Fact-checked by Grok 2 weeks ago

Well-order

In , particularly in , a well-order is a on a set such that every non-empty of the set has a least under that . This property distinguishes well-orders from other total orders, ensuring a form of "no infinite descending chains" in the structure. Well-orders are fundamental to the theory of ordinal numbers, where each well-ordered set is order-isomorphic to a unique ordinal, representing its order type. For example, the natural numbers equipped with the standard less-than relation form a well-order, as do all finite totally ordered sets and the even natural numbers under the usual order. In contrast, the integers or rational numbers with their standard orders are not well-ordered, since subsets like the negative integers lack a least element. The asserts that every non-empty set can be well-ordered, though constructing an explicit well-ordering often requires the and may not be intuitive for uncountable sets like the real numbers. This theorem, first proposed by in 1883 and formalized by Zermelo in 1904, is equivalent to the and plays a central role in transfinite arithmetic and proofs involving induction over arbitrary sets. Well-orders also underpin the principle of well-ordering for the natural numbers, which states that every non-empty subset of the naturals has a minimal element and serves as the basis for .

Basic Concepts

Definition

A well-order on a set X is a \leq on X such that every non-empty subset of X has a least element under \leq. A is a on X that is antisymmetric (if a \leq b and b \leq a, then a = b), transitive (if a \leq b and b \leq c, then a \leq c), and total (for any a, b \in X, either a \leq b or b \leq a). The corresponding strict well-order is the irreflexive relation < defined by a < b if and only if a \leq b and a \neq b; this strict relation is transitive and satisfies trichotomy: for any distinct a, b \in X, exactly one of a < b or b < a holds. The concept of a well-order was introduced by Georg Cantor in the context of his development of transfinite ordinal numbers.

Properties

A well-order on a set X is a total order, meaning that for any x, y \in X, exactly one of x < y, x = y, or y < x holds, ensuring the relation is trichotomous and comparable for all pairs. A fundamental property is the absence of infinite descending chains: there exists no infinite sequence (x_n)_{n \in \mathbb{N}} in X such that x_{n+1} < x_n for all n \in \mathbb{N}. Intuitively, this prevents any "endless regression" where elements can keep getting strictly smaller indefinitely, guaranteeing that any decreasing sequence must terminate after finitely many steps. This property implies that every non-empty subset A \subseteq X has a least element m \in A, meaning m \leq a for all a \in A. The least element is unique, as the total order ensures that if two elements both satisfy the condition, one must be smaller than or equal to the other, contradicting the assumption unless they coincide. The relation is well-founded, formally expressed as: for every sequence (x_n)_{n \in \mathbb{N}} in X, there exists some n such that x_{n+k} \geq x_n for all k \geq 0, ensuring no strict descent persists indefinitely. Every element x \in X decomposes X into the initial segment \{y \in X \mid y < x\}, which itself forms a well-ordered set, and the remainder X \setminus \{y \in X \mid y \leq x\}, preserving the structure of the order.

Examples

Natural Numbers

The natural numbers, denoted \mathbb{N} = \{0, 1, 2, \dots \}, are equipped with the standard arithmetic order \leq, where m \leq n if and only if n - m is a non-negative integer. This ordering makes (\mathbb{N}, \leq) a prototypical example of a well-ordered set, as every non-empty subset has a least element under the standard definition of well-ordering. To prove that every non-empty subset S \subseteq \mathbb{N} has a least element, proceed by contradiction. Suppose S has no least element. Define the property P(n) as "n \notin S". Then P(0) holds, since if $0 \in S, it would be the least element. Assume P(k) holds for all k \leq n; then n+1 \notin S, for otherwise n+1 would be the least element. By mathematical induction, P(n) holds for all n \in \mathbb{N}, implying S = \emptyset, a contradiction. Thus, S has a least element. This proof relies on the inductive structure of \mathbb{N} as constructed in axiomatic set theory. The order type of (\mathbb{N}, \leq) is \omega, the smallest infinite ordinal number. In set theory, \mathbb{N} is identified with the set of all finite ordinals, where each natural number n corresponds to the ordinal \{0, 1, \dots, n-1\}, establishing an order-isomorphism between \mathbb{N} and \omega. The well-ordering of \mathbb{N} is logically equivalent to the principle of mathematical induction, which states: If a property P satisfies P(0) and \forall n \in \mathbb{N}, P(n) \implies P(n+1), then P(n) holds for all n \in \mathbb{N}. This equivalence underscores how the absence of infinite descending chains in \mathbb{N} enables inductive proofs across mathematics.

Integers

The integers \mathbb{Z} consist of the set \{\dots, -2, -1, 0, 1, 2, \dots\} equipped with the standard ordering \leq, where for any a, b \in \mathbb{Z}, a \leq b if and only if b - a is a non-negative integer. This ordering is a total order on \mathbb{Z}, meaning any two elements are comparable. However, \mathbb{Z} fails to be a well-order because not every non-empty subset has a least element, as required by the definition of a well-order. A concrete counterexample is the subset of negative integers \{\dots, -3, -2, -1\}. This subset has no least element under \leq, since for any -n where n > 0, the element -(n+1) satisfies -(n+1) < -n. Equivalently, the negative integers form an infinite descending chain \dots > -3 > -2 > -1 with respect to the strict order <, which violates the least element property essential to well-orders. Thus, while totally ordered, \mathbb{Z} under the standard ordering is not a well-order.

Real Numbers

The real numbers \mathbb{R} are equipped with the standard order relation \leq, which is a total order satisfying the properties of an ordered field. This ordering is dense, meaning that for any two distinct real numbers a < b, there exists another real number c such that a < c < b. More specifically, there even exists a rational number r \in \mathbb{Q} between a and b, reflecting the density of the rationals within the reals. Under this standard ordering, the real numbers do not form a well-order because not every nonempty subset has a least element. A clear counterexample is the open interval (0,1), which contains no smallest element: for any x \in (0,1), there exists y = x/2 \in (0,1) such that y < x, and this process can continue indefinitely due to the density of the order. Similarly, the set of negative real numbers (-\infty, 0) lacks a least element, as it is unbounded below and any candidate can be surpassed by a smaller real number. This failure arises fundamentally from the density of the ordering, a property shared even by the countable dense subset \mathbb{Q} of rationals, which also lacks least elements in intervals like (0,1) \cap \mathbb{Q}. Although \mathbb{R} is complete—meaning every nonempty subset bounded below has a greatest lower bound (infimum)—this completeness does not guarantee the existence of a least element (minimum) in every nonempty subset, particularly in open intervals or unbounded-below sets where the infimum is either not achieved or lies outside the subset.

Relation to Set Theory

Ordinal Numbers

In set theory, the order type of a well-ordered set captures its structure up to isomorphism. Specifically, two well-ordered sets (X, \leq) and (Y, \leq') have the same order type if there exists an order-preserving bijection f: X \to Y, meaning f is a bijection such that x \leq x' if and only if f(x) \leq' f(x') for all x, x' \in X. Ordinal numbers serve as canonical representatives for these order types, labeling the isomorphism classes of well-ordered sets. Every well-ordered set (X, \leq) corresponds uniquely to an ordinal number \alpha, denoted by X \cong \alpha, such that there is an order-isomorphism between X and \alpha equipped with the standard membership relation \in. This assignment proceeds by transfinite recursion: the ordinal \alpha is the smallest ordinal greater than all initial segments of X, and for limit ordinals, \alpha = \sup\{\beta + 1 \mid \beta < \alpha\}, where the supremum is taken over the order types of proper initial segments. This uniqueness ensures that ordinals provide a complete classification of well-orders up to isomorphism. Finite ordinals correspond to the natural numbers under the usual ordering, where the ordinal n is the set \{0, 1, \dots, n-1\}. The smallest infinite ordinal is \omega, the order type of the natural numbers \mathbb{N} with the standard order. Successor ordinals like \omega + 1 extend \omega by adjoining a single element greater than all naturals, while \omega \cdot 2 concatenates two copies of \omega. Larger examples include \omega^2, \omega^\omega, and up to \varepsilon_0, the least fixed point of \alpha \mapsto \omega^\alpha. These illustrate the transfinite hierarchy beyond finite counts. Ordinal arithmetic includes operations like addition and multiplication, defined recursively to preserve order types. Ordinal addition \alpha + \beta is the order type of the disjoint union of \alpha and \beta where all elements of \beta follow those of \alpha. In contrast, the Hessenberg sum (or natural sum) \alpha \oplus \beta is commutative and defined by componentwise addition of coefficients in , mimicking finite arithmetic more closely. Similarly, ordinal multiplication \alpha \cdot \beta places \beta copies of \alpha in sequence, while the Hessenberg product \alpha \otimes \beta is also commutative via normal form coefficients. These operations highlight the non-commutative nature of standard ordinal arithmetic. Every ordinal \alpha admits a unique representation in Cantor normal form: \alpha = \omega^{\gamma_n} \cdot k_n + \omega^{\gamma_{n-1}} \cdot k_{n-1} + \dots + \omega^{\gamma_0} \cdot k_0, where \gamma_n > \gamma_{n-1} > \dots > \gamma_0, each k_i is a positive finite ordinal (), and the terms decrease strictly in exponents. This polynomial-like expansion in base \omega provides a way to express and compare ordinals, facilitating computations and proofs of uniqueness. introduced this form to systematize transfinite numbers beyond \omega.

Well-ordering Theorem

The well-ordering theorem states that for every set X, there exists a well-ordering \leq on X, meaning a such that every nonempty subset of X has a least element with respect to \leq. This well-ordering is not necessarily unique and cannot be constructed explicitly in general. proved the theorem in 1904, relying on what is now known as the to establish the existence of such an ordering via a non-constructive argument involving transfinite recursion on the power set of X. The proof addressed ongoing debates in , including Cantor's that every set admits a well-ordering, and it marked a pivotal use of the choice principle to resolve foundational questions about set comparability. A significant of the is that it identifies cardinal numbers with initial ordinals: every cardinal \kappa is the of some well-ordered set with no smaller well-ordered subsets of the same , ensuring that cardinalities like |\mathbb{R}| = 2^{\aleph_0} correspond to specific ordinals in the hierarchy of alephs. This framework supports the surjectivity from the class of ordinals onto the class of cardinals, enabling consistent comparisons across infinite sets. The theorem thus guarantees that sets without natural well-orderings, such as the integers \mathbb{Z} or the real numbers \mathbb{R} under their standard orders, can still be well-ordered through some alternative relation, though no explicit such ordering is provided by the proof itself.

Characterizations

Equivalent Formulations

A well-order on a set can equivalently be characterized as a total order with no infinite descending chains, meaning there does not exist a sequence (x_n)_{n \in \mathbb{N}} such that x_{n+1} < x_n for all n. This formulation emphasizes well-foundedness: every non-empty subset has a minimal element precisely because the absence of descending chains ensures that any chain must terminate at a least element. Another equivalent characterization is the hereditary property: the order induced on every subset is itself a well-order. In a well-ordered set (X, \leq), for any A \subseteq X, the restriction of \leq to A ensures that every non-empty subset of A has a least element with respect to this induced order, preserving the well-ordering structure throughout. Well-orders also admit a structural decomposition into successor and limit elements, mirroring the construction of ordinal numbers. Every element except the minimal one has an immediate predecessor (the maximal element strictly less than it), defining successor elements as \operatorname{succ}(x) = \min\{y \in X \mid y > x\}; limit elements, in contrast, lack an immediate predecessor and are the supremum of all preceding elements. This partition—where the corresponds to an ordinal \alpha with successors at \beta + 1 and limits at suprema of prior segments—provides a recursive way to analyze the order's transfinite extent. Recursively, well-orders can be built starting from the and iteratively adjoining successors or forming suprema of previous well-orders, directly paralleling the construction of ordinals. This transfinite ensures that any well-order is isomorphic to a unique ordinal, with the process terminating at limit stages by taking least upper bounds, thus generating all possible well-order types without gaps. In terms of strict orders, a strict well-order is equivalently an asymmetric, transitive, and well-founded strict , where well-foundedness again precludes infinite descending chains. This view highlights the irreflexive nature of the strict relation <, which inherits totality and the no-descent property from the non-strict version while ensuring no cycles or equivalences beyond equality.

Axiom of Choice Implications

The well-ordering theorem, which asserts that every set can be well-ordered, is equivalent to the (AC) over the Zermelo-Fraenkel axioms of set theory (ZF). In 1904, Ernst Zermelo proved the theorem using AC, by iteratively applying choice functions to select elements from the power set of the given set, constructing a well-ordering via transfinite recursion that builds ordinals corresponding to the set's cardinality. The converse holds because, given a well-ordering of the universe, one can define a choice function on any collection of non-empty sets by selecting the least element in each according to that ordering. Zorn's lemma, stating that every partially ordered set in which every chain has an upper bound contains a maximal element, is also equivalent to AC and thus to the well-ordering theorem. Max August Zorn introduced the lemma in 1935 as a tool for algebraic applications, proving it via AC by well-ordering the set of chains and extending them maximally. The equivalence arises because Zorn's lemma implies the well-ordering theorem: for any set, consider the partial order of its well-orderable subsets ordered by end-extension, apply Zorn's lemma to obtain a maximal such subset, which must be the entire set. The well-ordering theorem, via AC, has key implications for other results in set theory. It enables transfinite induction on any well-ordered set, allowing proofs by assuming the property holds for all predecessors and verifying the successor and limit cases, which is essential for constructing transfinite sequences and hierarchies. Independently of AC, Hartogs' theorem (1915) establishes that for any set A, there exists an ordinal \alpha (the Hartogs number of A) such that no injection from \alpha into A exists, providing a well-orderable cardinal strictly larger than any well-orderable subset of A; this bounds the size of well-orderings without requiring full AC. In ZF without AC, the well-ordering theorem fails: some sets, such as the real numbers \mathbb{R}, may not admit a well-ordering. Kurt Gödel (1938) showed the consistency of ZF + AC using the constructible universe L, where all sets are well-orderable. Paul Cohen (1963) proved the independence of AC from ZF via forcing, constructing models where \mathbb{R} has no well-ordering, as the continuum becomes incomparable to any aleph without choice principles. Zermelo's 1904 proof, which implicitly relied on AC before its explicit formulation, ignited intense debate among mathematicians like Henri Poincaré and Felix Hausdorff, who questioned the legitimacy of non-constructive choices in infinite settings, viewing it as an unjustified extension beyond finitary methods. This controversy highlighted AC's foundational role and led to its formal inclusion in Zermelo's 1908 axiomatization of set theory.

Topological Aspects

Order Topology

The order topology on a well-ordered set X is generated by taking as a subbasis the collection of all open rays of the form (a, \to) = \{ y \in X \mid y > a \} for each a \in X and all initial segments of the form (\leftarrow, b) = \{ y \in X \mid y < b \} for each b \in X. A basis for this topology consists of finite intersections of these subbasis elements, which yield open intervals (a, b) = \{ y \in X \mid a < y < b \} when both rays are present, along with the rays themselves. In the context of ordinal numbers, which form the canonical well-ordered sets, the basis elements reflect the structure of successors and limits. For a successor ordinal \beta + 1, the \{\beta\} is open in the order topology, as it equals (\beta - 1, \beta + 1) if \beta has an immediate predecessor, making such points isolated. For a limit ordinal \lambda, every neighborhood of \lambda contains a cofinal tail approaching \lambda from below, such as (\alpha, \lambda) for some \alpha < \lambda, since there is no immediate predecessor. The order topology on a well-ordered set is always Hausdorff, as for any distinct points a < b in X, the open sets (\leftarrow, b) and (a, \to) separate them. Regarding compactness, the ordinal space [0, \alpha) in the order topology is compact if and only if \alpha is a successor ordinal. For example, [0, \omega + 1) is compact, as it is a successor ordinal space and any open cover admits a finite subcover. In contrast, for the limit ordinal \alpha = \omega_1, the cover \{ [0, \beta) \mid \beta < \omega_1 \} has no finite subcover, as any finite collection leaves an uncovered tail.

Topological Properties

The order topology on a well-ordered set exhibits several distinctive topological properties, particularly in the case of ordinal spaces. These spaces are locally compact, meaning every point has a compact neighborhood. For an ordinal space [0, \alpha), where \alpha is a limit ordinal, each point \beta < \alpha admits the compact interval [0, \beta + 1] as a neighborhood, ensuring local compactness throughout. However, the entire space [0, \alpha) is not compact unless \alpha is finite or a successor ordinal, as infinite limit ordinals lead to non-compactness due to the absence of a finite subcover for the open cover of initial segments. Well-order topologies are scattered spaces, characterized by the absence of dense-in-itself subsets; every nonempty subspace contains an isolated point, namely its least element. This property follows directly from the well-ordering, where the minimal element of any subset is open in the induced topology. The Cantor-Bendixson rank of such a space, defined as the smallest ordinal \rho where the \rho-th iterated derived set is empty, measures the "scattered height" tied to the order type itself. For example, the space [0, \omega) has rank 1, as it is discrete with all points isolated and the first derived set empty. Regarding metrizability, the order topology on a well-ordered set is metrizable if and only if the order type is countable. Countable ordinals, such as \omega, admit a countable basis consisting of open intervals, and being regular Hausdorff spaces, they satisfy the conditions of the Urysohn metrization theorem, yielding a metric equivalent to the discrete topology on the naturals. In contrast, uncountable ordinals like \omega_1, the first uncountable ordinal, possess an uncountable basis and fail to be second countable, rendering them non-metrizable. A prominent example illustrating non-metrizability is the construction involving \omega_1, which underpins the long line—a space formed by \omega_1 \times [0,1) equipped with the lexicographic order topology. This yields a connected, locally Euclidean, non-metrizable manifold-like space that is sequentially compact but not compact, highlighting how well-orderings extend beyond familiar metric structures. Sequential properties vary by point type in ordinal spaces. Successor ordinals are first-countable, with a countable local basis of neighborhoods shrinking to the point, such as sequences of initial segments. However, limit ordinals of uncountable cofinality, like points in \omega_1, are not first-countable; any local basis requires cardinality equal to the cofinality, exceeding \aleph_0, which prevents sequential determination of limits at such points.

References

  1. [1]
    Well Ordered Set -- from Wolfram MathWorld
    A totally ordered set (A,<=) is said to be well ordered (or have a well-founded order) iff every nonempty subset of A has a least element.
  2. [2]
    [PDF] The Well-Ordering Theorem
    A set is well-ordered if any nonempty subset has a least element. The natural numbers are well-ordered, but integers are not with the usual order.
  3. [3]
  4. [4]
    Beweis, daß jede Menge wohlgeordnet werden kann
    Beweis, daß jede Menge wohlgeordnet werden kann. Aus einem an Herrn Hilbert gerichteten Briefe. Download PDF.
  5. [5]
    Beiträge zur Begründung der transfiniten Mengenlehre
    Cantor, G. Beiträge zur Begründung der transfiniten Mengenlehre. Math. Ann. 46, 481–512 (1895). https://doi.org/10.1007/BF02124929
  6. [6]
    well-order in nLab
    Dec 26, 2023 · A well-order is a relation that allows a set to be interpreted as an ordinal number, and is transitive, extensional, and well-founded.Idea · Definition · Examples · Interpretation as an ordinal...
  7. [7]
    3.7: The Well-Ordering Principle
    ### Properties of Well-Ordered Sets
  8. [8]
    Basic set theory - Stanford Encyclopedia of Philosophy
    The set of all finite ordinals is denoted by the Greek letter omega (\(\omega\)). Thus, \(\omega\) is just the set \(\mathbb{N}\) of natural numbers. \(\omega\ ...
  9. [9]
    Natural Numbers:Well-Ordering - Department of Mathematics at UTSA
    Nov 7, 2021 · First, well-order the real numbers (this is where the axiom of choice enters via the well-ordering theorem), giving a sequence ⟨ r α | α < β ⟩ ...
  10. [10]
    [PDF] The Well-Ordering Principle
    the validity of mathematical induction directly from the axioms of set theory. ... Every non-empty subset of the natural numbers has a least element. Proof ...
  11. [11]
    [PDF] 1. Peano's Axioms and Natural Numbers
    Lemma 1.9 (Well ordering of N). If n, m ∈ N, then exactly one of the following is true. Either n<m or n = m or m<n ...
  12. [12]
    [PDF] CARDINAL AND ORDINAL NUMBERS Contents 1. The Natural ...
    This construction of ω makes rigorous the intuitive description of the natural numbers as {0, 1, 2, 3, ..}, where the ellipsis represent the so on ad infinitum ...
  13. [13]
    [PDF] Math 310 Class Notes 4: The Well-ordering Principle
    Theorem 1. The mathematical induction is logically equivalent to the well-ordering principle. Proof. Part I. We show the well-ordering principle implies ...
  14. [14]
  15. [15]
    [PDF] An Introduction to Real Analysis John K. Hunter - UC Davis Math
    These are some notes on introductory real analysis. They cover the properties of the real numbers, sequences and series of real numbers, limits of functions ...
  16. [16]
    Order. Well-ordered sets
    The set of rational numbers is naturally totally ordered but not well ordered; the ordering is dense. The set of real numbers is naturally totally ordered ...
  17. [17]
    [PDF] Contributions to the Founding of the Theory of Transfinite Numbers
    and with the work of Cantor from 1870 to 1895. Some notes at the end contain a short account of the developments of the theory of transfinite numbers since ...
  18. [18]
    [PDF] On Carruth's Axioms for Natural Sums and Products1
    The first natural sum and product of ordinals, the so-called Hessenberg sum and Hessen- berg product, were introduced by Hessenberg in [5]. The novelty of these ...
  19. [19]
    [PDF] On Cantor's normal form theorem and algebraic number theory
    Jun 2, 2018 · Cantor introduced his normal form theorem as an ordinal polyno- mial for the countable ordinals of the second class up to the first epsilon.
  20. [20]
    Zermelo's Axiomatization of Set Theory
    Jul 2, 2013 · Indeed, Zermelo assumes that 'every set can be well-ordered' is equivalent to the Cantorian 'every cardinality is an aleph' (Zermelo 1904: 141).
  21. [21]
    well-ordering theorem in nLab
    ### Summary of Well-Ordering Theorem Content
  22. [22]
  23. [23]
    [PDF] CS 6110 S17 Lecture 6 Well-Founded Induction
    A binary relation ≺ is said to be well-founded if it has no infinite descending chains. An infinite descending chain is an infinite sequence of elements a0,a1, ...
  24. [24]
    [PDF] Day 12
    Jun 7, 2012 · Suppose that A with is well ordered. Then it is a linear order and there are no infinite descending chains. We want to show that it has the well ...
  25. [25]
    [PDF] Chapter 8 Ordered Sets
    The statement that any set of cardinals is well-ordered is clearly equivalent to the following theorem. Theorem 11.6 If is a nonempty collection of sets, then.<|control11|><|separator|>
  26. [26]
    245B, Notes 7: Well-ordered sets, ordinals, and Zorn's lemma ...
    Jan 28, 2009 · \rho(X) is isomorphic to X for every well-ordered set X. (In particular, if \rho(X) and \rho(Y) are equal, then X and Y are isomorphic.) · If ...
  27. [27]
    Well orders and the ordinal numbers - by Joel David Hamkins
    Dec 15, 2023 · A well order is a linear order with the further property that every nonempty subset of the domain has a least element.
  28. [28]
    [PDF] 6.2. The ordinals - People
    A set α is an ordinal iff it is transitive and well-ordered by ... An ordinal α is said to be a limit ordinal iff it is not 0 and is not a successor ordinal.
  29. [29]
    [PDF] 1 Initial Notation and Definitions
    Jan 21, 2016 · Strict total orders are defined in an analogous way. • A well order is a strict total order that is well-founded. • If B is a well order, hS, Bi ...
  30. [30]
    [PDF] Well-founded orderings - Stony Brook Computer Science
    Examples of strict orderings are the less-than and the greater-than relation on the integers. (The less-than-or-equal-to relation is a non-strict order,.
  31. [31]
    The Axiom of Choice - Stanford Encyclopedia of Philosophy
    Jan 8, 2008 · The Axiom of Choice ... Zermelo introduces axioms of set theory, explicitly formulates AC and uses it to prove the well-ordering theorem, thereby ...Origins and Chronology of the... · Mathematical Applications of... · Bibliography
  32. [32]
  33. [33]
    [PDF] Ordered Topological Spaces. Definition. Let S be a set. The relation ...
    If S is an ordered set with a first and last element and it has the least upper bound property, then S with the order topology is compact. Theorem 8.9. Let S be ...<|control11|><|separator|>
  34. [34]
    [PDF] 10. Orders and ω1
    The primary reason we introduce this topic is to define order topologies, and in particular the order topology on some new objects: ω1 and ω1 + 1.
  35. [35]
    The First Uncountable Ordinal | Dan Ma's Topology Blog
    Oct 11, 2009 · This is a basic discussion on the first uncountable ordinal \omega_1 and its immediate successor \omega_1+1.
  36. [36]
    [PDF] arXiv:1606.04025v2 [math.GN] 17 Nov 2016
    Nov 17, 2016 · It is well-known that ordinals are locally compact and scattered (for the last property we note that the smallest element of a subset A of X is ...Missing: compactness | Show results with:compactness
  37. [37]
    Countable Closed Ordinal Space is Metrizable - ProofWiki
    Theorem. Let Ω denote the first uncountable ordinal. Let Γ be a limit ordinal which strictly precedes Ω. Let [0..Γ] denote the closed ordinal space on Γ.
  38. [38]
    [PDF] Counterexamples in Topology - rexresearch1
    A counterexample, in its most restricted sense, is an example which dis- proves a famous conjecture. We choose to interpret the word more broadly, particularly ...