Fact-checked by Grok 2 weeks ago

Axiom of countability

In , the countability axioms are properties that require the existence of countable collections related to the open sets of a . The second axiom of countability is a fundamental such : a satisfies it if it admits a countable basis—a countable collection of open sets such that every open set in the can be expressed as a of basis elements. This axiom ensures that the has a "manageable" structure in terms of its open sets, distinguishing it from more general topological spaces that may require uncountably many basis elements. In contrast to the first axiom of countability, which requires only that every point in the space has a countable local basis of neighborhoods, the second axiom applies globally to the entire and is strictly stronger. Every is first-countable, but the converse does not hold; for example, the Sorgenfrey line (the real line with the ) is first-countable but not second-countable. Second countability implies several other desirable properties, including the Lindelöf condition (every open cover has a countable subcover) and separability (the existence of a countable dense subset). Classic examples of second-countable spaces include the real line \mathbb{R} with the standard , where the open intervals with rational endpoints form a countable basis, and more generally, any separable . Subspaces and countable products of second-countable spaces remain second-countable, making this axiom particularly useful for studying spaces and manifolds. In metrizable spaces, second countability is equivalent to separability and the Lindelöf property, highlighting its role in bridging topological and concepts.

Definitions

First axiom of countability

In , a topological space X satisfies the first axiom of countability, also known as being first countable, if every point in X has a countable local basis of open neighborhoods. This condition ensures that the neighborhood structure around each point can be "tamed" by a countable collection, allowing for more manageable local analysis compared to spaces without such restrictions. A local basis at a point x \in X, denoted \mathcal{B}_x, is a collection of open sets containing x such that for every open neighborhood U of x, there exists some V \in \mathcal{B}_x with x \in V \subseteq U. The first axiom requires that this \mathcal{B}_x is countable for each x, meaning it can be enumerated as a or list without repetition beyond necessity. This local countability contrasts with the second axiom of countability, which imposes a countable basis on the entire globally. Formally, X is first countable if for each x \in X, there exists a countable family of open neighborhoods \{U_n(x)\}_{n \in \mathbb{N}} such that for any U containing x, there is some n with U_n(x) \subseteq U. This axiom was introduced by in his foundational work on set-theoretic in 1914, where it served as a key local property to study and separation in abstract spaces.

Second axiom of countability

A topological space X satisfies the second axiom of countability, or is second-countable, if its topology admits a countable basis. This means there exists a countable collection \mathcal{B} = \{B_n \mid n \in \mathbb{N}\} of open subsets of X such that every open set in the topology of X can be expressed as a union of elements from \mathcal{B}. Formally, the collection \mathcal{B} forms a basis if for every U \subseteq X and every point x \in U, there exists some B_n \in \mathcal{B} such that x \in B_n \subseteq U. This condition ensures that \mathcal{B} generates the entire through arbitrary unions, providing a countable "building block" structure for all open sets. Any is also first-countable. To see this, fix a point x \in X; the subcollection \{B_n \in \mathcal{B} \mid x \in B_n\} is countable and serves as a local basis at x, since for any open neighborhood U of x, there exists B_n with x \in B_n \subseteq U. Thus, the global countability of \mathcal{B} yields countable local bases at each point via countable subcollections. In contrast to the first axiom of countability, which requires only a countable local basis at each individual point without a uniform global structure, the second axiom imposes a stronger condition by mandating a single countable basis for the whole space.

Equivalent formulations

Local bases in first countability

In a X, the first axiom of countability is equivalently formulated by requiring that every point x \in X possesses a countable local basis, that is, a countable collection \{B_n(x)\}_{n \in \mathbb{N}} of open neighborhoods of x such that for any open neighborhood U of x, there exists some n with B_n(x) \subseteq U. This local basis can often be chosen to be nested, satisfying B_1(x) \supseteq B_2(x) \supseteq \cdots, which simplifies constructions in proofs. The presence of such a countable local basis at each point enables a sequential characterization of topological notions that typically require nets or filters in more general spaces. Specifically, in a first-countable space, convergence can be described using sequences rather than nets: a net converges to a point if and only if there is a cofinal subsequence that converges to it as a sequence. This simplification arises because the countable basis allows one to extract a sequence from any convergent net by selecting points in successively smaller basis elements. A key in this context states that a X is first-countable , for every A \subseteq X, the \overline{A} equals A the set of all limits of convergent s in A. Equivalently, a point x lies in \overline{A} there exists a \{x_n\} \subseteq A such that x_n \to x. Thus, a F \subseteq X is closed it contains all limits of convergent sequences from F. This sequential criterion for closed sets holds precisely due to the countable local bases, which facilitate the construction of sequences witnessing limit points. Such formulations are central in , as explored in standard texts like Munkres' Topology. Second-countable spaces satisfy first countability, since a countable basis for the entire restricts to a countable local basis at each point.

Countable bases in second countability

The second axiom of countability, or second-countability, admits an equivalent formulation in terms of subbases. A subbasis for a X is a collection \mathcal{S} of open subsets whose finite intersections form a basis for the on X. A space satisfies the second axiom of countability it has a countable subbasis. To see this equivalence, note that any countable basis \mathcal{B} = \{B_n \mid n \in \mathbb{N}\} serves directly as a countable subbasis, as the collection of all finite intersections of elements from \mathcal{B} is countable and forms a basis for the , and \mathcal{B} generates the via arbitrary unions. Conversely, if \mathcal{S} = \{S_n \mid n \in \mathbb{N}\} is a countable subbasis, then the collection of all finite intersections \bigcap_{i=1}^k S_{n_i} (for k \in \mathbb{N}, n_i \in \mathbb{N}) is countable, as there are countably many finite subsets of \mathbb{N}, and this collection forms a basis for the . Second-countability also imposes restrictions on the cardinality of the topology itself. In a second-countable space with countable basis \mathcal{B}, every is a of at most countably many elements from \mathcal{B}, so the collection of all possible such unions has at most $2^{\aleph_0}, the . Thus, the has at most continuum many s.

Properties and implications

Implications between axioms

A second-countable topological space is always first-countable. To see this, let (X, \tau) be a with countable basis \{B_n \mid n \in \mathbb{N}\}. For any point x \in X, the collection \{B_n \mid x \in B_n\} forms a countable local basis at x, since for any open neighborhood U of x, there exists some B_k such that x \in B_k \subset U, ensuring the collection satisfies the local basis property. The converse does not hold: first-countability does not imply second-countability. An equipped with the provides a , as the sets form a countable local basis at each point, making the space first-countable, but any basis for the must include all singletons, which is uncountable. The countability axioms are defined for general topological spaces and hold independently of separation properties such as T1 or Hausdorffness, though they are frequently studied in or Hausdorff contexts to ensure desirable behaviors like metrizability. As a consequence, every is Lindelöf, meaning that every open cover admits a countable subcover.

Connections to compactness and metrizability

A compact is metrizable if and only if it is second-countable. This equivalence arises as a special case of Urysohn's metrization theorem, which states that every second-countable is metrizable, combined with the fact that compact s are (and hence ). In a compact second-countable space, the countable basis forms an open cover, and by compactness, there exists a finite subcollection of these basis elements that covers the entire space. This property underscores the interplay between countability and compactness, ensuring that such spaces are metrizable and exhibit controlled topological complexity. Second-countable spaces play a key role in dimension theory, particularly for finite-dimensional cases like topological manifolds, which are defined to be second-countable, Hausdorff, and locally Euclidean of dimension n, thereby possessing inductive dimension exactly n. While second-countable spaces can have infinite inductive dimension—for instance, the countable product \mathbb{R}^\mathbb{N} does so—the emphasis in applications often falls on finite-dimensional structures where second countability ensures manageable covering properties. In metric spaces, second countability is equivalent to separability (having a countable dense ) and to the Lindelöf property (every open cover admits a countable subcover). Additionally, in first-countable spaces, compactness implies sequential compactness.

Examples

Spaces satisfying both axioms

Euclidean spaces provide a fundamental example of topological spaces satisfying both the first and second axioms of countability. The space \mathbb{R}^n equipped with the standard possesses a countable basis consisting of open rectangles formed by products of open intervals with rational endpoints, which establishes second countability; first countability follows as a consequence of second countability in general topological spaces. Separable metric spaces also satisfy both axioms. In a separable , the existence of a countable dense allows the construction of a countable basis by taking open balls of rational radii centered at points of the dense , thereby proving second countability; spaces are inherently first countable due to the countable collection of balls of rational radii around each point. Finite-dimensional smooth manifolds are defined to be second countable. By construction, a smooth manifold is a second-countable that is locally , with the countable basis arising from the countable collection of domains; the local structure ensures first countability at each point. The , defined as the product \prod_{n=1}^\infty \left[0, \frac{1}{n}\right] with the , is a compact that is separable, hence second countable. Its first countability follows from the metric structure. Spaces satisfying both axioms, such as these examples, are Lindelöf, meaning every open cover admits a countable subcover.

Spaces satisfying one but not the other

While every is first-countable, the converse does not hold, as there exist topological spaces that satisfy the first axiom of countability but fail the second. This implication follows from the fact that a countable basis for the entire can be used to construct a countable local basis at each point by selecting those basis elements containing the point and shrinking them appropriately using the continuity of the . A classic example of a first-countable space that is not second-countable is the discrete on an X. In this topology, every subset is open, so the \{x\} forms a countable (in fact, finite) local basis at each point x \in X, satisfying first countability. However, any basis for the must include all singletons to generate the open sets, requiring an uncountable collection, so the space lacks a countable basis. Another example is the Sorgenfrey line, which is the real line \mathbb{R} equipped with the lower limit topology generated by the half-open intervals [a, b) for a < b. At each point x \in \mathbb{R}, the collection \{[x, x + \frac{1}{n}) \mid n \in \mathbb{N}\} serves as a countable local basis, making the space first-countable. Yet, it has no countable basis for the entire topology, as any countable collection of basis elements can only "cover" countably many such intervals in a way that fails to generate all open sets, necessitating an uncountable basis. The Moore plane, also known as the Niemytzki plane, provides a further : it consists of the closed upper half-plane \mathbb{R}^2_{\geq 0} where points above the x-axis have the usual Euclidean neighborhoods, while points on the x-axis have neighborhoods consisting of tangent open discs above the axis. Each point has a countable local basis—Euclidean balls for interior points and sequences of shrinking tangent discs for boundary points—ensuring first countability. Nevertheless, the space is not second-countable, as the boundary points require uncountably many distinct tangent disc bases to separate them properly. These examples are typically non-metrizable, highlighting how first countability alone does not guarantee the stronger uniformity of second countability.

Lindelöf property

The is a covering axiom in stating that a is Lindelöf if every open cover of the space admits a countable subcover. This condition ensures a form of "countable refinement" for open covers, weaker than but stronger than certain other global properties in non-metrizable settings. Second-countable spaces, which possess a countable basis for their , are always Lindelöf, as any open cover can be refined to a countable subcollection from the basis. However, the converse does not hold: there exist Lindelöf spaces that lack a countable basis. A classic example is the Sorgenfrey line, the real line equipped with the generated by half-open intervals [a, b); this space is hereditarily Lindelöf and separable but not second-countable, as any countable collection of basis elements fails to generate all open sets. In the context of separation axioms, every regular Lindelöf space is , meaning disjoint closed sets can be separated by disjoint open sets. This result ties the Lindelöf to countability through its with regularity, and since second-countability implies both regularity preservation and Lindelöf, it strengthens such guarantees in countable-basis settings.

Separability

In topology, a topological space X is said to be separable if it contains a countable dense subset, meaning there exists a countable set D \subseteq X such that every non-empty open subset of X intersects D. Every second-countable topological space is separable. To see this, let \mathcal{B} = \{B_n : n \in \mathbb{N}\} be a countable basis for the topology on X. For each non-empty B_n, select a point x_n \in B_n. The set D = \{x_n : n \in \mathbb{N}\} is countable and dense in X, since any non-empty open set U \subseteq X contains some basis element B_k, and thus intersects D at x_k. In the context of metric spaces, separability and second countability are equivalent properties. Specifically, a metric space is separable if and only if it is second-countable. This equivalence highlights the close relationship between the existence of a countable dense subset and the existence of a countable basis in metric topologies. A classic example is the real line \mathbb{R} with the standard Euclidean topology, which is separable because the set of rational numbers \mathbb{Q} is countable and dense in \mathbb{R}; every non-empty open interval in \mathbb{R} contains rational points. Consequently, \mathbb{R} is also second-countable, with the collection of open intervals with rational endpoints forming a countable basis.

References

  1. [1]
    [PDF] Section 30. The Countability Axioms
    Dec 8, 2016 · Definition. If a topological space X has a countable basis for its topology, then. X satisfies the Second Countability Axiom, or is second- ...
  2. [2]
    [PDF] 4. Countability
    First, take care to note that the definition does not say that every basis of the topology is countable, just that there is at least one countable basis. In ...
  3. [3]
    [PDF] (1) The first countability axiom. (2) The second
    We have studied four basic countability properties: (1) The first countability axiom. (2) The second countability axiom,. (3) The Lindel6f condition.
  4. [4]
    First axiom of countability - Encyclopedia of Mathematics
    Feb 7, 2011 · A concept in set-theoretic topology. A topological space satisfies the first axiom of countability if the defining system of neighbourhoods ...
  5. [5]
    [PDF] 3 COUNTABILITY AND CONNECTEDNESS AXIOMS
    X satisfies the first axiom of countability or is first countable iff for each x ∈ X, there is a countable neighbourhood basis at x. Any metrisable space ...
  6. [6]
    [PDF] Metric & Topological Spaces - alexwlchan
    A topological space (X, τ) is called secound countable if it has a countable base of open sets. Clearly, second countable implies first countable. Consider: for ...
  7. [7]
    [PDF] Topology (H) Lecture 13
    Apr 22, 2021 · In topology, an axiom of count- ability is a topological property that asserts the existence of a countable set with certain properties. There ...
  8. [8]
    [PDF] 5. Sequences, weak T-axioms, and first countability
    Definition 5.3. A topological space (X,Τ ) is said to be first countable if every point in X has a countable local basis. This is the key property that allows ...
  9. [9]
    [PDF] Class Notes for Math 871: General Topology, Instructor Jamie ...
    If X has a countable subbasis, then X is 2nd countable. Proof. If S is a countable subbasis then B = {S1 ∩ททท∩ Sn|n ≥ 0,Si ∈ S} is a countable basis.
  10. [10]
    [PDF] DESCRIPTIVE SET THEORY PROBLEM SET 1. Let X be a second ...
    Let X be a second-countable topological space. (a) Show that X has at most continuum many open subsets. (b) Let α,β,γ denote ordinals.
  11. [11]
    [PDF] Review of point-set topology Andrew Putman - Academic Web
    A space X is second countable if there is a countable basis for its topology. It is clear that all second countable spaces are first countable. It is not ...<|separator|>
  12. [12]
    [PDF] INTRODUCTION TO TOPOLOGY Contents 1. Basic concepts 1 2 ...
    A topological space (X, τ) is called first countable, if every point x ∈ X has a countable neighbourhood basis. The topological space (X, τ) is second countable ...
  13. [13]
    [PDF] 1 Hausdorff spaces - Math 535 - General Topology Additional notes
    Sep 14, 2012 · For example, consider X an uncountable set endowed with the discrete topology. Then X is first-countable but not second-countable. Remark 2.13.
  14. [14]
    [PDF] Chapter 4: Topological Spaces - UC Davis Mathematics
    The discrete topology is first countable, and if X is countable, then it is second countable. It is often useful to define a topology in terms of a base. T h ...
  15. [15]
    Lindelöf space - PlanetMath
    Mar 22, 2013 · A second-countable space is Lindelöf. A compact space is Lindelöf. A regular (http://planetmath.org/T3Space) ...
  16. [16]
    Second-countable implies Lindelof - Topospaces
    Jul 20, 2008 · This article gives the statement and possibly, proof, of an implication relation between two topological space properties.
  17. [17]
    [PDF] Normal Spaces, Regular Spaces, Urysohn metrization - UTK Math
    Urysohn metrization theorem. Let X be Hausdorff and second-countable. Then X is metrizable if and only if X is normal. The following lemma used in the proof ...
  18. [18]
    [PDF] Chapter 1: Smooth manifolds - Arizona Math
    Sep 2, 2010 · A topological manifold is Hausdorff, second countable, and locally Euclidean of dimension n. A manifold of dimension n is also called an n- ...
  19. [19]
    Sets of infinite Hausdorff dimension in a second countable metric ...
    Jul 18, 2020 · The infinite power RN is separable metrizable (hence second countable) and has infinite inductive dimension. This is a topologically defined ...finite dimensional compact Hausdorff space, not second countableIs second-countability a topological invariant? - Math Stack ExchangeMore results from math.stackexchange.com
  20. [20]
    Separability of a Metric Space Is Equivalent to the Existence of a ...
    Then the following conditions are equivalent: X is separable;. X is second countable, i.e., it has a countable open base;. X is Lindelöf ...
  21. [21]
    [PDF] Compactness, countability, function spaces: examples - UTK Math
    If X is locally compact, σ-compact (for instance, a topological manifold), then Cc(X, M) is second-countable if M is separable metric. For each i ≥ 1, let Bi be ...
  22. [22]
    [PDF] Lecture Notes on Topology for MAT3500/4500 following J. R. ... - UiO
    Nov 29, 2010 · A space having a countable basis at each of its points is said to satisfy the first countability axiom, or to be first-countable. Every ...
  23. [23]
    [PDF] Section 9.6. Separable Metric Spaces
    Dec 10, 2022 · A metric space X is separable provided there is a countable subset of X that is dense in X. Note 9.6.A. By the definition of “point of closure,” ...
  24. [24]
  25. [25]
    Hilbert Cube - an overview | ScienceDirect Topics
    H ω is a separable compact metric space, hence a second countable compactum and a Polish space. H ω is called the Hilbert cube. It is a compact subset of the ...
  26. [26]
    [PDF] Chapter III Topological Spaces
    b) Use part a) to prove that the Sorgenfrey line is not second countable. (Hint: Show that otherwise there would be a countable base of sets of the form but.
  27. [27]
    [PDF] MATH 561 - HOMEWORK 2
    Thus every point in the Moore plane has a countable local basis, and so the Moore plane is first-countable. However, the Moore plane is not second-countable.
  28. [28]
    [PDF] 10. Covering properties - TU Graz
    Observe that the Sorgenfrey line is hereditarily Lindelöf but not second countable. For metric spaces, however, we have the following important result. Theorem.
  29. [29]
    [PDF] A Comparison of Lindelof-type Covering Properties of Topological ...
    2.1 Definition. A topological space X is called Lindelöf if every open cover of X has a countable subcover. 2.2 Examples. The following are straightforward ...
  30. [30]
    Separable Space -- from Wolfram MathWorld
    A topological space having a countable dense subset. An example is the Euclidean space R^n with the Euclidean topology.
  31. [31]
    separable space - PlanetMath.org
    Mar 22, 2013 · All second-countable spaces are separable. A metric space is separable if and only if it is second-countable.
  32. [32]
    every second countable space is separable - PlanetMath
    Mar 22, 2013 · Let X X be a second countable space and let B ℬ be a countable base. For every non-empty set B B in B ℬ , choose a point xB∈B x B ∈ B . The ...