Fact-checked by Grok 2 weeks ago

Heine–Borel theorem

The Heine–Borel theorem is a cornerstone of and , asserting that a subset S of the \mathbb{R}^n is compact it is both closed and bounded. This equivalence provides a precise characterization of in finite-dimensional Euclidean spaces, where compactness means that every open cover of S admits a finite subcover. Named after the German mathematician Eduard Heine (1821–1881) and the French mathematician (1871–1956), the theorem emerged from late 19th-century efforts to rigorize . Heine contributed foundational ideas in his paper on of functions on closed bounded intervals, employing covering arguments that prefigured . Borel provided the first explicit statement and proof in his 1895 doctoral dissertation Sur quelques points de la théorie des fonctions, showing that a closed bounded interval in \mathbb{R} has the finite subcover property for countable open covers. Subsequent generalizations by Pierre Cousin (1895), (1904), and others extended the result to arbitrary open covers and higher dimensions, solidifying its modern form. The theorem's significance lies in its role as a bridge between metric properties (closedness and boundedness) and topological ones (), enabling key results such as the and the of continuous functions on compact sets. It holds specifically for \mathbb{R}^n with the standard but fails in infinite-dimensional spaces like function spaces, highlighting the peculiarities of finite-dimensional geometry.

Statement and Fundamentals

Formal Statement

In the context of Euclidean spaces, the Heine–Borel theorem provides a characterization of compact subsets. Consider \mathbb{R}^n, the n-dimensional space equipped with the standard induced by the Euclidean metric d(x, y) = \|x - y\|, where \|\cdot\| denotes the Euclidean . A K \subseteq \mathbb{R}^n is defined to be compact if, for every collection of open sets \{U_\alpha\}_{\alpha \in A} such that K \subseteq \bigcup_{\alpha \in A} U_\alpha, there exists a finite subcollection \{U_{\alpha_1}, \dots, U_{\alpha_m}\} with K \subseteq \bigcup_{i=1}^m U_{\alpha_i}. This property, known as having the finite subcover property for open covers, is a fundamental notion in . A subset K \subseteq \mathbb{R}^n is closed if its complement \mathbb{R}^n \setminus K is open, meaning that for every point x \notin K, there exists \varepsilon > 0 such that the open ball B(x, \varepsilon) = \{y \in \mathbb{R}^n : \|y - x\| < \varepsilon\} is contained in \mathbb{R}^n \setminus K. Equivalently, K contains all its limit points, though the topological definition via complements emphasizes the structure of the ambient space. A subset K \subseteq \mathbb{R}^n is bounded if the diameter \operatorname{diam}(K) = \sup\{\|x - y\| : x, y \in K\} < \infty, or equivalently, if there exists a point q \in \mathbb{R}^n and M > 0 such that K \subseteq B(q, M). These notions of closedness and boundedness capture sequential and metric constraints essential to the theorem. The Heine–Borel theorem asserts that a K \subseteq \mathbb{R}^n is compact it is both closed and bounded. This biconditional holds for any n \geq 1, with the case n=1 on the real line \mathbb{R} serving as the foundational instance, where compact subsets are precisely the closed and bounded intervals (including singletons and the ). In the proof of the direction from closed and bounded to compact, an \varepsilon/3 argument is often employed to refine open covers into finite subcovers by partitioning the space into smaller balls of radius \varepsilon/3, ensuring uniform control over the covering without delving into the full derivation here.

Key Implications

The Heine–Borel theorem plays a pivotal role in by providing a precise characterization of sets in spaces, establishing that is equivalent to being closed and bounded. This equivalence simplifies the study of topological properties in \mathbb{R}^n, where guarantees desirable analytical behaviors that fail in non-compact settings. Specifically, in spaces, a set ensures sequential , meaning every sequence in the set has a convergent with in the set. Furthermore, implies total boundedness, where for every \epsilon > 0, the set can be covered by finitely many balls of radius \epsilon, facilitating approximations and coverings essential for analysis. One of the theorem's key implications is its foundational support for the , which states that a continuous function on a compact set in \mathbb{R}^n attains its maximum and minimum values. The proof of this theorem relies directly on the Heine–Borel characterization, using compactness to extract finite subcovers and ensure the function's is compact and thus closed and bounded. Similarly, the theorem underpins the , asserting that every bounded sequence in \mathbb{R}^n has a convergent subsequence; in compact sets, this subsequence converges within the set, linking boundedness to sequential limits. In practical terms, these implications extend to and optimization problems in \mathbb{R}^n. For , compactness ensures that continuous functions on compact domains are Riemann integrable, as the derived from allows for bounded variation and finite partitions. In optimization, the guarantees the existence of global maxima and minima for continuous objective functions over closed and bounded feasible regions, a for methods in and . To illustrate the theorem's necessity, consider non-compact sets like the open (0,1) in \mathbb{R}, which is bounded but not closed and thus not compact. Continuous functions such as f(x) = x on (0,1) fail to attain a maximum, and sequences like \{1/n\} converge to a point outside the set, highlighting how the absence of leads to the breakdown of sequential and extremal properties.

Historical Context

Origins and Motivation

In the 1870s, mathematicians sought to establish rigorous foundations for calculus amid emerging paradoxes that undermined intuitive geometric arguments, such as Karl Weierstrass's 1872 construction of a continuous function nowhere differentiable, which highlighted the need for precise definitions of continuity and limits. This era of analysis was driven by efforts to resolve inconsistencies in classical results, including gaps in Peter Gustav Lejeune Dirichlet's 1829 principle for the pointwise convergence of Fourier series, which required functions to be piecewise continuous with bounded variation but left open questions about more general continuous functions. Heine's investigations into these convergence issues, particularly in his 1870 paper demonstrating that Fourier series of continuous functions of bounded variation converge to the function, underscored the importance of uniform convergence and bounded domains for analytical stability. A precursor to these developments appeared in Bernard Bolzano's 1817 purely analytic proof of the , where he introduced a novel theorem on the boundedness of s: every bounded of real numbers contains an , providing early insights into compactness-like properties essential for bounding continuous functions on intervals. Building directly on such foundations and the Fourier series motivations, Eduard Heine addressed the limitations of pointwise continuity by examining conditions under which sequences of continuous functions converge uniformly to a continuous . In his seminal 1872 paper "Die Elemente der Functionenlehre," published in Crelle's Journal für die reine und angewandte Mathematik, Heine proved that every on a closed bounded is and attains its values, thereby clarifying when bounded sets permit uniform limits without pathological behavior. This work, influenced by Weierstrass's epsilon-delta approach to limits, filled critical gaps in understanding bounded functions on closed , ensuring that holds precisely under these conditions and advancing the rigorization of .

Key Developments

Following Heine's foundational contributions in the mid-19th century, Émile Borel provided the first explicit statement and proof of the theorem in 1895, establishing that a closed bounded interval on the real line is compact with respect to countable open covers, as detailed in his paper "Sur quelques points de la théorie des fonctions" published in the Annales de l'École Normale Supérieure. That same year, Pierre Cousin extended Borel's result to arbitrary open covers in his work "Sur les fonctions de n variables complexes," providing a key generalization for uncountable collections. Borel extended this result to higher dimensions within his development of measure theory, applying it to multidimensional intervals to characterize measurable sets and facilitate integration over bounded regions, as elaborated in his 1898 monograph Leçons sur la théorie des fonctions. Henri Lebesgue further generalized the theorem around 1898, solidifying its application to closed bounded sets in \mathbb{R}^n with the full modern statement for arbitrary open covers. In the 1910s and early 1920s, the theorem's implications for compactness gained wider recognition among mathematicians. Sierpiński's 1927 paper "La notion de dérivée comme base d'une théorie des ensembles abstraits" in Mathematische Annalen further integrated derived sets into early general topology, contributing to the axiomatization of abstract spaces. The 1920s marked the theorem's pivotal transition from real analysis to general topology, driven by Felix Hausdorff's formalization of topological structures in his 1914 book Grundzüge der Mengenlehre (revised in 1927), where he defined compactness via finite subcovers of open sets and proved an abstract version of the Heine-Borel result for closed subsets of Euclidean spaces. This abstraction influenced subsequent advancements, notably Andrey Tychonoff's 1930 theorem, which generalized compactness to arbitrary products of compact spaces, building directly on Heine-Borel principles to extend the property beyond finite dimensions.

Proof in Euclidean Spaces

Proof of Compactness Implying Closed and Bounded

In \mathbb{R}^n equipped with the standard , is equivalent to sequential , meaning that a K \subseteq \mathbb{R}^n is compact every in K has a converging to a point in K. This equivalence holds more generally in any , where sequential implies via the construction of Lebesgue numbers for open covers and the extraction of convergent subsequences to contradict infinite covers without finite subcovers, while the converse follows from the fact that limit points of sequences in compact sets must lie within the set due to the of closed balls. In \mathbb{R}^n, this lemma facilitates proofs using sequences but can be avoided in favor of direct open cover arguments for the Heine–Borel theorem's one direction. To prove that every compact subset K \subseteq \mathbb{R}^n is bounded, consider the open cover of \mathbb{R}^n given by the open balls B(0, m) = \{ x \in \mathbb{R}^n : \|x\| < m \} for m = 1, 2, 3, \dots , where \|\cdot\| denotes the Euclidean norm and $0 is the origin. The restrictions of these balls to K form an open cover of K. By compactness of K, there exists a finite subcover B(0, m_1), \dots, B(0, m_k) such that K \subseteq \bigcup_{i=1}^k B(0, m_i). Let M = \max\{m_1, \dots, m_k\}; then K \subseteq B(0, M), so K is bounded. Equivalently, by contrapositive, if K is unbounded, then for every m \in \mathbb{N}, there exists x_m \in K with \|x_m\| > m, and the cover \{B(0, m) \cap K : m \in \mathbb{N}\} admits no finite subcover, contradicting compactness. To prove that every compact subset K \subseteq \mathbb{R}^n is closed, it suffices to show that its complement K^c = \mathbb{R}^n \setminus K is open. Fix any x \in K^c. Consider the collection of open sets U_m = \mathbb{R}^n \setminus \overline{B}(x, 1/m) for m = 1, 2, 3, \dots , where \overline{B}(x, r) is the closed ball of radius r centered at x. These sets form an open cover of K, since if some point y \in K were in every \overline{B}(x, 1/m), then \mathrm{dist}(x, y) \leq 1/m for all m, implying \mathrm{dist}(x, y) = 0 and y = x \in K, a contradiction. By compactness, there is a finite subcover U_{m_1}, \dots, U_{m_k} such that K \subseteq \bigcup_{i=1}^k U_{m_i}. Let N = \max\{m_1, \dots, m_k\}; then K \subseteq \mathbb{R}^n \setminus \overline{B}(x, 1/N), so B(x, 1/N) \subseteq K^c. Thus, B(x, 1/N) is an open neighborhood of x contained in K^c, proving K^c open and K closed. Equivalently, by contrapositive, if K is not closed, then there exists x \in \overline{K} \setminus K, and the cover \{U_m : m \in \mathbb{N}\} has no finite subcover, as points in K arbitrarily close to x require increasingly small $1/m.

Proof of Closed and Bounded Implying Compactness

The Heine–Borel theorem states that in Euclidean space \mathbb{R}^n, a subset is compact if and only if it is closed and bounded. The direction that compactness implies closed and bounded follows readily from properties of compact sets, but the converse—closed and bounded implies compact—requires more involved arguments, typically via sequential compactness or open cover exhaustion. These proofs rely on the completeness and finite-dimensional structure of \mathbb{R}^n, ensuring total boundedness for bounded sets.

Sequential Compactness Approach

To establish compactness, it suffices to show sequential compactness, as every sequentially compact of a is compact, and conversely in \mathbb{R}^n (equipped with the ), compactness, sequential compactness, and total boundedness plus are equivalent characterizations. A set K \subseteq \mathbb{R}^n is sequentially compact if every in K has a converging to a point in K. Assume K is closed and bounded. Boundedness ensures K lies within some large of finite radius, so by the , every sequence \{x_k\} \subseteq K has a convergent subsequence \{x_{k_j}\} with limit L \in \mathbb{R}^n. Since K is closed, L \in K. Thus, K is sequentially compact. The holds in \mathbb{R}^n by on dimension, reducing to the one-dimensional case where bounded sequences in closed intervals have convergent subsequences via the nested interval theorem.

Open Cover Approach via Total Boundedness

Alternatively, since a closed bounded K \subseteq \mathbb{R}^n is complete and totally bounded (as bounded sets in \mathbb{R}^n admit finite \varepsilon-nets for any \varepsilon > 0), it is compact. To see this directly via open covers, suppose \{U_\alpha\}_{\alpha \in A} is an open cover of K with no finite subcover. Let \alpha = \mathrm{diam}(K). Construct a sequence of nested nonempty closed subsets K_n \subseteq K inductively such that \mathrm{diam}(K_n) \leq \alpha / 2^n and no finite subcollection of \{U_\alpha\} covers K_n. For n=1, since K is totally bounded, cover K by finitely many closed balls of radius \alpha/2 with centers in K. At least one such ball B_1, say K_1 = B_1 \cap K, cannot be covered by finitely many U_\alpha. Proceed inductively: given K_n, cover it by finitely many closed balls of radius \alpha / 2^{n+1} with centers in K_n, and select K_{n+1} as one such ball intersected with K_n that admits no finite subcover. Pick x_n \in K_n; then \{x_n\} is Cauchy (since \mathrm{diam}(K_n) \to 0) and converges to some x \in K by completeness. There exists U \in \{U_\alpha\} with x \in U, so some r > 0 has B(x, r) \subseteq U. For large N, K_N \subseteq B(x, r) \subseteq U, so K_N is covered by the single set U, contradicting the construction. Thus, every open cover has a finite subcover.

Generalizations

To Metric Spaces

In metric spaces, the Heine–Borel theorem generalizes such that a A of a (X, d) is compact if and only if A is complete and totally bounded. Total boundedness means that for every \epsilon > 0, there exists a of points in X (an \epsilon-net) such that every point in A is within \epsilon of one of these points: \forall \epsilon > 0, \ \exists \{x_1, \dots, x_n\} \subset X \ \text{such that} \ A \subset \bigcup_{i=1}^n B(x_i, \epsilon), where B(x_i, \epsilon) denotes the open ball of radius \epsilon centered at x_i. This equivalence holds because compactness in metric spaces implies sequential compactness, which in turn ensures completeness (every Cauchy sequence converges in A) and total boundedness (no infinite discrete subsets). In complete metric spaces, compactness is thus equivalent to total boundedness plus completeness, but the original closed-and-bounded condition does not always suffice. For instance, the closed unit ball in the infinite-dimensional Hilbert space \ell^2 (sequences of square-summable real numbers with the standard norm) is closed and bounded but not compact, as it fails to be totally bounded—the orthonormal basis \{e_n\} (where e_n has 1 in the nth position and 0 elsewhere) has no finite $1/2-net, since \|e_m - e_n\| = \sqrt{2} for m \neq n. This illustrates how infinite dimensionality prevents closed bounded sets from being totally bounded. The Heine–Borel property also fails in incomplete metric spaces. Consider the space \mathbb{Q} of rational numbers with the usual metric d(p, q) = |p - q|; the subset [0, 1] \cap \mathbb{Q} is closed and bounded in \mathbb{Q} but not compact, as it is incomplete (Cauchy sequences like decimal approximations to \sqrt{2}/2 do not converge in \mathbb{Q}). In the specific case of \mathbb{R}^n with the Euclidean metric, closed and bounded subsets are totally bounded (due to finite dimensionality allowing finite covers by balls) and complete (as closed subsets of the complete space \mathbb{R}^n), recovering the classical Heine–Borel theorem.

To Topological Spaces

In general topological spaces, the Heine–Borel theorem lacks a direct analogue because the concept of boundedness is not intrinsically defined without additional structure, such as a or uniform structure; compactness is instead characterized topologically as every open cover admitting a finite subcover, often termed quasi-compactness to emphasize the absence of separation axioms. Without such structure, compactness does not correspond to being both closed and bounded, as "bounded" requires notions like in uniform spaces to quantify uniformity of closeness. For instance, in uniform spaces, a is bounded if for every , the can be covered by finitely many sets of less than that , but this fails to align with in purely topological settings lacking uniformity. A key partial generalization holds in Hausdorff spaces, where every compact subset is closed, reflecting the separation properties that prevent accumulation points outside the subset. However, even here, boundedness does not follow from compactness without further assumptions, as illustrated by the fact that the torus \mathbb{T}^2 = \mathbb{R}^2 / \mathbb{Z}^2 is compact but its universal cover \mathbb{R}^2 is unbounded, showing how compactness in quotient topologies can arise without bounded lifts in the covering space. In non-Hausdorff spaces, compactness need not imply closedness at all; a standard counterexample is the space X = \{a, b, c\} with open sets \emptyset, \{a\}, \{a,b\}, X, where \{b\} is compact—any open cover must include an open containing b, such as \{a,b\} or X, yielding a finite subcover—but \{b\} is not closed, as its closure is \{b,c\}. Tychonoff's theorem provides a significant extension, stating that the product of any collection of topological spaces, equipped with the , is ; this generalizes the finite-product case used in proofs for spaces and applies broadly without requiring boundedness or even Hausdorff assumptions on factors. in also connects to other covering properties: every is locally compact (each point has a compact neighborhood basis) and paracompact (every open cover admits a locally finite open refinement), but these implications reverse only under additional conditions like second-countability or Hausdorff separation, highlighting limitations in broader generalizations of the Heine–Borel . For contrast, while spaces may retain aspects of the theorem via total boundedness, general topological spaces prioritize open-cover definitions over distance-based criteria.

Heine–Borel Property

Definition and Characterization

The Heine–Borel property of a X is the condition that every closed and of X is compact. This property generalizes the classical Heine–Borel theorem, which characterizes compactness in spaces as precisely the closed and bounded subsets. In metric spaces, the Heine–Borel property holds every closed and bounded subset admits a finite subcover from any open cover. Equivalently, closed bounded subsets are totally bounded and complete. The \mathbb{R}^n possesses this property for all finite n, ensuring that aligns with closure and boundedness in these dimensions. In contrast, no infinite-dimensional has the Heine–Borel property, as closed balls in such spaces fail to be compact. To define boundedness in uniform spaces, where the property extends naturally, consider a uniform space (X, \mathcal{U}) with base of entourages \mathcal{U}. A subset A \subseteq X is bounded if, for every entourage U \in \mathcal{U}, there exists a positive integer n such that A \times A \subseteq U^n, where U^n denotes the n-fold composition of U with itself. This entourage-based notion captures the intuitive idea of finite "extent" without relying on a metric.

Examples in Specific Spaces

In Euclidean spaces \mathbb{R}^n for finite n, every closed and bounded subset is compact, and conversely, every compact subset is closed and bounded, fully embodying the Heine–Borel property. This characterization relies on the standard topology induced by the Euclidean metric, where boundedness ensures total boundedness and closedness preserves sequential compactness. In contrast, the infinite-dimensional Hilbert space \ell^2 of square-summable sequences lacks the Heine–Borel property. Its closed unit ball, defined as \{ x \in \ell^2 : \|x\|_2 \leq 1 \}, is closed and bounded but not compact, as it fails to be totally bounded; for instance, the orthonormal basis vectors form an infinite sequence with no convergent subsequence within the ball. This illustrates how infinite dimensionality prevents the equivalence between closed boundedness and compactness in normed spaces. Among topological vector spaces over \mathbb{R} or \mathbb{C}, those that are finite-dimensional possess the Heine–Borel property, as they are topologically isomorphic to \mathbb{R}^n or \mathbb{C}^n, inheriting compactness criteria from spaces. However, infinite-dimensional examples, such as Banach spaces beyond finite dimensions, do not; their closed unit balls are typically non-compact due to the absence of local . A locally bounded topological vector space with the Heine–Borel must thus be finite-dimensional, highlighting dimensionality as a key . Consider an uncountable set equipped with the discrete metric, where d(x,y) = 1 if x \neq y and d(x,x) = 0. The entire space is closed and bounded (with diameter 1), yet not compact, as the open cover consisting of singleton sets \{ \{x\} : x \in X \} admits no finite subcover. This metric space thus fails the Heine–Borel property, demonstrating that even bounded closed sets can lack compactness in non-complete or non-locally compact settings. The field of p-adic numbers \mathbb{Q}_p, for a prime p, equipped with its non-Archimedean , possesses the Heine–Borel property as a locally compact . Here, closed and bounded subsets are compact; for example, the p-adic integers \mathbb{Z}_p = \{ x \in \mathbb{Q}_p : |x|_p \leq 1 \} form a compact open ball, and the theorem extends to all such subsets via a modified covering argument analogous to the real case. This holds more broadly for locally compact non-Archimedean fields.

References

  1. [1]
    Heine-Borel Theorem - Department of Mathematics at UTSA
    Oct 27, 2021 · The Heine–Borel theorem, named after Eduard Heine and Émile Borel, states: For a subset S of Euclidean space R n , the following two statements are equivalent.
  2. [2]
    [PDF] Early Work Uniform Continuity to the Heine-Borel Theorem
    Sep 6, 2022 · simply means “every open cover has a finite subcover” and Borel was the first to publish a statement and proof of the key ideas asserted in ...
  3. [3]
  4. [4]
    [PDF] Section 26. Compact Sets
    Jul 27, 2016 · The (hopefully) familiar Heine-Borel Theorem states that a set of real numbers is compact if and only if it is closed and bounded (which we ...
  5. [5]
    [PDF] 15 | Heine-Borel Theorem
    The Heine-Borel Theorem states that a set A in R^n is compact if and only if A is closed and bounded.Missing: formal | Show results with:formal
  6. [6]
    [PDF] Rudin (1976) Principles of Mathematical Analysis.djvu
    This book is intended to serve as a text for the course in analysis that is usually taken by advanced undergraduates or by first-year students who study ...
  7. [7]
    Heine-Borel theorem in nLab
    ### Formal Statement of the Heine-Borel Theorem in \(\mathbb{R}^n\)
  8. [8]
    [PDF] An Introduction to Real Analysis John K. Hunter - UC Davis Math
    A subset of R is compact if and only if it is closed and bounded. This result follows from the Heine-Borel theorem, that every open cover of a closed, bounded ...Missing: implications | Show results with:implications
  9. [9]
    [PDF] Metric Topology
    (c) In a metric space: sequential compactness ⇔ compactness ⇒ total boundedness. (d) Heine-Borel Theorem. Let X be a subset of Rn equipped with the metric ...
  10. [10]
    [PDF] 2.4 The Extreme Value Theorem and Some of its Consequences
    The Extreme Value Theorem deals with the question of when we can be sure that for a given function f , (1) the values f (x) don't get too big or too small, (2) ...
  11. [11]
    The Jagged, Monstrous Function That Broke Calculus
    Jan 23, 2025 · Weierstrass' function, continuous but not differentiable, challenged calculus's intuition, forcing a shift to analysis and new definitions of ...
  12. [12]
    [PDF] Connections in Mathematical Analysis: The Case of Fourier Series
    May 15, 2017 · This is the question that. Heinrich Eduard Heine (1821-1881), of the University of Halle, posed himself, and in 1870 he showed that if a ...
  13. [13]
    [PDF] Bolzano, Cauchy and the intermediate value theorem - HAL
    Bolzano (1817) was the first to explicitly question the validity of proofs of the Intermediate Value. Theorem based on geometrical arguments, thus paving the ...
  14. [14]
    Die Elemente der Functionenlehre. - EuDML
    Heine, E.. "Die Elemente der Functionenlehre.." Journal für die reine und angewandte Mathematik 74 (1872): 172-188. <http://eudml.org/doc/148175>.
  15. [15]
    [PDF] Sur quelques points de la théorie des fonctions - Numdam
    Je donne pour ces fonctions une expression analytique (la somme d'une série de puissances et d'une série de Fourier) telle que les expres- sions analytiques des ...
  16. [16]
    Leçons sur la théorie des fonctions : Borel, Emile, 1871-1956
    Apr 11, 2006 · Leçons sur la théorie des fonctions. viii p., 1 l., 136 p. 26 cm. À quelques modifications près, ces leçons sont la reproduction de conférences faites à l'É ...Missing: Heine- higher dimensions
  17. [17]
    La notion de dérivée comme base d'une théorie des ensembles ...
    La notion de dérivée comme base d'une théorie des ensembles abstraits ... Article PDF. Download to read the full article text. Use our pre ...
  18. [18]
    La notion de dérivée comme base d'une théorie des ensembles ...
    Sierpinski, W.. "La notion de dérivée comme base d'une théorie des ensembles abstraits." Mathematische Annalen 97 (1927): 321-337. <http://eudml.org/doc ...
  19. [19]
    [PDF] Grundzüge der Mengenlehre
    Das vorliegende Werk will ein Lehrbuch und kein Bericht sein: es versucht die Hauptsachen der Mengenlehre ohne Voraussetzung.
  20. [20]
    [PDF] Über die topologische Erweiterung von Räumen - Digizeitschriften
    Titel: Über die topologische Erweiterung von Räumen. Autor: Tychonoff, A. Jahr: 1930. PURL: https://resolver.sub.uni-goettingen.de/purl?235181684_0102|log33.
  21. [21]
    [PDF] Compactness - Penn Math
    (a) implies (b): Since E is bounded it is contained in some closed interval I. This interval is compact (Theorem 2.40). But then E is a closed subset of a ...
  22. [22]
    [PDF] Notes on Compactness - Northwestern Math Department
    Proposition. A compact subspace of a metric space is closed and bounded. Proof. Let K be a compact subspace of a metric space M. The “ ...
  23. [23]
    [PDF] Topology of the Real Numbers - UC Davis Math
    Theorem 5.56 (Heine-Borel). A subset of R is compact if and only if it is closed and bounded. Page 13. 5.3. Compact sets. 101. Proof. The most important ...
  24. [24]
    [PDF] Section 27. Compact Subspaces of the Real Line
    Aug 1, 2016 · The Heine-Borel Theorem. A subspace A of Rn is compact if and only if it is closed and is bounded in the. Euclidean metric d or the ...Missing: implies | Show results with:implies
  25. [25]
    [PDF] compact sets in metric spaces notes for math 703
    This theorem is called the Heine-Borel theorem and is usually derived from the theorem that a closed bounded interval is compact. This latter theorem has at ...Missing: total | Show results with:total
  26. [26]
    Compactness and Heine-Borel - Advanced Analysis
    Jan 20, 2024 · A subset of a Hilbert space is compact if and only if it is closed, bounded, and close to finite-dimensional.Missing: total | Show results with:total
  27. [27]
    [PDF] 01. Metrics and topologies on vector spaces 1. Euclidean spaces
    Dec 30, 2019 · [12.1] Theorem: A set E in a metric space X has compact closure if and only if it is totally bounded. [12.2] Remark: Sometimes a set with ...<|control11|><|separator|>
  28. [28]
    [PDF] Analysis 1 Colloquium of Week 10 Compactness - Math (Princeton)
    Nov 19, 2014 · A metric space is said to have the Heine–Borel property if every closed and bounded subset is compact. Many metric spaces fail to have the Heine ...Missing: total | Show results with:total
  29. [29]
    Boundedness in uniform spaces and topological groups - EuDML
    Hejcman, Jan. "Boundedness in uniform spaces and topological groups." Czechoslovak Mathematical Journal 09.4 (1959): 544-563. <http://eudml.org/doc/ ...
  30. [30]
    Are all compact sets closed and bounded? – Math 320: Real Analysis
    Sep 29, 2016 · So this set and topology provide a counterexample that not all compact sets need to be closed. , such as being compact and not closed but also ...
  31. [31]
    [PDF] 16. Compactness
    We will specifically prove an important result from analysis called the Heine-Borel theorem that characterizes the compact subsets of Rn. This result is so ...Missing: history | Show results with:history
  32. [32]
    [PDF] Compactness in metric spaces
    The metric space X is said to be compact if every open covering has a finite subcovering.1. This abstracts the Heine–Borel property; indeed, the Heine–Borel.
  33. [33]
    [PDF] CONSTRUCTING METRICS WITH THE HEINE-BOREL PROPERTY
    By the Heine-Borel (HB) property of a metric space (X, d) we mean here that every closed bounded set is compact, i.e. bounded sets are totally bounded, and we ...
  34. [34]
    [PDF] arXiv:2101.06140v1 [math-ph] 15 Jan 2021
    Jan 15, 2021 · To this end, we recall that every second-countable, locally compact Hausdorff space can be endowed with a Heine-Borel metric (for details we ...
  35. [35]
    [PDF] boundedness on uniform spaces and it's mappings
    Feb 17, 2018 · Abstract: In this note we introduce a boundedness on uniform spaces. We study some properties of corresponding bounded sets.
  36. [36]
    Heine-Borel Theorem/Euclidean Space - ProofWiki
    Nov 4, 2023 · Theorem. Let n∈N>0. Let C be a subspace of the Euclidean space Rn. Then C is closed and bounded if and only if it is compact.
  37. [37]
    compactness of closed unit ball in normed spaces - PlanetMath.org
    Mar 22, 2013 · It follows that infinite dimensional (http://planetmath.org/Dimension2) normed spaces are not locally compact. Proof: •. (⟸) ( ⟸ ) This is ...<|separator|>
  38. [38]
    [PDF] Hilbert spaces
    A general result in a metric space is that any compact set is both closed and bounded, so this must be true in a Hilbert space. The Heine-Borel theorem gives a.
  39. [39]
    [PDF] TOPOLOGICAL VECTOR SPACES1 1. Definitions and basic facts.
    A locally bounded TVS with the Heine-Borel property is finite dimensional. (Follows since V is bounded and closed, hence compact.) 3. Seminorms, local convexity ...
  40. [40]
    [PDF] QMMM Warm-up 2: Real analysis
    Problem 8. Prove, using the Heine–Borel Theorem, that a finite dimensional normed space is locally compact.
  41. [41]
    [PDF] Examples for the failure of a Heine-Borel type theorem in metric ...
    Theorem: C0[a, b], together with the max distance is a complete metric space. This is easy to prove (and for that matter generalizes to C0(K) where K is any ...
  42. [42]
    [PDF] P-ADIC NUMBERS
    The ball B(a, p−m) is open since for every b ∈ B(a, p−m) we have. B(b, p−m) = B(a, p−m). To prove the compactness we modify the proof of the Heine-Borel theorem.
  43. [43]
    [PDF] Interpolation series for continuous functions on Π-adic ... - UTK Math
    (In fact, the Heine-Borel Theorem holds in all locally compact non-archimedean fields, a result due to. Schöbe [9].) In the special case, the complete field ...