaning that the intersection of any countable collection of dense open subsets is itself dense. Equivalently, a cannot be expressed as a countable union of nowhere dense sets; such unions are termed meager or of the first category. A is nowhere dense if its has empty interior, and meager sets thus represent "small" sets in this categorical sense. This theorem underscores that are of the second category in themselves, preventing them from being "small" in this regard. A sketch of the proof for complete metric spaces proceeds by contradiction or direct construction. To show the intersection \bigcap_{n=1}^\infty G_n is dense, where each G_n is dense and open in the complete space X, fix a nonempty open set [V](/page/V.) \subseteq X. Inductively construct a sequence of nonempty open sets U_n \subseteq G_n \cap V such that \overline{U_{n}} \supseteq U_{n+1} and \operatorname{diam}(U_n) < 1/n. The centers of these sets form a , which converges by completeness to a point in \bigcap_{n=1}^\infty G_n \cap V, hence the intersection is dense. For the locally compact Hausdorff version, the theorem holds similarly: such spaces are Baire spaces. The proof adapts the construction by selecting nested open sets [V_n](/page/V.) \subseteq G_n with compact closures \overline{V_n} satisfying \overline{V_{n+1}} \subseteq V_n and \overline{V_1} \subseteq V; the finite intersection property of these compacts ensures a nonempty intersection point in \bigcap_{n=1}^\infty G_n \cap V. Key corollaries illustrate the theorem's implications. The real line \mathbb{R}, as a complete metric space, cannot be a countable union of singletons, since each singleton is nowhere dense and their countable union would be meager. The set of irrational numbers \mathbb{R} \setminus \mathbb{Q} is comeager (residual), being the countable intersection of dense open sets (complements of enumerated rationals), while \mathbb{Q} is meager. In function spaces, such as the Banach space C[0,1] of continuous functions on [0,1] with the sup norm (which is complete metric), the set of nowhere differentiable functions is comeager, as the set of functions differentiable at some point is meager via unions over rationals and suitable open dense complements. The Banach fixed-point theorem, guaranteeing a unique fixed point for contractions on complete metric spaces, relies on completeness akin to the Baire theorem's foundation, though it uses successive approximations directly. Meager sets contrast with comeager (residual) sets, the latter being complements of meager sets and thus "large" in Baire spaces, containing dense intersections of open dense sets. Applications to function spaces abound, such as showing that generic continuous functions on [0,1] fail to be monotone on any interval or exhibit other pathological behaviors, leveraging the Baire property to identify residual subsets. Generalizations extend the theorem to uniform spaces, where completeness is replaced by uniform completeness ensuring Cauchy filters converge, preserving the Baire space property. Further, Čech-complete spaces—those completable to a complete uniform space in the Čech uniformity—are Baire spaces, providing a topological analogue of metric completeness without a metric.

Filters and convergence

A filter on a set X is a nonempty collection \mathcal{F} of subsets of X such that \emptyset \notin \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}. Typically, X \in \mathcal{F}. Introduced by in 1937 to generalize sequences for in arbitrary topological spaces, filters enable the definition of limits without countable indexing, which is vital in spaces lacking a countable local basis. An ultrafilter is a maximal filter \mathcal{U} on X, meaning that for every A \subseteq X, exactly one of A or X \setminus A belongs to \mathcal{U}. Principal ultrafilters are those generated by a fixed point p \in X, consisting of all supersets of \{p\}; they converge precisely to p. On infinite sets, free (non-principal) ultrafilters exist and contain no finite sets, providing a notion of "largeness" for cofinite or more general subsets. In a (X, \tau), a \mathcal{F} on X to a point x \in X (denoted \mathcal{F} \to x) if every open neighborhood of x belongs to \mathcal{F}, or equivalently, if the neighborhood \mathcal{N}_x = \{U \in \tau : x \in U\} is coarser than \mathcal{F}. This generalizes sequential : for a (x_n) in a first-countable space, the associated is the Fréchet filter on \mathbb{N}, comprising all cofinite subsets of \mathbb{N}, and \mathcal{F} \to x x_n \to x. The ultrafilter lemma asserts that every filter on X extends to an ultrafilter, a result that follows from the via but is strictly weaker than the full . This lemma underpins proofs of in general topology: the product of compact Hausdorff spaces is compact, as an ultrafilter on the product converges to a unique point if its projections converge in each factor. In sheaf theory over a topological space X, the stalk \mathcal{F}_x of a sheaf of sets \mathcal{F} at x \in X is the colimit of \mathcal{F}(U) over open neighborhoods U of x, consisting of germs—equivalence classes of sections s \in \mathcal{F}(U) where (U, s) \sim (V, t) if s and t agree on some W \subseteq U \cap V containing x. Filters facilitate the construction of these local "germs" by capturing the refinement of neighborhoods. Filters relate to nets by generating a cofinal directed set from their partial by inclusion, yielding a whose in X is equivalent to that of the ; conversely, every defines a convergent .

Research areas

Continuum and dimension theory

In continuum theory, a central object of study is the , defined as a compact connected . These spaces capture the intuitive notion of a continuous "filled" region without gaps, and their topological properties, such as arc connectedness (where every pair of points can be joined by an ), are fundamental to understanding embeddings and mappings in . A particularly important subclass consists of Peano continua, which are locally connected continua. These spaces admit continuous surjections from the unit interval [0,1], as established by the Hahn-Mazurkiewicz theorem, which characterizes them as precisely the compact, connected, locally connected, and metrizable spaces that are images of [0,1] under continuous maps. Among Peano continua, the —defined as the product \prod_{n=1}^\infty [0,1]—serves as a universal space, meaning every Peano continuum embeds homeomorphically into it. This universality underscores the Hilbert cube's role as an infinite-dimensional analogue of the unit interval, facilitating the study of embeddings and approximations in theory. Dimension theory in general topology extends these ideas by providing invariants to classify spaces beyond mere connectedness. The small inductive dimension, introduced by Menger and Urysohn, is defined recursively: \operatorname{ind}(X) = -1 if X is empty, and \operatorname{ind}(X) = n if every point in X has arbitrarily small neighborhoods whose boundaries have inductive dimension at most n-1, with the dimension being the supremum over such values. Formally, \operatorname{ind}(X) \leq n if for every x \in X and every open neighborhood U of x, there exists an open V \subseteq U such that \operatorname{ind}(\partial V) \leq n-1. This definition aligns with intuitive notions, yielding \operatorname{ind}(\mathbb{R}^k) = k. Complementing the inductive approach is the , which measures the minimal order of open covers. A X has covering dimension at most n if every finite open cover admits a refinement where no point lies in more than n+1 sets, and the dimension is the smallest such n. This notion, originating from Lebesgue's work on spaces, coincides with the inductive dimension for compact metric spaces and provides a combinatorial tool for analyzing coverings in continua. A cornerstone result in dimension theory is Brouwer's invariance of dimension theorem, which asserts that Euclidean spaces \mathbb{R}^n and \mathbb{R}^m are not homeomorphic if n \neq m. Proved by Brouwer in 1911 using fixed-point arguments and homology precursors, this theorem establishes as a topological , preventing "dimension collapse" under homeomorphisms and influencing the of continua.

Set-theoretic and algebraic topology

In set-theoretic topology, the weight of a topological space, denoted w(X), is defined as the minimal of a for the on X. This measures the "" of the space's open sets and plays a crucial role in classifying spaces under axioms like the (CH). Similarly, the density character d(X) is the smallest of a dense of X, providing insight into the minimal size needed to approximate the space densely. These interact with separation properties; for instance, under CH, there exist Moore spaces that are not metrizable, highlighting how set-theoretic assumptions can produce pathological that fail higher separation axioms like complete normality. Normality in topological spaces, the T_4 requiring disjoint closed sets to be separated by disjoint open sets, exhibits sensitivity to set-theoretic forcing and axioms beyond ZFC. (MA) combined with the negation of implies that certain paracompact spaces are but fail to be collectionwise , whereas can force the existence of non- spaces with countable dense subsets. These results underscore the independence of from ZFC, with often yielding counterexamples to conjectures like the normality of product spaces under specific conditions. Topological groups extend the structure of groups by equipping them with a topology where the group operations— and inversion—are continuous. A G thus combines with topological , enabling the study of and uniformity in group actions. For locally compact abelian s, establishes a between G and the dual group \hat{G} of continuous homomorphisms from G to the circle group \mathbb{T}, with the double dual \hat{\hat{G}} recovering G topologically. This duality, originally developed by in 1934, transforms problems in , such as the structure of compact abelian groups, into algebraic questions about discrete modules. Topological vector spaces generalize normed spaces by imposing a topology on a over \mathbb{R} or \mathbb{C} such that vector addition and are continuous. In these spaces, linear functionals are required to be continuous to preserve topological properties, distinguishing them from purely algebraic linear maps. The Hahn-Banach theorem, in its normed space version, guarantees that any continuous linear functional defined on a of a normed vector space extends to a continuous linear functional on the entire space while preserving the norm bound, a result pivotal for duality theory in Banach spaces. This extension property underpins separation theorems and the existence of dual spaces in functional analysis. Point-free topology, also known as theory, reformulates topological spaces without relying on points, instead using s—complete Heyting algebras of "open sets" closed under arbitrary joins and finite meets—to represent generalized spaces. A is the dual of a , where morphisms correspond to frame homomorphisms preserving the lattice operations, allowing the study of and intrinsically via algebraic means. This approach generalizes classical to settings where points may not exist or are pathological, such as in constructive , and frames serve as point-free analogs of sober spaces. In , pointless topology finds applications in and synthetic topology, where locales model computational processes and data types without explicit points, facilitating proofs in and . For example, frames represent domains of partial information, enabling the construction of continuous functions in programming semantics. Recent work post-2020 has explored locales in for synthetic differential geometry, bridging point-free methods with computational verification. The Bing metrization theorem characterizes metrizable spaces as those that are collectionwise normal and possess a basis, where the basis is a countable of discrete families of open sets. Independently, the Nagata-Smirnov metrization theorem states that a topological space is metrizable if and only if it is and has a \sigma-locally finite basis, meaning the basis decomposes into countably many locally finite collections. These theorems provide uniform criteria for spaces into structures, resolving the classical metrization problem for non-second-countable spaces via refinements of Urysohn's earlier result.

References

  1. [1]
    [PDF] Chapter 4: Topological Spaces - UC Davis Math
    Definition 4.1 A topology on a nonempty set X is a collection of subsets of X, called open sets, such that: (a) the empty set 0 and the set X are open;. (b) the ...
  2. [2]
    [PDF] A Review of General Topology. Part 1: First Concepts - CSUSM
    This document gives the first definitions and results of general topology. It is the first document of a series which reviews the basics of general topology ...
  3. [3]
    [PDF] definitions and theorems in general topology - UTK Math
    1. Basic definitions. A topology on a set X is defined by a family O of subsets of X, the open sets of the topology, satisfying the axioms: (i) ∅ and X are in ...
  4. [4]
    [PDF] Part 1. Prelude to Topology
    Topology took shape in the 19th century, with isolated results before. Euler's work on the Seven Bridges of Königsberg was an early topological statement.
  5. [5]
    [PDF] Topology - Harvard Mathematics Department
    1 Introduction. Topology is simply geometry rendered flexible. In geometry and analysis, we have the notion of a metric space, with distances specified between ...
  6. [6]
    [PDF] An outline summary of basic point set topology - UChicago Math
    Topological spaces. Definition 1.1. A topology on a set X is a set of subsets, called the open sets, which satisfies the following conditions.
  7. [7]
    Topological Spaces - Department of Mathematics at UTSA
    Oct 29, 2021 · A topological space is a set of points, along with a set of neighbourhoods for each point, satisfying a set of axioms relating points and neighbourhoods.
  8. [8]
    [PDF] 1. Introduction
    Broadly speaking, a topology on a given set X is simply a specified collection of subsets of X, which we will call open sets, that satisfies certain properties.
  9. [9]
    [PDF] Introduction to Topology - Cornell Math Department
    Sep 16, 2010 · Page 2. 1 Topological spaces. A topology is a geometric structure defined on a set. Basically it is given by declaring which subsets are “open” ...
  10. [10]
  11. [11]
    A history of Topology - MacTutor - University of St Andrews
    Cantor in 1872 introduced the concept of the first derived set, or set of limit points, of a set. He also defined closed subsets of the real line as subsets ...
  12. [12]
    [PDF] Papers on Topology - School of Mathematics
    Jul 31, 2009 · ... Poincaré before topology. In the introduction to his first major topology paper, the Analysis situs, Poincaré. (1895) announced his goal of ...
  13. [13]
    [PDF] Sur quelques points du calcul fonctionnel
    C'est ce que nous allons essayer de faire pour le Calcul Fonctionnel et en par- ticulier pour la th6orie des ensembles abstraits. 7- Si l'on examine avec soin ...
  14. [14]
    Grundzüge der Mengenlehre : Hausdorff, Felix, 1868-1942
    Dec 2, 2008 · Grundzüge der Mengenlehre. by: Hausdorff, Felix, 1868-1942. Publication date: 1914. Topics: Set theory. Publisher: Leipzig Viet. Collection ...Missing: separation axioms neighborhoods
  15. [15]
    Kazimierz Kuratowski (1896 - 1980) - Biography - MacTutor
    Other major contributions by Kuratowski were to compactness and metric spaces. He was the author of Topologie, referred to above, which was the crowning ...
  16. [16]
    Pavel Urysohn (1898 - 1924) - Biography - MacTutor
    The first question that Egorov posed was to find a general intrinsic topological definition of a curve which when restricted to the plane became Cantor's notion ...Missing: Alexandrov neighborhood 1920s
  17. [17]
    Kolmogorov's ideas in the theory of operations on sets - IOPscience
    Introduction to the theory of operations on sets. § 2. The Kolmogorov hierarchy. § 3. The R-transform. § 4. Inductive analysis. Indices.
  18. [18]
    Roman Sikorski - Biography - MacTutor - University of St Andrews
    Roman Sikorski was a Polish mathematician whose research interests included Boolean algebras, mathematical logic, functional analysis, the theory of ...Missing: Borel 1930s
  19. [19]
    A Topology for Spaces of Transformations - jstor
    3, July, 1946. A TOPOLOGY FOR SPACES OF TRANSFORMATIONS'. By RICHARD F ... facts about metric spaces to facts about uniform spaces. In this connection ...
  20. [20]
    Sur l'opération Ā de l'Analysis Situs - EuDML
    Kuratowski, Casimir. "Sur l'opération Ā de l'Analysis Situs." Fundamenta Mathematicae 3.1 (1922): 182-199. <http://eudml.org/doc/213290>.
  21. [21]
    [PDF] 4 | Basis, Subbasis, Subspace
    It can difficult to describe all open sets explicitly, so topological spaces are often defined by giving either a basis or a subbasis of a topology.
  22. [22]
    [PDF] 7. Subspaces
    Definition 2.1. Let (X,Τ ) be a topological space, and let Y Ç X be any subset. We define the subspace topology ΤY on Y (we will sometimes write Τsubspace for ...
  23. [23]
    [PDF] Introduction to Topology - Cornell Math Department
    Sep 23, 2010 · A subset of a topological space has a naturally induced topology, called the subspace topology. In geometry, the subspace topology is the source ...
  24. [24]
    [PDF] 19 | Quotient Spaces
    19.11 Definition. Let X be a topological space and let ∼ be an equivalence relation on X. The quotient topology on the set X/∼ is the topology where a set ...<|control11|><|separator|>
  25. [25]
    [PDF] Math 190: Quotient Topology Supplement
    The quotient topology is one of the most ubiquitous constructions in algebraic, combinatorial, and differential topology. It is also among the most difficult ...
  26. [26]
    [PDF] Real Projective Space: An Abstract Manifold
    Mar 10, 2017 · We can consider three types of sets in the quotient space. The first consists of regions of the circle away from (1,0), which map back to open ...
  27. [27]
    Discrete Topology -- from Wolfram MathWorld
    The largest topology contains all subsets as open sets, and is called the discrete topology. In particular, every point in X is an open set in the discrete ...
  28. [28]
    Trivial Topology -- from Wolfram MathWorld
    Also called indiscrete topology, the trivial topology is the smallest topology on a set , namely the one in which the only open sets are the empty set and the ...Missing: definition | Show results with:definition<|control11|><|separator|>
  29. [29]
    Example of Continuous Functions in Topological Spaces
    If \mathcal O_X is the indiscrete topology and \mathcal O_Y is the discrete topology, then only the constant functions f:X\to Y are continuous. Example 4. If f: ...
  30. [30]
    Discrete Set -- from Wolfram MathWorld
    On any reasonable space, a finite set is discrete. A set is discrete if it has the discrete topology, that is, if every subset is open. In the case of a subset ...
  31. [31]
    [PDF] Finite Spaces Handouts 1 - UChicago Math
    Jun 19, 2012 · Definition 1.1. 5. For any infinite set X, the cofinite topology is defined as Ucof = {U | X\U is finite } ∪ {∅} In other words, close sets are ...
  32. [32]
    [PDF] Chapter 4 Topologies and interiors (postponable)
    Then Σ = {S ⊆ Ω : S = ∅ or S is cofinite} is the resulting cofinite topology. ... In 4.2 we defined the topology in terms of the interior. That can be done ...
  33. [33]
    [PDF] Chapter III Topological Spaces
    If is a finite set, then the cofinite topology is the same as the discrete topology on . ... (using the properties of a homeomorphism and the definition of.
  34. [34]
    [PDF] MATH4530–Topology. PrelimI Solutions
    Sep 30, 2021 · Solution: (1) Compact: Any infinite set with finite complement topology is compact. The proof is as follows. Let X be an infinite set with the ...
  35. [35]
    [PDF] Math 4853 homework 51. Let X be a set with the cofinite topology ...
    For a set X with cofinite topology, every subspace of X has the cofinite topology, meaning every subset of X is compact.
  36. [36]
    [PDF] REAL ANALYSIS MATH 608 HOMEWORK #1 Problem 1. (1) Show ...
    (2) By definition every finite set is closed in the cofinite topology, so in particular singletons are closed. ... It is not second countable. Indeed if B.
  37. [37]
    [PDF] Sequences in the cofinite topology
    Show that if {xn} is a sequence in R with no repeated terms, then {xn} converges to every real number. 2. Consider the constant sequence xn = 0 ∀n. Show this ...<|control11|><|separator|>
  38. [38]
    [PDF] MAT327 - Lecture 1
    May 8, 2019 · Definition : Co-countable Topology. We define the co-countable topology on a set X as: {U ⊆ X : X \ U is countable} ∪ {∅}. Showing that ...
  39. [39]
    [PDF] About these notes - MIT ESP
    (a) The cofinite topology: A set V ⊆ X is open ... We say that V is open (in this topology) iff the following property is true: “For every ... Prove that the ...
  40. [40]
    [PDF] Non-Hausdorff T1 Properties - arXiv
    Apr 1, 2024 · The co-countable topology on an uncountable set is k1H but not T2. To see this, first observe that the space is hyperconnected: every pair ...
  41. [41]
    [PDF] Study of Compactness with respect to some special Topologies on IR
    Property 4: R with Co-countable topology is not compact. Proof. Assume (R,τ4) is compact. For each, ∈ Q,Qx = (R-Q)U1xl be an open set.
  42. [42]
    order topology
    ### Definition of Order Topology
  43. [43]
    [PDF] Section 14. The Order Topology
    May 29, 2016 · (1) Comparability: For every x, y ∈ A for which x 6= y, either x<y or y<x. (2) Nonreflexivity: For no x ∈ A does the relation x<x hold.
  44. [44]
    [PDF] Topology of the Real Numbers - UC Davis Math
    We usually abbreviate “sequentially compact” to “compact,” but sometimes we need to distinguish explicitly between the sequential definition of compactness.
  45. [45]
    Sorgenfrey topology - Encyclopedia of Mathematics
    Nov 26, 2023 · The Sorgenfrey line serves as a counterexample to several topological properties, see, for example, [a3]. For example, it is not metrizable (cf.Missing: Steinhaus | Show results with:Steinhaus
  46. [46]
    [PDF] Subspaces of the Sorgenfrey Line by Dennis K. Burke, Miami ...
    The Sorgenfrey Line S is an elementary example [S] of a topological space almost always introduced and studied in the first basic topology course. Many of the ...Missing: Steinhaus | Show results with:Steinhaus
  47. [47]
    topology of the complex plane - PlanetMath
    Mar 22, 2013 · If we identify R2 ℝ 2 and C ℂ , it is clear that the above topology coincides with topology induced by the Euclidean metric on R2 ℝ 2 .<|separator|>
  48. [48]
    11.1 Complex numbers
    A complex number is just a pair ( x , y ) ∈ R 2 on which we define multiplication (see below). We call the set the complex numbers and denote it by ...
  49. [49]
    [PDF] Closed sets and the Zariski topology
    Therefore, if k is infinite, the Zariski topology on k is not Hausdorff. Definition 1.5. In a topological space X, the ...
  50. [50]
    [PDF] 1.1. Algebraic sets and the Zariski topology. We have said in the ...
    ... Zariski topology. For example, in the usual complex topology, the affine line A1 (i. e. the complex plane) is reducible because it can be written e. g. as ...<|separator|>
  51. [51]
    line with two origins in nLab
    Sep 19, 2021 · The line with two origins is the topological space which results by “gluing” a copy of the real line identically to itself, except at the origin.
  52. [52]
    Line with two origins - Topospaces
    May 31, 2016 · This article describes a standard counterexample to some plausible but false implications. In other words, it lists a pathology that may be useful to keep in ...
  53. [53]
    [PDF] Section 18. Continuous Functions
    Jun 11, 2016 · Our definition is based entirely on open sets and we should in fact state that f is continuous relative to the topologies on X and Y . Lemma 18.
  54. [54]
    [PDF] 6. Continuity and homeomorphisms
    In particular we will define a special type of function—a continuous function— between topological spaces in such a way that some amount of the topological ...
  55. [55]
    [PDF] topbook.pdf - Topology Without Tears by Sidney A. Morris
    Jan 6, 2021 · I aim in this book to provide a thorough grounding in general topology. ... continuity which lends itself more to generalization. 1The ...
  56. [56]
    [PDF] 1 Nets and sequences - Humboldt-Universität zu Berlin
    In metric spaces, a standard theorem states that sequential continuity is equivalent to continuity. In arbitrary topological spaces this is no longer true, but ...
  57. [57]
    [PDF] Filters in Analysis and Topology - David R. MacIver
    Jul 1, 2004 · By continuity there is a neighbourhood U of x such that g(U) ⊆ V . But U ∈ F, so g(U) ∈ g(F). Thus, as g(F) is a filter, V ∈ g(F). Hence ...
  58. [58]
    [PDF] CHARACTERIZING CONTINUITY IN TOPOLOGICAL SPACES*
    Velleman proved that a mapping of IR to IR is continuous iff the images of compact sets are compact and ones of connected sets are connected.
  59. [59]
    [PDF] 6. Continuity and homeomorphisms
    Before we do anything with continuous functions, here are some examples. We will start with examples in and around the real numbers, where the reading is ...
  60. [60]
    homeomorphism - PlanetMath.org
    using the techniques of topology, there is no way of ...
  61. [61]
    diffeomorphism in nLab
    Jun 28, 2017 · Observation 2.1. Every diffeomorphism is in particular a homeomorphism between the underlying topological spaces. The converse in general fails.
  62. [62]
    [PDF] Math 396. Gluing topologies, the Hausdorff condition, and examples
    The ability to glue without having to similtaneously confront other structures (such as metrics) is a big advantage of working with topological spaces rather ...
  63. [63]
    Compact Space -- from Wolfram MathWorld
    A topological space is compact if every open cover of X has a finite subcover. In other words, if X is the union of a family of open sets, there is a finite ...
  64. [64]
    Heine-Borel Theorem -- from Wolfram MathWorld
    The Heine-Borel theorem states that a subspace of R^n (with the usual topology) is compact iff it is closed and bounded.
  65. [65]
    properties of compact spaces - PlanetMath.org
    Mar 22, 2013 · A closed subset of a compact space is compact · A compact subspace of a Hausdorff space is closed · The continuous image of a compact space is ...
  66. [66]
    [PDF] The Tychonoff theorem
    Dec 16, 2020 · We saw that the product of two compact spaces equipped with product topology is compact, following the similar argument, we could show that this ...Missing: Andrey 1930
  67. [67]
    [PDF] Chapter 3. Connectedness and Compactness
    Jul 14, 2016 · Space X is connected if there is no separation of X. Note. An alternative definition is that X is connected if and only if the only subsets of X ...
  68. [68]
    [PDF] 18. Connectedness
    A topological space (X,T ) is connected if and only if every continuous func- tion f : X → {0,1} is constant (where {0,1} has the discrete topology). Proof.
  69. [69]
    [PDF] R. Engelking: General Topology Introduction 1 Topological spaces
    Jan 1, 2012 · 4) Connectedness is an invariant of continuous mappings. Theorem. (6.1.7) A subspace C of a topological space X is connected if and only if for ...
  70. [70]
    [PDF] Spaces that are connected but not path connected - Keith Conrad
    A topological space X is called connected if it's impossible to write X as a union of two nonempty disjoint open subsets: if X = U ∪ V where U and V are open ...<|separator|>
  71. [71]
    [PDF] Section 31. The Separation Axioms
    Aug 30, 2016 · Normal Spaces ⊂ Regular Spaces ⊂ Hausdorff Spaces ⊂ Tychonoff Spaces. We show by example below that the first two inclusions are proper. Note.
  72. [72]
    [PDF] MA-231 Topology - IISc Math
    Nov 29, 2004 · The T2 axiom, introduced by Hausdorff under the name “separability” is probably the most useful separation axiom especially where convergence is ...
  73. [73]
    [PDF] MA651 Topology. Lecture 6. Separation Axioms.
    A T0 space is sometimes called a Kolmogorov space and a T1 space, a Fréchet space. A T2 is called a Hausdorff space. Each of axioms in Definition (36.1) is ...
  74. [74]
    [PDF] Grundzüge der Mengenlehre
    Das vorliegende Werk will ein Lehrbuch und kein Bericht sein: es versucht die Hauptsachen der Mengenlehre ohne Voraussetzung.
  75. [75]
    Über die Mächtigkeit der zusammenhängenden Mengen - EuDML
    Urysohn, P.. "Über die Mächtigkeit der zusammenhängenden Mengen." Mathematische Annalen 94 (1925): 262-295. <http://eudml.org/doc/159111>.Missing: zusammenfügbaren | Show results with:zusammenfügbaren<|control11|><|separator|>
  76. [76]
    [PDF] Section 30. The Countability Axioms
    Dec 8, 2016 · If a topological space X has a countable basis for its topology, then. X satisfies the Second Countability Axiom, or is second-countable. Note.Missing: general ccc
  77. [77]
    Separability, the Countable Chain Condition and the Lindelöf ... - jstor
    In this section, we will attempt to show how our results can be applied in general topology. (4.1) EXAMPLE. Souslin spaces. Recall that a Souslin space is a.
  78. [78]
    Sur quelques points du calcul fonctionnel | Rendiconti del Circolo ...
    Dec 23, 2008 · Download PDF ... Cite this article. Fréchet, M.M. Sur quelques points du calcul fonctionnel. Rend. Circ. Matem. Palermo 22, 1–72 (1906).
  79. [79]
    [PDF] METRIC SPACES 1. Introduction As calculus developed, eventually ...
    It is a set on which a notion of distance between each pair of elements is defined, and in which notions from calculus in. R (open and closed intervals, ...
  80. [80]
    Urysohn metrization theorem - Encyclopedia of Mathematics
    Oct 15, 2014 · A compact or countably compact Hausdorff space is metrizable if and only if it has a countable base: indeed, it is homeomorphic to a subset of the Hilbert cube.Missing: paper | Show results with:paper
  81. [81]
    59 SHORTER NOTICES - Project Euclid
    By André Weil. Paris, Hermann, 1937. 39 pp. A space E with uniform structure is a topological space restricted as follows. There exists on £ a system of ...
  82. [82]
    [PDF] Notes on Uniform Structures - UC 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.
  83. [83]
    [PDF] NOTES ON BAIRE'S THEOREM Example. A complete metric space ...
    A Hausdorff topological space X is a Baire space if countable intersections of open dense subsets of X are dense. Thus Baire's theorem says complete metric ...Missing: statement | Show results with:statement
  84. [84]
    None
    ### Summary of Baire Category Theorem Content from http://homepages.math.uic.edu/~rosendal/WebpagesMathCourses/MATH511-notes/DST%20notes%20-%20BaireCategory09.pdf
  85. [85]
    [PDF] Baire Spaces
    Theorem 5.5 (Baire category theorem for locally compact Hausdorff space). Every locally compact Hausdorff space X is a Baire space. Proof. Let {Un} n∈N.<|control11|><|separator|>
  86. [86]
    [PDF] Interpreter for topologists - University of Florida
    Definition 1.4 Suppose that M |= X,Y are topological spaces and f : X → Y is a continuous function. Suppose that π: X → ˆX and χ: Y → ˆY are interpretations. An.
  87. [87]
    [PDF] filter convergence and tychonoff's theorem
    Definition. A filter on a set S is a nonempty family F of subsets of S such that (i)–(iv) hold: (i) 0 /∈ F;. (ii) if A ∈ F and B ⊇ A, then B ∈ F;.
  88. [88]
    [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 ...
  89. [89]
    [PDF] Nets and filters (are better than sequences)
    We will also define a type of object called a filter and show that filters also furnish us with a type of convergence which turns out to be equivalent to net ...
  90. [90]
    Section 6.11 (0078): Stalks—The Stacks project
    6.11 Stalks. Let X be a topological space. Let x \in X be a point. Let \mathcal{F} be a presheaf of sets on X. The stalk of \mathcal{F} at x is the set.
  91. [91]
    [PDF] Continuum Theory, Cantor Sets, and the Topology of Dimension
    A Peano continuum is a Peano space which is a continuum. Theorem (Hahn-Mazurkiewicz). Every Peano continuum is a continuous image of the closed interval [0,1].
  92. [92]
    [PDF] ON SPACE-FILLING CURVES AND THE HAHN-MAZURKIEWICZ ...
    These are notes on space-filling curves, looking at a few examples and proving the Hahn-Mazurkiewicz theorem. ... We start by looking at Peano's original example.
  93. [93]
    [PDF] Hyperspaces of Peano continua are Hilbert cubes
    The main results of this paper are that the hyperspace 2x of nonempty closed subsets of a nondegenerate Peano continuum X is homeomorphic to the Hilbert Cube Q, ...
  94. [94]
    [PDF] notions of dimension - Cornell Mathematics
    Jun 9, 2010 · The topological dimension is defined as the smallest value of m for which this statement holds. We denote it by dimT (X).
  95. [95]
    Lebesgue covering dimension (Chapter 1)
    This chapter concentrates on one of these definitions, the Lebesgue covering dimension, which we will denote by dim(X), and refer to simply as the covering ...
  96. [96]
    Brouwer: Development of Topology, Invariance of Dimension
    Nov 1, 2025 · In fact, Brouwer published two proofs of dimensional invariance, the first in 1911 and the second in 1913, both full of deep insights. Moreover, ...Missing: paper | Show results with:paper
  97. [97]
    [1307.6625] Dimension-Raising Maps in a Large Scale - arXiv
    Jul 25, 2013 · In this paper we introduce a new notion of finite-to-one like map in a large scale setting. Using this notion we formulate a dimension-raising ...Missing: non- 2020s topology
  98. [98]
    Set-Theoretic Topology - ScienceDirect.com
    Under continuum hypothesis (CH), many of the known constructions of S- and L-spaces can be modified to produce strong S- and L-spaces. The existence of ...
  99. [99]
    [PDF] Normality and Martin's Axiom - Biblioteka Nauki
    It is known [13] that Martin's Axiom is independent of the ZFC axioms of set theory. It is implied by the Continuum Hypothesis. Nevertheless, it is also.
  100. [100]
    [PDF] Pontryagin Duality and the Structure of, Locally Compact Abelian ...
    Definition. Let G1 and G₂ be topological groups. A map f: G1 G2 is said to be a continuous homomorphism if it is both a homomorphism of groups and continuous.
  101. [101]
    [PDF] Some notes on Pontryagin duality of abelian topological groups - arXiv
    Oct 28, 2025 · Some observations on topological abelian groups and their duality appear to be useful to recall, notably those with an emphasis on ...
  102. [102]
    [PDF] Hahn-Banach theorems 1. Continuous Linear Functionals 2 ...
    Jul 17, 2008 · Second, continuous linear functionals on subspaces of a locally convex topological vectorspace have continuous extensions to the whole space.
  103. [103]
    245B, Notes 6: Duality and the Hahn-Banach theorem - Terence Tao
    Jan 26, 2009 · The Hahn-Banach theorem is true when X, Y are real normed vector spaces. Proof. This is a standard “Zorn's lemma” argument. Fix Y, X, \lambda .
  104. [104]
    [PDF] Notes on point-free topology
    Abstract Point-free topology is the study of the category of locales and localic maps and its dual category of frames and frame homomorphisms.
  105. [105]
    [PDF] Pointfree topology and constructive mathematics - arXiv
    Nov 23, 2024 · A locale is spatial if its opens are completely determined by the points they contain. Not all locales are spatial.
  106. [106]
    [PDF] LOCALES ARE NOT POINTLESS
    LOCALES ARE NOT POINTLESS. STEVEN VICKERS. Department of Computing, Imperial College. 180 Queen's Gate, London SW7 2BZ, United Kingdom. E-mail: sjv@doc.ic.ac.uk.<|separator|>
  107. [107]
    [PDF] and Bing-Nagata-Smirnov Metrization Theorems - DiVA portal
    In this subject the Bing-Nagata-Smirnov metrization theorem is crucial in order to verify the existence of metrics on certain sub-structures on manifold- like ...
  108. [108]
    [PDF] Section 40. The Nagata-Smirnov Metrization Theorem
    Oct 1, 2016 · The Nagata-Smirnov Metrization Theorem. A topological space X is metrizable if and only if X is regular and has a basis that is countably ...
  109. [109]
    [PDF] Large cardinals beyond HOD - arXiv
    Sep 12, 2025 · Abstract. Exacting and ultraexacting cardinals are large cardi- nal numbers compatible with the Zermelo-Fraenkel axioms of set.
  110. [110]
    Nonlinearity and illfoundedness in the hierarchy of large cardinal ...
    Apr 30, 2025 · In this paper I present counterexamples, natural instances of nonlinearity and illfoundedness in the hierarchy of large cardinal consistency strength.