Fact-checked by Grok 2 weeks ago

Normal space

In , a normal space is a X in which, for any two disjoint closed subsets A and B of X, there exist disjoint open subsets U and V of X such that A \subseteq U and B \subseteq V. This separation property strengthens the axioms for and is often denoted as the T_4 when combined with the T_1 condition (where singletons are closed sets, equivalent to the Hausdorff property). The definition typically assumes T_1, ensuring that points are closed and that the space distinguishes distinct points with disjoint open neighborhoods. Normal spaces form a key part of the hierarchy of separation axioms in , where every normal space is (disjoint closed sets and points can be separated by open sets) and every is Hausdorff, with these inclusions being proper. A fundamental consequence is , which states that in a normal space, for any disjoint closed sets A and B, there exists a f: X \to [0,1] such that f(A) = \{0\} and f(B) = \{1\}, enabling the embedding of such spaces into metric-like structures under additional conditions. The further extends this, allowing continuous real-valued functions defined on closed subspaces to be extended to the entire space. Notable examples of normal spaces include all metric spaces equipped with their standard topology, as the distance function allows separation of closed sets by open balls. Compact Hausdorff spaces are also , as are well-ordered sets under the and regular second-countable spaces. However, is not preserved under arbitrary products or subspaces; for instance, the uncountable product of intervals (0,1)^J for uncountable J is regular but not , and the Sorgenfrey plane (product of Sorgenfrey lines) fails despite each factor being . These properties make normal spaces essential in metrization theorems, such as the Urysohn metrization theorem, which characterizes second-countable T_1 spaces as metrizable.

Core Definitions

Formal Definition

In topology, a closed set in a topological space is the complement of an open set. A topological space X is called normal if, given any two disjoint closed subsets A and B of X, there exist disjoint open subsets U and V of X such that A \subseteq U and B \subseteq V. A normal space that is also T_1—meaning that it satisfies the where sets are closed, allowing points to be separated from closed sets by open neighborhoods—is denoted a T_4 space.

Equivalent Formulations

In T_1 spaces, every normal is , with normality extending the separation property from points and disjoint closed sets to arbitrary pairs of disjoint closed sets. An alternative characterization states that a topological space is normal if and only if, for every closed set E and every open set U containing E, there exists an open set V such that E \subseteq V \subseteq \overline{V} \subseteq U. Normality can also be defined in terms of function extension: a space is normal if and only if every continuous real-valued function defined on a closed subset extends to a continuous function on the entire space. This is known as Tietze's extension theorem. Urysohn's lemma provides another equivalent formulation: a space is normal if and only if, for any two disjoint closed sets E and F, there exists a continuous function f: X \to [0,1] such that f(E) = \{0\} and f(F) = \{1\}. The concept of normal spaces was introduced by Felix Hausdorff in 1914, playing a pivotal role in the early development of general topology by formalizing higher separation axioms beyond Hausdorff spaces.

Examples

Normal Spaces

Metric spaces provide a fundamental class of normal topological spaces. In a metric space (X, d), any two disjoint closed sets A and B can be separated by open sets constructed using open balls of radius equal to half the infimum of the distances between points in A and B, ensuring the openness and disjointness required by the axiom. This property holds because the metric induces a topology where such balls form a basis, allowing precise control over neighborhoods around closed sets. Compact s also satisfy . Every compact Hausdorff space is normal, as the compactness ensures that closed subsets are compact and can be separated using the Hausdorff property combined with finite subcovers to construct disjoint open neighborhoods. further supports this by guaranteeing that products of compact Hausdorff spaces remain compact and Hausdorff, preserving normality. Euclidean spaces \mathbb{R}^n exemplify spaces through their structure, inheriting the separation properties of spaces directly from the . Similarly, topological manifolds, being locally and Hausdorff with a second-countable , are due to their metrizable local charts that extend globally via the manifold's structure. Finite products of spaces are in the . For instance, the product of two spaces inherits the because projections allow lifting separations from each factor to disjoint open sets in the product. Infinite products under the Tychonoff can be if the factors are compact, as the resulting space is compact Hausdorff by . A example is the unit interval [0,1] with the standard , which is compact and Hausdorff, hence . This space satisfies the T4 , combining with the T1 property inherent to Hausdorff spaces.

Non-Normal Spaces

The Niemytzki plane, also known as the plane, is defined as the upper half-plane \mathbb{R} \times [0, \infty) equipped with a topology where the basis consists of all open disks in the upper half-plane for points with positive y-coordinate, and for points on the x-axis (y=0), the basis elements are open disks tangent to the x-axis at that point and lying entirely in the upper half-plane. This space is Hausdorff and completely regular but fails to be normal. Specifically, let A = \mathbb{Q} \times \{0\} be the set of points on the x-axis with rational x-coordinates, and B = (\mathbb{R} \setminus \mathbb{Q}) \times \{0\} the points with irrational x-coordinates; both A and B are closed and disjoint in the Niemytzki plane, yet there do not exist disjoint open sets containing them, as any open neighborhood of a point in A intersects every open neighborhood of nearby points in B due to the tangent disk basis. The Sorgenfrey line is the real line \mathbb{R} with the , generated by basis elements [a, b) for a < b. This space is normal, hereditarily Lindelöf, and paracompact, but its product with itself, the Sorgenfrey plane \mathbb{R}_l \times \mathbb{R}_l, is not normal. The failure arises from the sets P = \{(p, -p) \mid p \in \mathbb{R} \setminus \mathbb{Q}\} (the anti-diagonal over irrationals) and Q = \{(q, -q) \mid q \in \mathbb{Q}\}&#36; (over rationals), which are both closed and disjoint; however, they cannot be separated by disjoint open sets because any basic open neighborhood in the product topology around points in PandQ$ will overlap due to the half-open intervals aligning along the anti-diagonal. The Tychonoff plank is the product space ([0, \omega_1] \times [0, \omega]) equipped with the , where \omega_1 is the first uncountable ordinal and \omega is the first infinite ordinal. The deleted Tychonoff plank, obtained by removing the point (\omega_1, \omega), is completely regular but not . In this space, the sets C = \{\omega_1\} \times [0, \omega) and D = [0, \omega_1) \times \{\omega\} are closed and disjoint, but no disjoint open sets separate them, as any open neighborhood of C must include points arbitrarily close to (\omega_1, \omega) from below in the second coordinate, which inevitably intersects neighborhoods of D near the deleted point. These examples illustrate pathologies that do not occur in metric spaces, which are always .

Key Properties

Separation Properties

A fundamental separation property of normal spaces is encapsulated in , which states that if X is a normal space and A, B \subseteq X are disjoint closed sets, then there exists a f: X \to [0,1] such that f(A) = \{0\} and f(B) = \{1\}. Intuitively, this lemma provides a continuous "separator" that distinguishes the two closed sets by mapping one to the zero level and the other to the unit level, with intermediate values ensuring continuity across the space; it relies on the T1 condition to ensure singletons are closed, making the property hold specifically for T4 spaces (normal and T1). The extends this separability to real-valued functions: in a normal space X, if A \subseteq X is closed and g: A \to \mathbb{R} is continuous, then there exists a continuous extension G: X \to \mathbb{R} such that G|_A = g. This theorem underscores the flexibility of normal spaces in extending local continuous data globally while preserving the function's boundedness if the original is bounded. In paracompact T4 spaces, every open cover admits a locally finite open refinement to which is subordinate, consisting of continuous functions \{\phi_i\} such that each \phi_i \geq 0, \sum \phi_i = 1, and \operatorname{supp}(\phi_i) is contained in the corresponding refinement set. Compact Hausdorff spaces, being normal, are paracompact, meaning every open cover has a locally finite open refinement. Moreover, that is also paracompact satisfies metrizability under additional conditions, such as possessing a countable basis.

Embedding and Extension Theorems

Normal spaces, being completely regular, admit a canonical embedding into a product of closed intervals. Specifically, every space X is homeomorphic to a of the Tychonoff cube [0,1]^I for some I, where the is constructed using a family of continuous functions from X to [0,1] that separate points from closed sets. This theorem, originally established by Tychonoff, highlights the functional richness of normal spaces and facilitates their study within the broader class of completely regular spaces. A key extension theorem for normal spaces arises from their complete regularity, enabling the construction of the \beta X. For any normal space X, \beta X is a compact into which X embeds densely as a , and every continuous bounded real-valued on X extends uniquely to a on \beta X. This compactification, developed independently by Stone and Čech, preserves the topological structure of X while extending it to a compact setting, making it invaluable for analyzing limits and extensions in normal spaces. Metrization theorems provide conditions under which spaces admit a compatible , thereby embedding them into spaces. The Urysohn metrization theorem states that every second-countable space is metrizable, as implies the required regularity and Hausdorff properties. This result, due to Urysohn, is particularly powerful for spaces with countable bases, such as manifolds or separable spaces, allowing the transfer of tools like completeness and . For compact spaces, metrizability follows under second countability, strengthening the theorem's applicability in bounded settings where compactness ensures . The metrization theorem extends these ideas to a broader class of normal spaces without assuming second countability. It asserts that a normal space is metrizable it has a basis, meaning a basis that is a countable of discrete families of open sets. Named after , this theorem characterizes metrizability through base conditions that align with normality's separation capabilities, enabling metrization for certain locally compact or paracompact normal spaces beyond the second-countable case.

Relations to Separation Axioms

Comparisons with T₀ to T₃ Axioms

In , the separation axioms T₀ through T₄ form a of increasingly stringent conditions on how well points and sets can be distinguished using open sets in a . A normal space, often denoted as satisfying T₄ when combined with T₁, implies all weaker axioms in this chain, providing a structured progression from basic point separation to the separation of disjoint closed sets. The weakest axiom, T₀ (also known as Kolmogorov quotient), requires that for any two distinct points x and y in the , there exists an open set containing one but not the other. This ensures a minimal level of distinguishability between points. Every normal satisfies T₀, as the stronger separation properties guarantee such open sets exist. Next, T₁ (Fréchet ) strengthens T₀ by requiring that every set \{x\} is closed, which is equivalent to the condition that for distinct points x and y, there are open sets containing x but not y, and vice versa. requires T₁ for the T₄ designation, as the definition of T₄ explicitly includes T₁ alongside the normality condition for separating disjoint closed sets. T₂ (Hausdorff space) further refines separation by demanding that any two distinct points have disjoint open neighborhoods. A normal space implies T₂ provided it also satisfies T₁, since the ability to separate points from closed sets and closed sets from each other cascades to point-point separation. T₃ (regular Hausdorff space) combines regularity—where a point and a disjoint have disjoint open neighborhoods—with T₁ (or sometimes T₀ in alternative conventions). A T₄ space is precisely a T₃ space augmented by the condition that any two disjoint can be separated by disjoint open neighborhoods. The implications form a strict chain: T₄ ⇒ T₃ ⇒ T₂ ⇒ T₁ ⇒ T₀, where each step adds a layer of separation power, though the reverse implications do not hold. For instance, there exist spaces that are T₃ but not T₄, satisfying point-closed set separation yet failing to separate certain pairs of disjoint closed sets, and similarly for weaker levels relative to T₄.

Stronger Notions of Normality

A completely normal space is defined as a where every pair of separated sets possesses disjoint open neighborhoods. Two sets A and B are separated if A \cap \overline{B} = \emptyset and B \cap \overline{A} = \emptyset. This condition strengthens the by ensuring separation not just for disjoint closed sets but for sets that are mutually disjoint from each other's closures. Completely normal spaces imply normality, as disjoint closed sets are a special case of , but the converse does not hold. For instance, all spaces are completely normal, since subspaces of spaces are metrizable and thus normal. In contrast, the Tychonoff plank—defined as ([0, \omega_1] \times [0, \omega]) with the product —is normal but not completely normal, as the deleted Tychonoff plank (removing the point (\omega_1, \omega)) is a non-normal subspace. Equivalently, the subspace consisting of the "residual boundaries" \{ \omega_1 \} \times [0, \omega) \cup ([0, \omega_1) \times \{ \omega \} fails to be normal in the subspace topology, since the components \{ \omega_1 \} \times [0, \omega) and ([0, \omega_1) \times \{ \omega \} are disjoint closed sets that cannot be separated by disjoint open sets. Hereditarily normal spaces provide an equivalent formulation to complete : a space is hereditarily normal if every , endowed with the , is . This equivalence arises because the separation property for ensures that induced topologies on arbitrary subspaces preserve . spaces exemplify hereditarily spaces, inheriting this property from their metrizability. The Tychonoff plank serves as a , being yet containing a non- as noted above. Monotonically spaces extend in spaces with order structure, requiring a separation for disjoint closed sets. Specifically, for disjoint closed sets A and B, there exists a function U assigning to each such pair an U(A, B) containing A and disjoint from B, such that if A' \subseteq A and B' \subseteq B are disjoint closed sets, then U(A', B') \subseteq U(A, B). This monotonicity ensures consistent refinement of neighborhoods. Every generalized ordered space (GO-space), including linearly ordered topological spaces, is monotonically , highlighting its relevance to ordered topologies. Collectionwise normal spaces strengthen normality by handling families of closed sets simultaneously. A T_1 space is collectionwise normal if, for every discrete collection \{F_i\}_{i \in I} of closed sets, there exists a discrete collection \{U_i\}_{i \in I} of open sets such that F_i \subseteq U_i for each i and the U_i are pairwise disjoint. This property exceeds mere pairwise separation and is crucial in dimension theory, where it facilitates inductive constructions for embedding spaces into spaces. Examples include all compact spaces, while certain planes provide counterexamples to weaker forms. Completely normal spaces that are also Lindelöf are paracompact, as the Lindelöf property combined with (which complete normality implies under T_1) yields a locally finite open refinement for every open cover via standard theorems on regular Lindelöf spaces.

References

  1. [1]
    Normal Space -- from Wolfram MathWorld
    A normal space is a topological space in which for any two disjoint closed sets C,D there are two disjoint open sets U and V such that C subset= U and D subset ...
  2. [2]
    normal space in nLab
    Oct 19, 2021 · A normal space is a space (typically a topological space) which satisfies one of the stronger separation axioms.
  3. [3]
    [PDF] Section 31. The Separation Axioms
    Aug 30, 2016 · So RK is an example of a nonregular Hausdorff space showing that the set of regular spaces is a proper subset of the set of Hausdorff spaces.
  4. [4]
    [PDF] Section 32. Normal Spaces
    Aug 20, 2016 · Every regular space with a countable basis is normal. Note. The following result shows that every metrizable space is normal, so that we can ...
  5. [5]
    [PDF] Normal Spaces, Regular Spaces, Urysohn metrization. Definition.
    NORMAL SPACES, REGULAR SPACES, URYSOHN METRIZATION. Definition. A topological space X is normal if it is Hausdorff and open sets separate disjoint closed sets: ...
  6. [6]
    Separation Axioms -- from Wolfram MathWorld
    A list of five properties of a topological space X expressing how rich the population of open sets is.<|control11|><|separator|>
  7. [7]
  8. [8]
    [PDF] Chapter 7 Separation Properties
    There is another famous characterization of normal spaces in terms of . It is a result about. GР\С. “extending” continuous real-valued functions defined on ...
  9. [9]
    Grundzüge der Mengenlehre : Hausdorff, Felix, 1868-1942
    Dec 2, 2008 · Grundzüge der Mengenlehre ; Publication date: 1914 ; Topics: Set theory ; Publisher: Leipzig Viet ; Collection: gerstein; toronto; ...Missing: normal spaces
  10. [10]
    [PDF] 12. Metric spaces and metrizability
    Exactly the same proof shows that every metrizable space is normal. The ... Let (X, d) be a metric space (just a metric space, no topology needed for this.
  11. [11]
    [PDF] 4 COMPACTNESS AXIOMS
    Theorem 4.7 Every compact Hausdorff space is normal. Proof. Let A and B be disjoint closed subsets of the compact Hausdorff space X. Then A and B are compact.
  12. [12]
    [PDF] Manifolds.pdf - University of South Carolina
    Abstract. A manifold is a connected Hausdorff space in which every point has a neighborhood homeomorphic to Euclidean n-space (n is unique).<|separator|>
  13. [13]
    [PDF] Topology - The Research Repository @ WVU
    (i) The product of two regular spaces is a regular space. (ii) The product of two normal spaces is a normal space. Ex. 2.
  14. [14]
    Another Proof that the Niemytzki Plane is Not Normal - jstor
    The proof is simple, uses only the nested set theorem, and works just as well for the Sorgenfrey plane. The Niemytzki plane can be envisioned by thinking of hot ...Missing: non- | Show results with:non-
  15. [15]
    [PDF] Some Properties of the Sorgenfrey Line and the Sorgenfrey Plane
    plane is not normal, and hence the product of two normal spaces need not be normal. The proof that the Sorgenfrey plane is not normal and many of the lemmas ...
  16. [16]
    [PDF] The Tychonoff Plank - G Eric Moorhouse
    our proof that X = {x ... Not every subspace of a normal space is normal. The standard counterexample for demonstrating this is the Punctured Tychonoff Plank.
  17. [17]
    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: PS zusammenfüngenden
  18. [18]
    An extension of Tietze's theorem. - Project Euclid
    An extension of Tietze's theorem. J. Dugundji. Download PDF + Save to My Library. Pacific J. Math. 1(3): 353-367 (1951).
  19. [19]
    [PDF] Partitions of unity - webspace.science.uu.nl
    In a normal space X, if A ⊂ U ⊂ X with A-closed and U-open in X, then there exists an open V in X such that A ⊂ V ⊂ V ⊂ U. Proof. Since A ⊂ U, A and X ...
  20. [20]
    Completely Regular Space -- from Wolfram MathWorld
    In any case, every completely regular space is regular, and the converse is not true. See also. Completely Regular Graph, Tychonoff Space. This entry ...
  21. [21]
    [PDF] 9. Stronger separation axioms
    Munkres does not give a name to the property we call “regular”. Counterexamples in Topology, on the other hand, defines “regular” and “T3” in exactly the ...
  22. [22]
    [PDF] MH 7500 THEOREMS Definition. A topological space is an ordered ...
    Definition. A space X is a T0-space if whenever x, y ∈ X with x 6= y, there is an open set containing one of these points but not the other. If there is ...
  23. [23]
    Monotonically Normal Spaces - jstor
    2.2. Lemma. Any monotonically normal space has a monotone normality oper- ator G satisfying G(A, B) n G(B, A) = 0 for any pair (A, B) of disjoint closed. sets.
  24. [24]
    [PDF] Collectionwise Pre-Normality in Topological Spaces
    A space X is called a collectionwise normal space if and only if X is a T1-space and for every discrete family F = {Fs}s∈S of closed subsets of X, there exits a ...
  25. [25]
    [PDF] “The Lindelöf Property” - University of Birmingham
    Feb 25, 2002 · It is an important result that regular Lindelöf spaces are paracompact, from which it follows that they are (collectionwise) normal.