Fact-checked by Grok 2 weeks ago

Finite topological space

A finite topological space is a whose underlying set is a , equipped with a collection of subsets satisfying the axioms of a : it includes the and the whole space, and is closed under arbitrary unions and finite intersections. These spaces provide concrete examples for studying fundamental topological concepts, such as , connectedness, and , without the complications of infinite sets. Finite topological spaces are particularly notable for their minimal open neighborhoods, which form a basis for the and allow exhaustive enumeration of all open sets due to the finiteness of the underlying set. In the T_0 case—where distinct points have distinct neighborhood systems—these spaces correspond bijectively to finite partially ordered sets (posets), with the order defined by inclusion of minimal neighborhoods, enabling translations between and . Every finite T_0 space contains isolated points, and its closure and interior operators can be described in terms of the associated partial order. Beyond pedagogy, finite topological spaces have applications in algebraic topology, where they model the weak homotopy types of finite CW-complexes, as shown by their singular homology and homotopy groups matching those of associated simplicial complexes. They also play a role in digital and computational topology for analyzing image data, shape theory of compact metric spaces, and even ringed spaces in scheme theory, highlighting their utility despite often failing stronger separation axioms like T_1 unless discrete. The number of distinct topologies on a set of n elements grows rapidly, with known enumerations up to moderate n, underscoring the combinatorial richness of these structures.

Fundamentals

Definition of Finite Topologies

A finite topological space is a topological space (X, \tau) in which the underlying set X is finite, meaning |X| < \infty, and \tau is a topology on X. A topology \tau on a set X is a subset of the power set \mathcal{P}(X), consisting of the empty set \emptyset, the full set X, and collections of subsets (open sets) that are closed under arbitrary unions and finite intersections. The power set \mathcal{P}(X) denotes the set of all subsets of X, which for finite X is itself finite with cardinality $2^{|X|}. On a X, any collection of subsets that includes \emptyset and X and is closed under arbitrary unions and finite intersections automatically forms a \tau \subseteq \mathcal{P}(X), as the finiteness of X ensures that arbitrary unions remain within \mathcal{P}(X) without exceeding the finite number of possible subsets. This construction guarantees that the closed sets, defined as the complements of the open sets in \tau, also form a collection closed under arbitrary intersections and finite unions, providing a perspective on the structure. In finite topological spaces, the finiteness of the open sets in \tau equivalently implies that the closed sets are finite, since each closed set is the complement of an open set and \mathcal{P}(X) is finite. Moreover, every of a finite topological space is itself finite, as it inherits a from a of the finite set X. Finite topological spaces have been utilized since the early development of to construct counterexamples, such as compact spaces that are non-metrizable due to failure of Hausdorff separation. In 1994, emphasized their value in his essay "On Proof and Progress in Mathematics," describing the of finite topologies as "an oddball topic that can lend good insight to a variety of questions."

Relation to Alexandroff Spaces and Preorders

An Alexandroff space is a topological space in which arbitrary intersections of open sets are open. Every finite topological space is an Alexandroff space, as the intersection of any collection of open sets in a finite space reduces to a finite intersection, which remains open by definition of a topology. In particular, finite T_0 spaces—those satisfying the Kolmogorov separation axiom, where distinct points have distinct minimal open neighborhoods—are precisely the Alexandroff spaces arising from partial orders. The specialization preorder on a topological space (X, \tau) is defined by x \leq y if and only if x \in \overline{\{y\}}, the closure of the singleton \{y\}. This relation is reflexive and transitive, forming a on X. In a finite T_0 space, the specialization preorder is antisymmetric, hence a partial order (poset), because the T_0 condition ensures that if x \leq y and y \leq x, then the minimal open neighborhoods of x and y coincide, implying x = y. The open sets in such a space are exactly the upper sets with respect to this poset: a subset U \subseteq X is open if whenever x \in U and x \leq y, then y \in U. There is a bijection between the T_0 topologies on a finite set X and the partial orders on X. Given a finite poset (X, \leq), the corresponding Alexandroff topology has as its open sets the unions of principal up-sets \uparrow x = \{y \in X \mid x \leq y\} for x \in X, which are precisely the up-sets in the poset. Conversely, from a finite T_0 topology on X, the specialization preorder yields a poset whose associated up-sets recover the original topology. This equivalence holds because, in the finite case, the T_0 condition guarantees that the specialization preorder uniquely determines the minimal open neighborhoods, and thus the entire topology via arbitrary unions (which coincide with finite unions). To sketch the recovery: for a finite T_0 space (X, \tau), the minimal open neighborhood U_x of each x \in X satisfies U_x = \uparrow x = \{y \in X \mid x \leq y\} under the specialization preorder, and every open set is a union of such U_x. Thus, applying the up-set construction to the specialization poset regenerates \tau. Every finite poset induces a unique T_0 topology in this manner, allowing order-theoretic properties, such as chains or antichains, to be translated directly into topological features like connectedness or the number of components.

Examples

Trivial and Small Cases (n ≤ 2)

For the , which has zero points, there is a unique : the collection consisting solely of the itself as the only . A set with a single point p, denoted \{p\}, admits exactly one : \{\emptyset, \{p\}\}. In this case, the discrete (where all subsets are open) and the indiscrete (where only the and the whole space are open) coincide, as these are the only possible subsets. For a two-point set \{a, b\}, there are four distinct labeled topologies. These are:
  • The indiscrete topology: \{\emptyset, \{a, b\}\}, with only the empty set and the full space open.
  • The discrete topology: \{\emptyset, \{a\}, \{b\}, \{a, b\}\}, where every subset is open.
  • The Sierpiński topology with a distinguished: \{\emptyset, \{a\}, \{a, b\}\}.
  • The Sierpiński topology with b distinguished: \{\emptyset, \{b\}, \{a, b\}\}.
Up to , these reduce to three inequivalent : the indiscrete one, the one, and the Sierpiński topology (where the two labeled variants are homeomorphic via relabeling). The Sierpiński topology provides a basic example of a non-Hausdorff T_0 , satisfying the Kolmogorov but not the Hausdorff condition, as the two points cannot be separated by disjoint open neighborhoods. A key feature of the Sierpiński space is that its open sets form a chain under inclusion: \emptyset \subset \{a\} \subset \{a, b\} (or analogously for the b-distinguished variant), illustrating a simple linear structure absent in the discrete or indiscrete cases. This corresponds to the associated specialization preorder, where one point precedes the other but not vice versa.

Cases with 3 Points

On a set with three elements, such as {a, b, c}, there are exactly 29 distinct topologies when labeling the points. This count marks the first case where finite topologies display significant variety, including the initial appearance of spaces with non-trivial connected components, such as a disconnected of an and a connected two-point . Up to , these 29 topologies reduce to 9 inequivalent classes, categorized by their collections, T0 status, and , revealing patterns like chains, forks, and asymmetric structures. Classification proceeds by associating each topology to a preorder on the set, where the open sets are the upsets—subsets closed upward under the specialization relation x \leq y if every open containing y contains x. For construction, one specifies binary decisions for each pair of points: whether x \leq y, y \leq x, both (equivalence), or neither (incomparable), then takes the upsets generated by the principal upsets \uparrow x = \{y \mid x \leq y\} as the basis, ensuring closure under arbitrary unions and finite intersections. This preorder-based approach systematically enumerates all possibilities, with T0 topologies corresponding to antisymmetric preorders (partial orders) and non-T0 ones allowing equivalences. Key examples highlight emerging properties. The discrete topology has all 8 subsets open, is T0, but disconnected with three components. The indiscrete topology admits only \emptyset and {a,b,c} as open, is connected and hyperconnected (every nonempty open is dense), but not T0. The linear order topology for the chain poset a < b < c features open sets \emptyset, \{a\}, \{a,b\}, \{a,b,c\}, yielding a connected T0 space that fails T1 since closures of singletons overlap. A T0 but not T1 example is the poset with two minimal elements a, b both below c (), where opens are \emptyset, \{a\}, \{b\}, \{a,b\}, \{a,b,c\}; singletons {a} and {b} are open but {c} is not closed, as its closure includes everything. Fork-like structures, such as the poset with c maximal and a, b minimal elements below it, produce connected T0 topologies with a single branch point. Hyperconnected spaces beyond the indiscrete include those where one point specializes to all others, ensuring no proper clopen subsets exist. The on two points appears as a in several three-point extensions, underscoring its role in building asymmetric connected examples. The following table summarizes the 9 homeomorphism classes, with representative proper open sets (beyond \emptyset and the full set), T0 status, and connectivity:
ClassRepresentative Proper OpensT0?Connected?
Discrete (D3){a}, {b}, {c}, {a,b}, {a,c}, {b,c}YesNo
Indiscrete (I3)NoneNoYes
Included point (one open singleton){a}NoYes
Two open singletons, disconnected{a}, {b}YesNo
Chain of three (linear){a}, {a,b}YesYes
V-poset (two mins below one max){a}, {b}, {a,b}YesYes
Inverted V (one min below two maxs){b,c}NoYes
Sierpiński-like (one min, two branches){a}, {a,b}, {a,c}YesYes
Paired indiscrete (equivalence on two, isolated){a,b}NoNo

Cases with 4 Points

There are distinct topologies on a labeled 4-point set. Up to , these reduce to 33 inequivalent types, which can be classified by invariants such as the minimal number of open sets, the structure of the , or the rank (height) of the associated in the T0 case. Of these, 16 are T0 topologies, corresponding to the 16 non-isomorphic posets on 4 elements. Representative examples among the 33 types include the discrete topology, in which all 16 subsets are open, and the indiscrete topology, with only the and the full space open. A particularly interesting non-T0 example is the pseudocircle, defined on points {a, b, c, d} with basis opens {a, c}, {b, d}, {a, b, c}, and {a, b, d}; this space is weakly equivalent to the circle S^1 and exhibits non-trivial despite its finiteness. Other notable types include spaces with multiple components, such as the disjoint union of a 2-point discrete space and a Sierpiński space (where one singleton is open), illustrating disconnectedness while preserving compactness. Approximations to finite projective planes appear in certain configurations, such as topologies where open sets represent "lines" through points in a near-pencil arrangement, modeling incidence structures with 4 points and 6 lines akin to a degenerate projective geometry. With 4 points, symmetry becomes more prominent, as the symmetric group S_4 acts on the set, and many of the 33 types admit non-trivial homeomorphisms (automorphisms) fixing the topology; notably, over half are Alexandroff spaces isomorphic to those induced by preorders, with the T0 subset directly tied to poset realizations. These classifications can be verified and explored computationally using , which enumerates all posets up to via its combinat module and extends to preorder-based topologies, or , which handles transitive digraphs equivalent to finite topologies through its and poset packages. For instance, hyperconnected topologies on 4 points, where no proper non-empty is both open and closed, emerge frequently in such enumerations, highlighting the prevalence of indecomposable structures.

Core Properties

Compactness, Countability, and Separation Axioms

Finite topological spaces exhibit strong compactness properties due to their finite underlying sets. A topological space is compact if every open cover admits a finite subcover. In a finite space X, any open cover \{U_i\}_{i \in I} of X can be refined to a finite subcollection by selecting, for each point x \in X, an open set U_{i_x} containing x; since X is finite, this subcollection is finite and covers X. Thus, every finite topological space is compact. Moreover, compactness implies the Lindelöf property, where every open cover has a countable subcover; here, the subcover is finite, hence countable. Finite spaces are also paracompact, as every open cover has a locally finite open refinement: the finite subcover from compactness is locally finite, since each point in the finite space intersects only finitely many sets in any cover. Countability axioms hold trivially for finite spaces. A space is second-countable if it has a countable basis for its . The power set of a finite set X is finite, so the collection of all open sets in any on X is finite and serves as a basis; thus, every finite topological is second-countable. Separability requires a countable dense ; the entire X is finite (hence countable) and dense in itself, so finite spaces are separable. Separation axioms simplify significantly in finite spaces, often reducing to discrete cases. A space satisfies the T0 axiom (Kolmogorov) if for distinct points x, y \in X, there exists an open set containing one but not the other; this is common in finite topologies, particularly Alexandroff spaces equivalent to preorders via the specialization order. The T1 axiom (Fréchet) requires singletons to be closed. In finite spaces, T1 holds if and only if the topology is discrete: if T1, then for any subset U \subseteq X, its complement H = X \setminus U is a finite union of closed singletons, hence closed, so U is open; thus, all subsets are open. Conversely, the discrete topology clearly satisfies T1. No finite space is regular or higher without being discrete, as T1 already forces discreteness, and higher axioms like T2 (Hausdorff) coincide with T1 in this setting. In finite spaces, the R0 axiom (symmetric separation via closures) is equivalent to T0, reflecting the preorder structure where symmetry and antisymmetry align under finiteness constraints. Finite spaces provide counterexamples to generalizations of metrizability in compact settings: while compact spaces are second-countable, finite compact spaces need not be metrizable unless T1 (i.e., ), as non-T1 examples like the are compact but fail Hausdorff separation, hence non-metrizable.

Specialization Preorder and Connectivity

In finite topological spaces, the provides a natural way to encode topological structure through . For a finite space X, the specialization \leq is defined by x \leq y if and only if x belongs to the of the \{y\}, or equivalently, if the minimal open neighborhood U_y of y contains x. This relation is reflexive and transitive, forming a on X; it becomes a partial order (poset) precisely when X is T_0, establishing a between finite T_0 topological spaces and finite posets. The can be viewed as a on the points of X, with an arc from x to y if x < y (strict inequality). In this , the underlying undirected captures comparability relations regardless of direction. A finite space X is connected if and only if its specialization is order-connected, meaning that for any two points x, y \in X, there exists a finite sequence x = z_0, z_1, \dots, z_n = y such that consecutive elements are comparable in the preorder (i.e., z_i \leq z_{i+1} or z_{i+1} \leq z_i). This condition is equivalent to the underlying undirected of the preorder being weakly connected, often referred to as the preorder being "" in the sense of rather than total comparability. In finite spaces, connected components correspond to the equivalence classes under this order-connectedness relation, and the upset of open sets (in the T_0 case) translates poset components into disjoint unions of principal opens. Unlike in spaces, where connectedness may not align directly with preorder properties due to accumulation points, finite spaces exhibit this tight equivalence, allowing topological connectivity to be fully determined by the preorder's . In finite topological spaces, connectedness is equivalent to path-connectedness; the connected components are precisely the components. Path-connectedness in finite spaces aligns with connectedness. Continuous paths exist between points in the same via suitable parameterizations that respect the minimal neighborhoods, such as step functions switching at interior points of [0,1]. For comparable points x \leq y, paths can be constructed in appropriate directions consistent with the order. Full path-connectedness holds for connected finite spaces, including chains (total orders). For instance, the , with points \{0,1\} and non-trivial \{0\}, corresponds to the chain $0 \leq 1. It is connected, hyperconnected, and T_0, and path-connected, as continuous paths join 0 and 1 using step functions (e.g., constant then switch). Stronger connectivity notions further exploit the . A finite space is hyperconnected (or irreducible) if every pair of non-empty open sets intersects; this holds if and only if the specialization has a greatest element, whose is the entire space X, ensuring all upsets overlap at that element. Dually, the space is ultraconnected if every pair of non-empty closed sets intersects, equivalent to the having a least element. In finite posets, these properties manifest in the upset/ structure, where connected components via the preorder graph directly inform the topology's irreducibility.

Advanced Features

Metrizability and Additional Structures

Finite topological spaces admit pseudometrics under specific conditions related to their separation properties. A finite space is pseudometrizable if and only if it satisfies the R0 separation axiom, which requires that for distinct points x, y, if x is in the closure of \{y\} then y is in the closure of \{x\}. In R0 spaces, the specialization preorder is an equivalence relation, partitioning the space into equivalence classes where points within a class are indistinguishable (have identical neighborhood systems). A pseudometric can be defined by setting d(x, y) = 0 if x and y are in the same equivalence class and d(x, y) = 1 otherwise; this generates the topology, with open balls of radius less than 1 being the equivalence classes (indiscrete within) and larger balls the whole space or unions. For general finite spaces, quasi-pseudometrics or partial pseudometrics can model the preorder, assigning distances based on the order, such as d(x, y) = 1 if x \leq y and x \neq y, d(x, y) = 0 otherwise, but these are asymmetric unless R0 holds. For full metrizability, which requires a genuine metric (with d(x, y) = 0 implying x = y), finite topological spaces are metrizable only if they are discrete. This follows because metrizability implies the Hausdorff (T2) property, and any finite Hausdorff space must be discrete, as distinct points can be separated by disjoint open sets, leading to singleton opens. The second-countability required for metrizability in compact spaces is automatically satisfied in finite settings, reinforcing that non-discrete finite spaces fail regularity or other axioms needed for a separating metric. Additional structures on finite spaces, such as uniformities, are inherently limited. Every finite topological space admits a unique quasi-uniform structure compatible with its , generated from the , but true uniform structures exist only if the space is R0, in which case the uniformity is precompact and induced by the pseudometric. These uniformities are trivial in the sense that entourages are finite unions of equivalence classes from the specialization relation, reflecting the of finite spaces. For algebraic structures, paratopological groups—where multiplication is jointly continuous but inversion may not be—on finite sets reduce to topologies, as non-discrete multiplications would violate finiteness without yielding continuous inverses. Finite topologies find practical applications in through structures like the Khalimsky line, a T0 topology on the integers where even points have discrete neighborhoods and odd points connect adjacent evens. This topology models digital connectivity without paradoxes like thin tunnels, enabling robust algorithms for segmentation and boundary detection in binary images. The Khalimsky plane extends this to 2D grids, preserving topological invariants for tasks.

Homotopy in Algebraic Topology

Finite topological spaces play a significant role in , particularly through their weak equivalences to CW-complexes. A result establishes that every finite topological space X admits a weak equivalence to a finite obtained via the of its associated , where the arises from the specialization order on X. This subdivision constructs the order complex K(X), whose geometric realization |K(X)| is weakly equivalent to X, preserving all groups. Such equivalences allow finite spaces to model the weak types of finite CW-complexes, enabling the computation of invariants like and groups using combinatorial tools from poset theory. A notable example is the pseudocircle, a 7-point finite T₀-space constructed as a poset that is weakly equivalent to the circle S^1. This space features a non-trivial \pi_1 \cong \mathbb{Z}, demonstrating that finite spaces can capture infinite cyclic groups despite their discrete nature. The poset structure consists of points arranged to mimic the looping behavior of , with minimal and maximal elements forming chains that induce the desired homotopy type upon subdivision. This construction highlights how finite models can approximate continuous manifolds while remaining computationally tractable. The of a finite topological space is computable via paths in its , specifically through the edge-path group of the order complex K(X), which aligns with the classical presentation using loops and relations. This approach extends to finite models of higher-dimensional spheres or tori; for instance, minimal finite T₀-spaces exist that are weakly equivalent to S^n for n \geq 1 or products like S^1 \times S^1, with matching those of the targets. Developments since the , including Barmak's comprehensive treatment of finite space , have advanced these models, with applications in for , where finite spaces approximate image topologies to analyze connectivity and holes in pixelated data. Finite spaces uniquely classify the weak types of compacta through extensions of their models, where the is obtained by iteratively removing beat points (elements beat by others in the ). This classification aligns with Scott's model framework for finite spectra, providing a combinatorial basis for stable categories via poset representations. Path components in rely on the underlying connectivity of the , ensuring that these models faithfully reproduce algebraic invariants.

Enumeration

Total Number of Topologies

The total number of distinct topologies on a labeled with n elements, denoted T(n), is a well-studied enumerative quantity in combinatorial . This count includes all possible topologies without regard to equivalence, treating points as distinguishable. The values begin with T(0) = 1, T(1) = 1, T(2) = 4, T(3) = 29, and T(4) = 355, and continue to grow rapidly, reaching T(18) = 261492535743634374805066126901117203 as computed via exhaustive algorithmic .
nT(n)
01
11
24
329
4355
56942
6209527
79535241
8642779354
963260289423
These computations rely on the between topologies and quasi-orders (reflexive, transitive ) on the set, allowing representation as transitive digraphs where open sets correspond to upset . Enumeration proceeds recursively by building the subset lattice and ensuring closure under unions and finite intersections, often implemented computationally with over possible basis elements or relation matrices. No closed-form formula exists for T(n). For unlabeled topologies, counted up to homeomorphism (i.e., isomorphism of the specialization preorder), the sequence p(n) (OEIS A001930) starts with p(1) = 1, p(2) = 3, p(3) = 9, and p(4) = 33. Higher terms include p(5) = 139 and p(6) = 718, with values known up to n=16 via generation of non-isomorphic transitive digraphs using canonical labeling and automorphism group computations.
np(n)
11
23
39
433
5139
6718
74535
835979
As of 2025, no additional terms beyond n=18 for T(n) or n=16 for p(n) have been published, though improvements in algorithms have facilitated related enumerations in poset theory. The growth of T(n) is super-exponential, with \log_2 T(n) \sim n^2 / 4 as n \to \infty.

Number of T0 Topologies

The number of T0 topologies on a with n labeled elements, denoted T0(n), is equal to the number of partial orders on that set. This equivalence arises from the bijection between T0 topologies and partial orders, established via the specialization preorder, where the closure operator defines the order relation x ≤ y if x belongs to the of {y}. The sequence T0(n) is cataloged in OEIS as A001035 and has been computed using recursive enumeration of transitive antisymmetric relations or equivalent combinatorial structures. Representative values for small n are provided in the following table:
nT0(n)
11
23
319
4219
54231
6130023
76129859
8431723379
944511042511
Enumeration of T0(n) typically involves counting the possible upset families closed under unions and containing the whole set, corresponding to the ideals in the poset, or directly tallying valid binary relations satisfying reflexivity, , and antisymmetry; computational algorithms, such as over potential covering relations, confirm these values for n up to 18. For the unlabeled case (up to ), the number of distinct poset structures on n elements is given by OEIS A000112. While the total number of topologies T(n) (OEIS A000798) encompasses all preorders via a similar bijection, T0(n) grows more slowly due to the additional antisymmetry constraint, which eliminates non-distinguishable points in the specialization order. Connections to Dedekind numbers M(n) (OEIS A000372), which count antichains or downsets in the boolean lattice on n elements and relate to free distributive lattices or Dedekind-free posets (posets embeddable as downsets without fixed points under complement), provide context for subsets of T0 structures, though the full enumeration remains focused on all partial orders. No significant analytical advances have emerged post-2023, but enhanced algorithms enable extensions to larger n through optimized relation generation.

References

  1. [1]
  2. [2]
  3. [3]
    Topological Space -- from Wolfram MathWorld
    A topological space, also called an abstract topological space, is a set X together with a collection of open subsets T that satisfies the four conditions.
  4. [4]
    [PDF] On Proof and Progress in Mathematics
    A random continuing example is an understanding of finite topological spaces, an oddball topic that can lend good insight to a variety of questions but that ...
  5. [5]
    [PDF] FINITE TOPOLOGICAL SPACES Contents 1. Alexandroff spaces ...
    Here we relate Alexandroff spaces to the combinatorial notions of preorder and partial order. Definition 2.1. A preorder on a set X is a reflexive and ...
  6. [6]
    [PDF] On homotopy types of Alexandroff spaces - arXiv
    Feb 4, 2009 · With every topological space X we associate its specialization preorder, that is a preorder P(X)=(X, ≤) such that x ≤ y iff y ∈ {x}. This ...<|control11|><|separator|>
  7. [7]
    specialization topology in nLab
    ### Summary of Specialization Topology from nLab
  8. [8]
    [PDF] FINITE TOPOLOGICAL SPACES 1. Introduction - UChicago Math
    A topology on a set X is a set of subsets (open sets) with properties. For finite X, topologies correspond to reflexive and transitive relations ≤ on X.
  9. [9]
    [PDF] Finite Topologies and Partitions
    In this paper, we compute these numbers for k ≤ 17, and arbitrary n, as well as tN0(n, k), the number of all unlabeled non-T0 topologies on E with k open sets, ...
  10. [10]
    The Number of Finite Topologies - jstor
    KLEITMAN AND B. ROTHSCHILD [June the (n+l)st vertex with at most n/64 ... 1970] THE NUMBER OF FINITE TOPOLOGIES 281. (B,+,), or at most n/2 vertices (D ...
  11. [11]
    A000798 - OEIS
    J. I. Brown and S. Watson, The number of complements of a topology on n points is at least 2^n (except for some special cases), Discr. Math.
  12. [12]
    A001930 - OEIS
    J. A. Wright, There are 718 6-point topologies, quasi-orderings and transgraphs, Notices Amer. Math. Soc., 17 (1970), p. 646, Abstract #70T-A106.
  13. [13]
    A000112 - OEIS
    Also the number of unlabeled T_0 topologies with n points. For example, non ... A000798 (labeled topologies), A001035 (labeled posets), A001930 ...
  14. [14]
    Singular homology groups and homotopy groups of finite topological ...
    Project Euclid, September 1966, Singular homology groups and homotopy groups of finite topological spaces, Michael C. McCord.
  15. [15]
    FINITE TOPOLOGICAL SPACES
    If f:F->G is a homeomorphism, then / induces a map of minimal bases, which preserves inclusions and numbers of elements in each base set, so F and G determine ...
  16. [16]
    None
    Below is a merged summary of the properties of finite topological spaces based on the provided segments, combining all information into a single, comprehensive response. To retain maximum detail and ensure clarity, I will use a structured format with tables where appropriate, followed by a narrative summary. The information is synthesized from all four segments, avoiding redundancy while preserving all unique details.
  17. [17]
    Finite Topological Spaces and Quasi-Uniform Structures
    Nov 20, 2018 · The study of quasi-uniform spaces provides a more natural and obviously equivalent characterization of finite topological spaces.
  18. [18]
    [PDF] Every Finite Topology is Generated by a Partial Pseudometric
    Abstract. Given any preorder " on a finite set X, we present an algorithm to construct a partial pseudometric p on X which generates " in the sense that a ...
  19. [19]
    A finite topological space is metrizable iff it is discrete
    Sep 23, 2019 · Let X be a finite topological space. X is metrizable if and only if the topology is discrete. Suppose X is metrizable. That means there exists a metric d on X.How does one show a topological space is metrizable? Using text ...A question about metrizability - Mathematics Stack ExchangeMore results from math.stackexchange.com
  20. [20]
  21. [21]
    [PDF] On Some Classes of Paratopological Groups
    May 3, 2011 · Lemmas 3.4 and 3.5 below will be used to prove that closures of sub- groups of finite products of almost topological groups are subgroups.
  22. [22]
    The Khalimsky Line as a Foundation for Digital Topology
    This study considers the Khalimsky line, that is, the integers, equipped with the topology in which a set is open iff whenever it contains an even integer, it ...
  23. [23]
    [PDF] On Khalimsky Topology and Applications on the Digital Image ...
    May 7, 2015 · Abstract. The main goal of this work is to deal with the Khalimsky digital topology and its application in segmentation.
  24. [24]
    Algebraic Topology of Finite Topological Spaces and Applications
    This volume deals with the theory of finite topological spaces and its relationship with the homotopy and simple homotopy theory of polyhedra.
  25. [25]
    Strong homotopy in finite topological adjacency category
    Aug 15, 2021 · The present paper investigates a strong homotopy (ie, SA-homotopy) in a finite topological adjacency category.
  26. [26]
    Shape of compacta as extension of weak homotopy of finite spaces
    Oct 6, 2021 · We construct a category that classifies compact Hausdorff spaces by their shape and finite topological spaces by their weak homotopy type.
  27. [27]
    [PDF] Finite spaces and larger contexts J. P. May - UChicago Math
    hold for general topological spaces, but it does hold for metric spaces. Actually, what is relevant is not the metric but something it implies. Definition ...
  28. [28]
    THE NUMBER OF FINITE TOPOLOGIES1 infinity. 276
    The logarithm (base 2) of the number of distinct topologies on a set of re elements is shown to be asymptotic to »2/4 as n goes to infinity. Let A be a set with ...
  29. [29]
    [PDF] The Number of Topologies on a Finite Set
    In this paper, we compute T(n, k) for 2 ≤ k ≤ 12, as well as the total number of labeled. T0 topologies on X having n+4, n+5, n+6 open sets. We also give ...
  30. [30]
    [PDF] Counting Unlabelled Topologies and Transitive Relations
    We also successfully recovered the numbers given by Pfeiffer [3] and the values of p(n) up to n = 15 given by Brinkmann and McKay [1]. n q(n) s(n) t(n). 1. 1.
  31. [31]
  32. [32]
    A001035 - OEIS
    J. I. Brown and S. Watson, The number of complements of a topology on n points is at least 2^n (except for some special cases), Discr.
  33. [33]