fbpx
Wikipedia

Compact space

In mathematics, specifically general topology, compactness is a property that seeks to generalize the notion of a closed and bounded subset of Euclidean space[1] by making precise the idea of a space having no "punctures" or "missing endpoints", i.e. that the space not exclude any limiting values of points. For example, the open interval (0,1) would not be compact because it excludes the limiting values of 0 and 1, whereas the closed interval [0,1] would be compact. Similarly, the space of rational numbers is not compact, because it has infinitely many "punctures" corresponding to the irrational numbers, and the space of real numbers is not compact either, because it excludes the two limiting values and . However, the extended real number line would be compact, since it contains both infinities. There are many ways to make this heuristic notion precise. These ways usually agree in a metric space, but may not be equivalent in other topological spaces.

Per the compactness criteria for Euclidean space as stated in the Heine–Borel theorem, the interval A = (−∞, −2] is not compact because it is not bounded. The interval C = (2, 4) is not compact because it is not closed. The interval B = [0, 1] is compact because it is both closed and bounded.

One such generalization is that a topological space is sequentially compact if every infinite sequence of points sampled from the space has an infinite subsequence that converges to some point of the space.[2]

The Bolzano–Weierstrass theorem states that a subset of Euclidean space is compact in this sequential sense if and only if it is closed and bounded.

Thus, if one chooses an infinite number of points in the closed unit interval [0, 1], some of those points will get arbitrarily close to some real number in that space. For instance, some of the numbers in the sequence 1/2, 4/5, 1/3, 5/6, 1/4, 6/7, ... accumulate to 0 (while others accumulate to 1). The same set of points would not accumulate to any point of the open unit interval (0, 1), so the open unit interval is not compact. Although subsets (subspaces) of Euclidean space can be compact, the entire space itself is not compact, since it is not bounded. For example, considering (the real number line), the sequence of points 0,  1,  2,  3, ... has no subsequence that converges to any real number.

Compactness was formally introduced by Maurice Fréchet in 1906 to generalize the Bolzano–Weierstrass theorem from spaces of geometrical points to spaces of functions. The Arzelà–Ascoli theorem and the Peano existence theorem exemplify applications of this notion of compactness to classical analysis. Following its initial introduction, various equivalent notions of compactness, including sequential compactness and limit point compactness, were developed in general metric spaces.[3] In general topological spaces, however, these notions of compactness are not necessarily equivalent. The most useful notion — and the standard definition of the unqualified term compactness — is phrased in terms of the existence of finite families of open sets that "cover" the space in the sense that each point of the space lies in some set contained in the family. This more subtle notion, introduced by Pavel Alexandrov and Pavel Urysohn in 1929, exhibits compact spaces as generalizations of finite sets. In spaces that are compact in this sense, it is often possible to patch together information that holds locally — that is, in a neighborhood of each point — into corresponding statements that hold throughout the space, and many theorems are of this character.

The term compact set is sometimes used as a synonym for compact space, but also often refers to a compact subspace of a topological space.

Historical development

In the 19th century, several disparate mathematical properties were understood that would later be seen as consequences of compactness. On the one hand, Bernard Bolzano (1817) had been aware that any bounded sequence of points (in the line or plane, for instance) has a subsequence that must eventually get arbitrarily close to some other point, called a limit point. Bolzano's proof relied on the method of bisection: the sequence was placed into an interval that was then divided into two equal parts, and a part containing infinitely many terms of the sequence was selected. The process could then be repeated by dividing the resulting smaller interval into smaller and smaller parts — until it closes down on the desired limit point. The full significance of Bolzano's theorem, and its method of proof, would not emerge until almost 50 years later when it was rediscovered by Karl Weierstrass.[4]

In the 1880s, it became clear that results similar to the Bolzano–Weierstrass theorem could be formulated for spaces of functions rather than just numbers or geometrical points. The idea of regarding functions as themselves points of a generalized space dates back to the investigations of Giulio Ascoli and Cesare Arzelà.[5] The culmination of their investigations, the Arzelà–Ascoli theorem, was a generalization of the Bolzano–Weierstrass theorem to families of continuous functions, the precise conclusion of which was that it was possible to extract a uniformly convergent sequence of functions from a suitable family of functions. The uniform limit of this sequence then played precisely the same role as Bolzano's "limit point". Towards the beginning of the twentieth century, results similar to that of Arzelà and Ascoli began to accumulate in the area of integral equations, as investigated by David Hilbert and Erhard Schmidt. For a certain class of Green's functions coming from solutions of integral equations, Schmidt had shown that a property analogous to the Arzelà–Ascoli theorem held in the sense of mean convergence — or convergence in what would later be dubbed a Hilbert space. This ultimately led to the notion of a compact operator as an offshoot of the general notion of a compact space. It was Maurice Fréchet who, in 1906, had distilled the essence of the Bolzano–Weierstrass property and coined the term compactness to refer to this general phenomenon (he used the term already in his 1904 paper[6] which led to the famous 1906 thesis).

However, a different notion of compactness altogether had also slowly emerged at the end of the 19th century from the study of the continuum, which was seen as fundamental for the rigorous formulation of analysis. In 1870, Eduard Heine showed that a continuous function defined on a closed and bounded interval was in fact uniformly continuous. In the course of the proof, he made use of a lemma that from any countable cover of the interval by smaller open intervals, it was possible to select a finite number of these that also covered it. The significance of this lemma was recognized by Émile Borel (1895), and it was generalized to arbitrary collections of intervals by Pierre Cousin (1895) and Henri Lebesgue (1904). The Heine–Borel theorem, as the result is now known, is another special property possessed by closed and bounded sets of real numbers.

This property was significant because it allowed for the passage from local information about a set (such as the continuity of a function) to global information about the set (such as the uniform continuity of a function). This sentiment was expressed by Lebesgue (1904), who also exploited it in the development of the integral now bearing his name. Ultimately, the Russian school of point-set topology, under the direction of Pavel Alexandrov and Pavel Urysohn, formulated Heine–Borel compactness in a way that could be applied to the modern notion of a topological space. Alexandrov & Urysohn (1929) showed that the earlier version of compactness due to Fréchet, now called (relative) sequential compactness, under appropriate conditions followed from the version of compactness that was formulated in terms of the existence of finite subcovers. It was this notion of compactness that became the dominant one, because it was not only a stronger property, but it could be formulated in a more general setting with a minimum of additional technical machinery, as it relied only on the structure of the open sets in a space.

Basic examples

Any finite space is compact; a finite subcover can be obtained by selecting, for each point, an open set containing it. A nontrivial example of a compact space is the (closed) unit interval [0,1] of real numbers. If one chooses an infinite number of distinct points in the unit interval, then there must be some accumulation point in that interval. For instance, the odd-numbered terms of the sequence 1, 1/2, 1/3, 3/4, 1/5, 5/6, 1/7, 7/8, ... get arbitrarily close to 0, while the even-numbered ones get arbitrarily close to 1. The given example sequence shows the importance of including the boundary points of the interval, since the limit points must be in the space itself — an open (or half-open) interval of the real numbers is not compact. It is also crucial that the interval be bounded, since in the interval [0,∞), one could choose the sequence of points 0, 1, 2, 3, ..., of which no sub-sequence ultimately gets arbitrarily close to any given real number.

In two dimensions, closed disks are compact since for any infinite number of points sampled from a disk, some subset of those points must get arbitrarily close either to a point within the disc, or to a point on the boundary. However, an open disk is not compact, because a sequence of points can tend to the boundary — without getting arbitrarily close to any point in the interior. Likewise, spheres are compact, but a sphere missing a point is not since a sequence of points can still tend to the missing point, thereby not getting arbitrarily close to any point within the space. Lines and planes are not compact, since one can take a set of equally-spaced points in any given direction without approaching any point.

Definitions

Various definitions of compactness may apply, depending on the level of generality. A subset of Euclidean space in particular is called compact if it is closed and bounded. This implies, by the Bolzano–Weierstrass theorem, that any infinite sequence from the set has a subsequence that converges to a point in the set. Various equivalent notions of compactness, such as sequential compactness and limit point compactness, can be developed in general metric spaces.[3]

In contrast, the different notions of compactness are not equivalent in general topological spaces, and the most useful notion of compactness — originally called bicompactness — is defined using covers consisting of open sets (see Open cover definition below). That this form of compactness holds for closed and bounded subsets of Euclidean space is known as the Heine–Borel theorem. Compactness, when defined in this manner, often allows one to take information that is known locally — in a neighbourhood of each point of the space — and to extend it to information that holds globally throughout the space. An example of this phenomenon is Dirichlet's theorem, to which it was originally applied by Heine, that a continuous function on a compact interval is uniformly continuous; here, continuity is a local property of the function, and uniform continuity the corresponding global property.

Open cover definition

Formally, a topological space X is called compact if each of its open covers has a finite subcover.[7] That is, X is compact if for every collection C of open subsets of X such that

 ,

there is a finite subcollection FC such that

 

Some branches of mathematics such as algebraic geometry, typically influenced by the French school of Bourbaki, use the term quasi-compact for the general notion, and reserve the term compact for topological spaces that are both Hausdorff and quasi-compact. A compact set is sometimes referred to as a compactum, plural compacta.

Compactness of subsets

A subset K of a topological space X is said to be compact if it is compact as a subspace (in the subspace topology). That is, K is compact if for every arbitrary collection C of open subsets of X such that

 

there is a finite subcollection FC such that

 

Compactness is a "topological" property. That is, if  , with subset Z equipped with the subspace topology, then K is compact in Z if and only if K is compact in Y.

Characterization

If X is a topological space then the following are equivalent:

  1. X is compact; i.e., every open cover of X has a finite subcover.
  2. X has a sub-base such that every cover of the space, by members of the sub-base, has a finite subcover (Alexander's sub-base theorem).
  3. X is Lindelöf and countably compact.[8]
  4. Any collection of closed subsets of X with the finite intersection property has nonempty intersection.
  5. Every net on X has a convergent subnet (see the article on nets for a proof).
  6. Every filter on X has a convergent refinement.
  7. Every net on X has a cluster point.
  8. Every filter on X has a cluster point.
  9. Every ultrafilter on X converges to at least one point.
  10. Every infinite subset of X has a complete accumulation point.[9]
  11. For every topological space Y, the projection   is a closed mapping[10] (see proper map).

Bourbaki defines a compact space (quasi-compact space) as a topological space where each filter has a cluster point (i.e., 8. in the above).[11]

Euclidean space

For any subset A of Euclidean space, A is compact if and only if it is closed and bounded; this is the Heine–Borel theorem.

As a Euclidean space is a metric space, the conditions in the next subsection also apply to all of its subsets. Of all of the equivalent conditions, it is in practice easiest to verify that a subset is closed and bounded, for example, for a closed interval or closed n-ball.

Metric spaces

For any metric space (X, d), the following are equivalent (assuming countable choice):

  1. (X, d) is compact.
  2. (X, d) is complete and totally bounded (this is also equivalent to compactness for uniform spaces).[12]
  3. (X, d) is sequentially compact; that is, every sequence in X has a convergent subsequence whose limit is in X (this is also equivalent to compactness for first-countable uniform spaces).
  4. (X, d) is limit point compact (also called weakly countably compact); that is, every infinite subset of X has at least one limit point in X.
  5. (X, d) is countably compact; that is, every countable open cover of X has a finite subcover.
  6. (X, d) is an image of a continuous function from the Cantor set.[13]
  7. Every decreasing nested sequence of nonempty closed subsets S1S2 ⊇ ... in (X, d) has a nonempty intersection.
  8. Every increasing nested sequence of proper open subsets S1S2 ⊆ ... in (X, d) fails to cover X.

A compact metric space (X, d) also satisfies the following properties:

  1. Lebesgue's number lemma: For every open cover of X, there exists a number δ > 0 such that every subset of X of diameter < δ is contained in some member of the cover.
  2. (X, d) is second-countable, separable and Lindelöf – these three conditions are equivalent for metric spaces. The converse is not true; e.g., a countable discrete space satisfies these three conditions, but is not compact.
  3. X is closed and bounded (as a subset of any metric space whose restricted metric is d). The converse may fail for a non-Euclidean space; e.g. the real line equipped with the discrete metric is closed and bounded but not compact, as the collection of all singletons of the space is an open cover which admits no finite subcover. It is complete but not totally bounded.

Ordered Spaces

For an ordered space (X, <) (i.e. a totally ordered set equipped with the order topology), the following are equivalent:

  1. (X, <) is compact.
  2. Every subset of X has a supremum (i.e. a least upper bound) in X.
  3. Every subset of X has an infimum (i.e. a greatest lower bound) in X.
  4. Every nonempty closed subset of X has a maximum and a minimum element.

An ordered space satisfying (any one of) these conditions is called a complete lattice.

In addition, the following are equivalent for all ordered spaces (X, <), and (assuming countable choice) are true whenever (X, <) is compact. (The converse in general fails if (X, <) is not also metrizable.):

  1. Every sequence in (X, <) has a subsequence that converges in (X, <).
  2. Every monotone increasing sequence in X converges to a unique limit in X.
  3. Every monotone decreasing sequence in X converges to a unique limit in X.
  4. Every decreasing nested sequence of nonempty closed subsets S1S2 ⊇ ... in (X, <) has a nonempty intersection.
  5. Every increasing nested sequence of proper open subsets S1S2 ⊆ ... in (X, <) fails to cover X.

Characterization by continuous functions

Let X be a topological space and C(X) the ring of real continuous functions on X. For each pX, the evaluation map   given by evp(f) = f(p) is a ring homomorphism. The kernel of evp is a maximal ideal, since the residue field C(X)/ker evp is the field of real numbers, by the first isomorphism theorem. A topological space X is pseudocompact if and only if every maximal ideal in C(X) has residue field the real numbers. For completely regular spaces, this is equivalent to every maximal ideal being the kernel of an evaluation homomorphism.[14] There are pseudocompact spaces that are not compact, though.

In general, for non-pseudocompact spaces there are always maximal ideals m in C(X) such that the residue field C(X)/m is a (non-Archimedean) hyperreal field. The framework of non-standard analysis allows for the following alternative characterization of compactness:[15] a topological space X is compact if and only if every point x of the natural extension *X is infinitely close to a point x0 of X (more precisely, x is contained in the monad of x0).

Hyperreal definition

A space X is compact if its hyperreal extension *X (constructed, for example, by the ultrapower construction) has the property that every point of *X is infinitely close to some point of X*X. For example, an open real interval X = (0, 1) is not compact because its hyperreal extension *(0,1) contains infinitesimals, which are infinitely close to 0, which is not a point of X.

Sufficient conditions

  • A closed subset of a compact space is compact.[16]
  • A finite union of compact sets is compact.
  • A continuous image of a compact space is compact.[17]
  • The intersection of any non-empty collection of compact subsets of a Hausdorff space is compact (and closed);
    • If X is not Hausdorff then the intersection of two compact subsets may fail to be compact (see footnote for example).[a]
  • The product of any collection of compact spaces is compact. (This is Tychonoff's theorem, which is equivalent to the axiom of choice.)
  • In a metrizable space, a subset is compact if and only if it is sequentially compact (assuming countable choice)
  • A finite set endowed with any topology is compact.

Properties of compact spaces

  • A compact subset of a Hausdorff space X is closed.
    • If X is not Hausdorff then a compact subset of X may fail to be a closed subset of X (see footnote for example).[b]
    • If X is not Hausdorff then the closure of a compact set may fail to be compact (see footnote for example).[c]
  • In any topological vector space (TVS), a compact subset is complete. However, every non-Hausdorff TVS contains compact (and thus complete) subsets that are not closed.
  • If A and B are disjoint compact subsets of a Hausdorff space X, then there exist disjoint open set U and V in X such that AU and BV.
  • A continuous bijection from a compact space into a Hausdorff space is a homeomorphism.
  • A compact Hausdorff space is normal and regular.
  • If a space X is compact and Hausdorff, then no finer topology on X is compact and no coarser topology on X is Hausdorff.
  • If a subset of a metric space (X, d) is compact then it is d-bounded.

Functions and compact spaces

Since a continuous image of a compact space is compact, the extreme value theorem holds for such spaces: a continuous real-valued function on a nonempty compact space is bounded above and attains its supremum.[18] (Slightly more generally, this is true for an upper semicontinuous function.) As a sort of converse to the above statements, the pre-image of a compact space under a proper map is compact.

Compactifications

Every topological space X is an open dense subspace of a compact space having at most one point more than X, by the Alexandroff one-point compactification. By the same construction, every locally compact Hausdorff space X is an open dense subspace of a compact Hausdorff space having at most one point more than X.

Ordered compact spaces

A nonempty compact subset of the real numbers has a greatest element and a least element.

Let X be a simply ordered set endowed with the order topology. Then X is compact if and only if X is a complete lattice (i.e. all subsets have suprema and infima).[19]

Examples

  • Any finite topological space, including the empty set, is compact. More generally, any space with a finite topology (only finitely many open sets) is compact; this includes in particular the trivial topology.
  • Any space carrying the cofinite topology is compact.
  • Any locally compact Hausdorff space can be turned into a compact space by adding a single point to it, by means of Alexandroff one-point compactification. The one-point compactification of   is homeomorphic to the circle S1; the one-point compactification of   is homeomorphic to the sphere S2. Using the one-point compactification, one can also easily construct compact spaces which are not Hausdorff, by starting with a non-Hausdorff space.
  • The right order topology or left order topology on any bounded totally ordered set is compact. In particular, Sierpiński space is compact.
  • No discrete space with an infinite number of points is compact. The collection of all singletons of the space is an open cover which admits no finite subcover. Finite discrete spaces are compact.
  • In   carrying the lower limit topology, no uncountable set is compact.
  • In the cocountable topology on an uncountable set, no infinite set is compact. Like the previous example, the space as a whole is not locally compact but is still Lindelöf.
  • The closed unit interval [0, 1] is compact. This follows from the Heine–Borel theorem. The open interval (0, 1) is not compact: the open cover   for n = 3, 4, ...  does not have a finite subcover. Similarly, the set of rational numbers in the closed interval [0,1] is not compact: the sets of rational numbers in the intervals   cover all the rationals in [0, 1] for n = 4, 5, ...  but this cover does not have a finite subcover. Here, the sets are open in the subspace topology even though they are not open as subsets of  .
  • The set   of all real numbers is not compact as there is a cover of open intervals that does not have a finite subcover. For example, intervals (n − 1, n + 1), where n takes all integer values in Z, cover   but there is no finite subcover.
  • On the other hand, the extended real number line carrying the analogous topology is compact; note that the cover described above would never reach the points at infinity and thus would not cover the extended real line. In fact, the set has the homeomorphism to [−1, 1] of mapping each infinity to its corresponding unit and every real number to its sign multiplied by the unique number in the positive part of interval that results in its absolute value when divided by one minus itself, and since homeomorphisms preserve covers, the Heine-Borel property can be inferred.
  • For every natural number n, the n-sphere is compact. Again from the Heine–Borel theorem, the closed unit ball of any finite-dimensional normed vector space is compact. This is not true for infinite dimensions; in fact, a normed vector space is finite-dimensional if and only if its closed unit ball is compact.
  • On the other hand, the closed unit ball of the dual of a normed space is compact for the weak-* topology. (Alaoglu's theorem)
  • The Cantor set is compact. In fact, every compact metric space is a continuous image of the Cantor set.
  • Consider the set K of all functions f :   → [0, 1] from the real number line to the closed unit interval, and define a topology on K so that a sequence   in K converges towards fK if and only if   converges towards f(x) for all real numbers x. There is only one such topology; it is called the topology of pointwise convergence or the product topology. Then K is a compact topological space; this follows from the Tychonoff theorem.
  • Consider the set K of all functions f : [0, 1] → [0, 1] satisfying the Lipschitz condition |f(x) − f(y)| ≤ |x − y| for all xy ∈ [0,1]. Consider on K the metric induced by the uniform distance   Then by Arzelà–Ascoli theorem the space K is compact.
  • The spectrum of any bounded linear operator on a Banach space is a nonempty compact subset of the complex numbers  . Conversely, any compact subset of   arises in this manner, as the spectrum of some bounded linear operator. For instance, a diagonal operator on the Hilbert space   may have any compact nonempty subset of   as spectrum.

Algebraic examples

See also

Notes

  1. ^ Let X = {a, b} ∪  , U = {a} ∪  , and V = {b} ∪  . Endow X with the topology generated by the following basic open sets: every subset of   is open; the only open sets containing a are X and U; and the only open sets containing b are X and V. Then U and V are both compact subsets but their intersection, which is  , is not compact. Note that both U and V are compact open subsets, neither one of which is closed.
  2. ^ Let X = {a, b} and endow X with the topology {X, ∅, {a}}. Then {a} is a compact set but it is not closed.
  3. ^ Let X be the set of non-negative integers. We endow X with the particular point topology by defining a subset UX to be open if and only if 0 ∈ U. Then S := {0} is compact, the closure of S is all of X, but X is not compact since the collection of open subsets {{0, x} : xX} does not have a finite subcover.

References

  1. ^ "Compactness". Encyclopaedia Britannica. mathematics. Retrieved 2019-11-25 – via britannica.com.
  2. ^ Engelking, Ryszard (1977). General Topology. Warsaw, PL: PWN. p. 266.
  3. ^ a b "Sequential compactness". www-groups.mcs.st-andrews.ac.uk. MT 4522 course lectures. Retrieved 2019-11-25.
  4. ^ Kline 1990, pp. 952–953; Boyer & Merzbach 1991, p. 561
  5. ^ Kline 1990, Chapter 46, §2
  6. ^ Frechet, M. 1904. Generalisation d'un theorem de Weierstrass. Analyse Mathematique.
  7. ^ Weisstein, Eric W. "Compact Space". mathworld.wolfram.com. Retrieved 2019-11-25.
  8. ^ Howes 1995, pp. xxvi–xxviii.
  9. ^ Kelley 1955, p. 163
  10. ^ Bourbaki 2007, § 10.2. Theorem 1, Corollary 1.
  11. ^ Bourbaki 2007, § 9.1. Definition 1.
  12. ^ Arkhangel'skii & Fedorchuk 1990, Theorem 5.3.7
  13. ^ Willard 1970 Theorem 30.7.
  14. ^ Gillman & Jerison 1976, §5.6
  15. ^ Robinson 1996, Theorem 4.1.13
  16. ^ Arkhangel'skii & Fedorchuk 1990, Theorem 5.2.3; Closed set in a compact space is compact at PlanetMath.; Closed subsets of a compact set are compact at PlanetMath.
  17. ^ Arkhangel'skii & Fedorchuk 1990, Theorem 5.2.2; See also Compactness is preserved under a continuous map at PlanetMath.
  18. ^ Arkhangel'skii & Fedorchuk 1990, Corollary 5.2.1
  19. ^ Steen & Seebach 1995, p. 67

Bibliography

  • Alexandrov, Pavel; Urysohn, Pavel (1929). "Mémoire sur les espaces topologiques compacts". Koninklijke Nederlandse Akademie van Wetenschappen te Amsterdam, Proceedings of the Section of Mathematical Sciences. 14.
  • Arkhangel'skii, A.V.; Fedorchuk, V.V. (1990). "The basic concepts and constructions of general topology". In Arkhangel'skii, A.V.; Pontrjagin, L.S. (eds.). General Topology I. Encyclopedia of the Mathematical Sciences. Vol. 17. Springer. ISBN 978-0-387-18178-3..
  • Arkhangel'skii, A.V. (2001) [1994], "Compact space", Encyclopedia of Mathematics, EMS Press.
  • Bolzano, Bernard (1817). Rein analytischer Beweis des Lehrsatzes, dass zwischen je zwey Werthen, die ein entgegengesetzes Resultat gewähren, wenigstens eine reele Wurzel der Gleichung liege. Wilhelm Engelmann. (Purely analytic proof of the theorem that between any two values which give results of opposite sign, there lies at least one real root of the equation).
  • Borel, Émile (1895). "Sur quelques points de la théorie des fonctions". Annales Scientifiques de l'École Normale Supérieure. 3. 12: 9–55. doi:10.24033/asens.406. JFM 26.0429.03.
  • Bourbaki, Nicolas (2007). Topologie générale. Chapitres 1 à 4. Berlin: Springer. doi:10.1007/978-3-540-33982-3. ISBN 978-3-540-33982-3.
  • Boyer, Carl B. (1959). The history of the calculus and its conceptual development. New York: Dover Publications. MR 0124178.
  • Boyer, Carl Benjamin; Merzbach, Uta C (1991). A History of Mathematics (2nd ed.). John Wiley & Sons. ISBN 978-0-471-54397-8.
  • Arzelà, Cesare (1895). "Sulle funzioni di linee". Mem. Accad. Sci. Ist. Bologna Cl. Sci. Fis. Mat. 5 (5): 55–74.
  • Arzelà, Cesare (1882–1883). "Un'osservazione intorno alle serie di funzioni". Rend. Dell' Accad. R. Delle Sci. dell'Istituto di Bologna: 142–159.
  • Ascoli, G. (1883–1884). "Le curve limiti di una varietà data di curve". Atti della R. Accad. Dei Lincei Memorie della Cl. Sci. Fis. Mat. Nat. 18 (3): 521–586.
  • Fréchet, Maurice (1906). "Sur quelques points du calcul fonctionnel". Rendiconti del Circolo Matematico di Palermo. 22 (1): 1–72. doi:10.1007/BF03018603. hdl:10338.dmlcz/100655. S2CID 123251660.
  • Gillman, Leonard; Jerison, Meyer (1976). Rings of continuous functions. Springer-Verlag.
  • Howes, Norman R. (23 June 1995). Modern Analysis and Topology. Graduate Texts in Mathematics. New York: Springer-Verlag Science & Business Media. ISBN 978-0-387-97986-1. OCLC 31969970. OL 1272666M.
  • Kelley, John (1955). General topology. Graduate Texts in Mathematics. Vol. 27. Springer-Verlag.
  • Kline, Morris (1990) [1972]. Mathematical thought from ancient to modern times (3rd ed.). Oxford University Press. ISBN 978-0-19-506136-9.
  • Lebesgue, Henri (1904). Leçons sur l'intégration et la recherche des fonctions primitives. Gauthier-Villars.
  • Robinson, Abraham (1996). Non-standard analysis. Princeton University Press. ISBN 978-0-691-04490-3. MR 0205854.
  • Scarborough, C.T.; Stone, A.H. (1966). "Products of nearly compact spaces" (PDF). Transactions of the American Mathematical Society. 124 (1): 131–147. doi:10.2307/1994440. JSTOR 1994440. (PDF) from the original on 2017-08-16..
  • Steen, Lynn Arthur; Seebach, J. Arthur Jr. (1995) [1978]. Counterexamples in Topology (Dover Publications reprint of 1978 ed.). Berlin, New York: Springer-Verlag. ISBN 978-0-486-68735-3. MR 0507446.
  • Willard, Stephen (1970). General Topology. Dover publications. ISBN 0-486-43479-6.

External links

  • Countably compact at PlanetMath.
  • Sundström, Manya Raman (2010). "A pedagogical history of compactness". arXiv:1006.4131v1 [math.HO].

This article incorporates material from Examples of compact spaces on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.

compact, space, compactness, redirects, here, other, uses, compactness, disambiguation, mathematics, specifically, general, topology, compactness, property, that, seeks, generalize, notion, closed, bounded, subset, euclidean, space, making, precise, idea, spac. Compactness redirects here For other uses see Compactness disambiguation In mathematics specifically general topology compactness is a property that seeks to generalize the notion of a closed and bounded subset of Euclidean space 1 by making precise the idea of a space having no punctures or missing endpoints i e that the space not exclude any limiting values of points For example the open interval 0 1 would not be compact because it excludes the limiting values of 0 and 1 whereas the closed interval 0 1 would be compact Similarly the space of rational numbers Q displaystyle mathbb Q is not compact because it has infinitely many punctures corresponding to the irrational numbers and the space of real numbers R displaystyle mathbb R is not compact either because it excludes the two limiting values displaystyle infty and displaystyle infty However the extended real number line would be compact since it contains both infinities There are many ways to make this heuristic notion precise These ways usually agree in a metric space but may not be equivalent in other topological spaces Per the compactness criteria for Euclidean space as stated in the Heine Borel theorem the interval A 2 is not compact because it is not bounded The interval C 2 4 is not compact because it is not closed The interval B 0 1 is compact because it is both closed and bounded One such generalization is that a topological space is sequentially compact if every infinite sequence of points sampled from the space has an infinite subsequence that converges to some point of the space 2 The Bolzano Weierstrass theorem states that a subset of Euclidean space is compact in this sequential sense if and only if it is closed and bounded Thus if one chooses an infinite number of points in the closed unit interval 0 1 some of those points will get arbitrarily close to some real number in that space For instance some of the numbers in the sequence 1 2 4 5 1 3 5 6 1 4 6 7 accumulate to 0 while others accumulate to 1 The same set of points would not accumulate to any point of the open unit interval 0 1 so the open unit interval is not compact Although subsets subspaces of Euclidean space can be compact the entire space itself is not compact since it is not bounded For example considering R 1 displaystyle mathbb R 1 the real number line the sequence of points 0 1 2 3 has no subsequence that converges to any real number Compactness was formally introduced by Maurice Frechet in 1906 to generalize the Bolzano Weierstrass theorem from spaces of geometrical points to spaces of functions The Arzela Ascoli theorem and the Peano existence theorem exemplify applications of this notion of compactness to classical analysis Following its initial introduction various equivalent notions of compactness including sequential compactness and limit point compactness were developed in general metric spaces 3 In general topological spaces however these notions of compactness are not necessarily equivalent The most useful notion and the standard definition of the unqualified term compactness is phrased in terms of the existence of finite families of open sets that cover the space in the sense that each point of the space lies in some set contained in the family This more subtle notion introduced by Pavel Alexandrov and Pavel Urysohn in 1929 exhibits compact spaces as generalizations of finite sets In spaces that are compact in this sense it is often possible to patch together information that holds locally that is in a neighborhood of each point into corresponding statements that hold throughout the space and many theorems are of this character The term compact set is sometimes used as a synonym for compact space but also often refers to a compact subspace of a topological space Contents 1 Historical development 2 Basic examples 3 Definitions 3 1 Open cover definition 3 2 Compactness of subsets 3 3 Characterization 3 3 1 Euclidean space 3 3 2 Metric spaces 3 3 3 Ordered Spaces 3 3 4 Characterization by continuous functions 3 3 5 Hyperreal definition 4 Sufficient conditions 5 Properties of compact spaces 5 1 Functions and compact spaces 5 2 Compactifications 5 3 Ordered compact spaces 6 Examples 6 1 Algebraic examples 7 See also 8 Notes 9 References 10 Bibliography 11 External linksHistorical development EditIn the 19th century several disparate mathematical properties were understood that would later be seen as consequences of compactness On the one hand Bernard Bolzano 1817 had been aware that any bounded sequence of points in the line or plane for instance has a subsequence that must eventually get arbitrarily close to some other point called a limit point Bolzano s proof relied on the method of bisection the sequence was placed into an interval that was then divided into two equal parts and a part containing infinitely many terms of the sequence was selected The process could then be repeated by dividing the resulting smaller interval into smaller and smaller parts until it closes down on the desired limit point The full significance of Bolzano s theorem and its method of proof would not emerge until almost 50 years later when it was rediscovered by Karl Weierstrass 4 In the 1880s it became clear that results similar to the Bolzano Weierstrass theorem could be formulated for spaces of functions rather than just numbers or geometrical points The idea of regarding functions as themselves points of a generalized space dates back to the investigations of Giulio Ascoli and Cesare Arzela 5 The culmination of their investigations the Arzela Ascoli theorem was a generalization of the Bolzano Weierstrass theorem to families of continuous functions the precise conclusion of which was that it was possible to extract a uniformly convergent sequence of functions from a suitable family of functions The uniform limit of this sequence then played precisely the same role as Bolzano s limit point Towards the beginning of the twentieth century results similar to that of Arzela and Ascoli began to accumulate in the area of integral equations as investigated by David Hilbert and Erhard Schmidt For a certain class of Green s functions coming from solutions of integral equations Schmidt had shown that a property analogous to the Arzela Ascoli theorem held in the sense of mean convergence or convergence in what would later be dubbed a Hilbert space This ultimately led to the notion of a compact operator as an offshoot of the general notion of a compact space It was Maurice Frechet who in 1906 had distilled the essence of the Bolzano Weierstrass property and coined the term compactness to refer to this general phenomenon he used the term already in his 1904 paper 6 which led to the famous 1906 thesis However a different notion of compactness altogether had also slowly emerged at the end of the 19th century from the study of the continuum which was seen as fundamental for the rigorous formulation of analysis In 1870 Eduard Heine showed that a continuous function defined on a closed and bounded interval was in fact uniformly continuous In the course of the proof he made use of a lemma that from any countable cover of the interval by smaller open intervals it was possible to select a finite number of these that also covered it The significance of this lemma was recognized by Emile Borel 1895 and it was generalized to arbitrary collections of intervals by Pierre Cousin 1895 and Henri Lebesgue 1904 The Heine Borel theorem as the result is now known is another special property possessed by closed and bounded sets of real numbers This property was significant because it allowed for the passage from local information about a set such as the continuity of a function to global information about the set such as the uniform continuity of a function This sentiment was expressed by Lebesgue 1904 who also exploited it in the development of the integral now bearing his name Ultimately the Russian school of point set topology under the direction of Pavel Alexandrov and Pavel Urysohn formulated Heine Borel compactness in a way that could be applied to the modern notion of a topological space Alexandrov amp Urysohn 1929 showed that the earlier version of compactness due to Frechet now called relative sequential compactness under appropriate conditions followed from the version of compactness that was formulated in terms of the existence of finite subcovers It was this notion of compactness that became the dominant one because it was not only a stronger property but it could be formulated in a more general setting with a minimum of additional technical machinery as it relied only on the structure of the open sets in a space Basic examples EditAny finite space is compact a finite subcover can be obtained by selecting for each point an open set containing it A nontrivial example of a compact space is the closed unit interval 0 1 of real numbers If one chooses an infinite number of distinct points in the unit interval then there must be some accumulation point in that interval For instance the odd numbered terms of the sequence 1 1 2 1 3 3 4 1 5 5 6 1 7 7 8 get arbitrarily close to 0 while the even numbered ones get arbitrarily close to 1 The given example sequence shows the importance of including the boundary points of the interval since the limit points must be in the space itself an open or half open interval of the real numbers is not compact It is also crucial that the interval be bounded since in the interval 0 one could choose the sequence of points 0 1 2 3 of which no sub sequence ultimately gets arbitrarily close to any given real number In two dimensions closed disks are compact since for any infinite number of points sampled from a disk some subset of those points must get arbitrarily close either to a point within the disc or to a point on the boundary However an open disk is not compact because a sequence of points can tend to the boundary without getting arbitrarily close to any point in the interior Likewise spheres are compact but a sphere missing a point is not since a sequence of points can still tend to the missing point thereby not getting arbitrarily close to any point within the space Lines and planes are not compact since one can take a set of equally spaced points in any given direction without approaching any point Definitions EditVarious definitions of compactness may apply depending on the level of generality A subset of Euclidean space in particular is called compact if it is closed and bounded This implies by the Bolzano Weierstrass theorem that any infinite sequence from the set has a subsequence that converges to a point in the set Various equivalent notions of compactness such as sequential compactness and limit point compactness can be developed in general metric spaces 3 In contrast the different notions of compactness are not equivalent in general topological spaces and the most useful notion of compactness originally called bicompactness is defined using covers consisting of open sets see Open cover definition below That this form of compactness holds for closed and bounded subsets of Euclidean space is known as the Heine Borel theorem Compactness when defined in this manner often allows one to take information that is known locally in a neighbourhood of each point of the space and to extend it to information that holds globally throughout the space An example of this phenomenon is Dirichlet s theorem to which it was originally applied by Heine that a continuous function on a compact interval is uniformly continuous here continuity is a local property of the function and uniform continuity the corresponding global property Open cover definition Edit Formally a topological space X is called compact if each of its open covers has a finite subcover 7 That is X is compact if for every collection C of open subsets of X such that X x C x displaystyle X bigcup x in C x there is a finite subcollection F C such that X x F x displaystyle X bigcup x in F x Some branches of mathematics such as algebraic geometry typically influenced by the French school of Bourbaki use the term quasi compact for the general notion and reserve the term compact for topological spaces that are both Hausdorff and quasi compact A compact set is sometimes referred to as a compactum plural compacta Compactness of subsets Edit A subset K of a topological space X is said to be compact if it is compact as a subspace in the subspace topology That is K is compact if for every arbitrary collection C of open subsets of X such that K c C c displaystyle K subseteq bigcup c in C c there is a finite subcollection F C such that K c F c displaystyle K subseteq bigcup c in F c Compactness is a topological property That is if K Z Y displaystyle K subset Z subset Y with subset Z equipped with the subspace topology then K is compact in Z if and only if K is compact in Y Characterization Edit If X is a topological space then the following are equivalent X is compact i e every open cover of X has a finite subcover X has a sub base such that every cover of the space by members of the sub base has a finite subcover Alexander s sub base theorem X is Lindelof and countably compact 8 Any collection of closed subsets of X with the finite intersection property has nonempty intersection Every net on X has a convergent subnet see the article on nets for a proof Every filter on X has a convergent refinement Every net on X has a cluster point Every filter on X has a cluster point Every ultrafilter on X converges to at least one point Every infinite subset of X has a complete accumulation point 9 For every topological space Y the projection X Y Y displaystyle X times Y to Y is a closed mapping 10 see proper map Bourbaki defines a compact space quasi compact space as a topological space where each filter has a cluster point i e 8 in the above 11 Euclidean space Edit For any subset A of Euclidean space A is compact if and only if it is closed and bounded this is the Heine Borel theorem As a Euclidean space is a metric space the conditions in the next subsection also apply to all of its subsets Of all of the equivalent conditions it is in practice easiest to verify that a subset is closed and bounded for example for a closed interval or closed n ball Metric spaces Edit For any metric space X d the following are equivalent assuming countable choice X d is compact X d is complete and totally bounded this is also equivalent to compactness for uniform spaces 12 X d is sequentially compact that is every sequence in X has a convergent subsequence whose limit is in X this is also equivalent to compactness for first countable uniform spaces X d is limit point compact also called weakly countably compact that is every infinite subset of X has at least one limit point in X X d is countably compact that is every countable open cover of X has a finite subcover X d is an image of a continuous function from the Cantor set 13 Every decreasing nested sequence of nonempty closed subsets S1 S2 in X d has a nonempty intersection Every increasing nested sequence of proper open subsets S1 S2 in X d fails to cover X A compact metric space X d also satisfies the following properties Lebesgue s number lemma For every open cover of X there exists a number d gt 0 such that every subset of X of diameter lt d is contained in some member of the cover X d is second countable separable and Lindelof these three conditions are equivalent for metric spaces The converse is not true e g a countable discrete space satisfies these three conditions but is not compact X is closed and bounded as a subset of any metric space whose restricted metric is d The converse may fail for a non Euclidean space e g the real line equipped with the discrete metric is closed and bounded but not compact as the collection of all singletons of the space is an open cover which admits no finite subcover It is complete but not totally bounded Ordered Spaces Edit For an ordered space X lt i e a totally ordered set equipped with the order topology the following are equivalent X lt is compact Every subset of X has a supremum i e a least upper bound in X Every subset of X has an infimum i e a greatest lower bound in X Every nonempty closed subset of X has a maximum and a minimum element An ordered space satisfying any one of these conditions is called a complete lattice In addition the following are equivalent for all ordered spaces X lt and assuming countable choice are true whenever X lt is compact The converse in general fails if X lt is not also metrizable Every sequence in X lt has a subsequence that converges in X lt Every monotone increasing sequence in X converges to a unique limit in X Every monotone decreasing sequence in X converges to a unique limit in X Every decreasing nested sequence of nonempty closed subsets S1 S2 in X lt has a nonempty intersection Every increasing nested sequence of proper open subsets S1 S2 in X lt fails to cover X Characterization by continuous functions Edit Let X be a topological space and C X the ring of real continuous functions on X For each p X the evaluation map ev p C X R displaystyle operatorname ev p colon C X to mathbb R given by evp f f p is a ring homomorphism The kernel of evp is a maximal ideal since the residue field C X ker evp is the field of real numbers by the first isomorphism theorem A topological space X is pseudocompact if and only if every maximal ideal in C X has residue field the real numbers For completely regular spaces this is equivalent to every maximal ideal being the kernel of an evaluation homomorphism 14 There are pseudocompact spaces that are not compact though In general for non pseudocompact spaces there are always maximal ideals m in C X such that the residue field C X m is a non Archimedean hyperreal field The framework of non standard analysis allows for the following alternative characterization of compactness 15 a topological space X is compact if and only if every point x of the natural extension X is infinitely close to a point x0 of X more precisely x is contained in the monad of x0 Hyperreal definition Edit A space X is compact if its hyperreal extension X constructed for example by the ultrapower construction has the property that every point of X is infinitely close to some point of X X For example an open real interval X 0 1 is not compact because its hyperreal extension 0 1 contains infinitesimals which are infinitely close to 0 which is not a point of X Sufficient conditions EditA closed subset of a compact space is compact 16 A finite union of compact sets is compact A continuous image of a compact space is compact 17 The intersection of any non empty collection of compact subsets of a Hausdorff space is compact and closed If X is not Hausdorff then the intersection of two compact subsets may fail to be compact see footnote for example a The product of any collection of compact spaces is compact This is Tychonoff s theorem which is equivalent to the axiom of choice In a metrizable space a subset is compact if and only if it is sequentially compact assuming countable choice A finite set endowed with any topology is compact Properties of compact spaces EditA compact subset of a Hausdorff space X is closed If X is not Hausdorff then a compact subset of X may fail to be a closed subset of X see footnote for example b If X is not Hausdorff then the closure of a compact set may fail to be compact see footnote for example c In any topological vector space TVS a compact subset is complete However every non Hausdorff TVS contains compact and thus complete subsets that are not closed If A and B are disjoint compact subsets of a Hausdorff space X then there exist disjoint open set U and V in X such that A U and B V A continuous bijection from a compact space into a Hausdorff space is a homeomorphism A compact Hausdorff space is normal and regular If a space X is compact and Hausdorff then no finer topology on X is compact and no coarser topology on X is Hausdorff If a subset of a metric space X d is compact then it is d bounded Functions and compact spaces Edit Since a continuous image of a compact space is compact the extreme value theorem holds for such spaces a continuous real valued function on a nonempty compact space is bounded above and attains its supremum 18 Slightly more generally this is true for an upper semicontinuous function As a sort of converse to the above statements the pre image of a compact space under a proper map is compact Compactifications Edit Every topological space X is an open dense subspace of a compact space having at most one point more than X by the Alexandroff one point compactification By the same construction every locally compact Hausdorff space X is an open dense subspace of a compact Hausdorff space having at most one point more than X Ordered compact spaces Edit A nonempty compact subset of the real numbers has a greatest element and a least element Let X be a simply ordered set endowed with the order topology Then X is compact if and only if X is a complete lattice i e all subsets have suprema and infima 19 Examples EditAny finite topological space including the empty set is compact More generally any space with a finite topology only finitely many open sets is compact this includes in particular the trivial topology Any space carrying the cofinite topology is compact Any locally compact Hausdorff space can be turned into a compact space by adding a single point to it by means of Alexandroff one point compactification The one point compactification of R displaystyle mathbb R is homeomorphic to the circle S1 the one point compactification of R 2 displaystyle mathbb R 2 is homeomorphic to the sphere S2 Using the one point compactification one can also easily construct compact spaces which are not Hausdorff by starting with a non Hausdorff space The right order topology or left order topology on any bounded totally ordered set is compact In particular Sierpinski space is compact No discrete space with an infinite number of points is compact The collection of all singletons of the space is an open cover which admits no finite subcover Finite discrete spaces are compact In R displaystyle mathbb R carrying the lower limit topology no uncountable set is compact In the cocountable topology on an uncountable set no infinite set is compact Like the previous example the space as a whole is not locally compact but is still Lindelof The closed unit interval 0 1 is compact This follows from the Heine Borel theorem The open interval 0 1 is not compact the open cover 1 n 1 1 n textstyle left frac 1 n 1 frac 1 n right for n 3 4 does not have a finite subcover Similarly the set of rational numbers in the closed interval 0 1 is not compact the sets of rational numbers in the intervals 0 1 p 1 n and 1 p 1 n 1 textstyle left 0 frac 1 pi frac 1 n right text and left frac 1 pi frac 1 n 1 right cover all the rationals in 0 1 for n 4 5 but this cover does not have a finite subcover Here the sets are open in the subspace topology even though they are not open as subsets of R displaystyle mathbb R The set R displaystyle mathbb R of all real numbers is not compact as there is a cover of open intervals that does not have a finite subcover For example intervals n 1 n 1 where n takes all integer values in Z cover R displaystyle mathbb R but there is no finite subcover On the other hand the extended real number line carrying the analogous topology is compact note that the cover described above would never reach the points at infinity and thus would not cover the extended real line In fact the set has the homeomorphism to 1 1 of mapping each infinity to its corresponding unit and every real number to its sign multiplied by the unique number in the positive part of interval that results in its absolute value when divided by one minus itself and since homeomorphisms preserve covers the Heine Borel property can be inferred For every natural number n the n sphere is compact Again from the Heine Borel theorem the closed unit ball of any finite dimensional normed vector space is compact This is not true for infinite dimensions in fact a normed vector space is finite dimensional if and only if its closed unit ball is compact On the other hand the closed unit ball of the dual of a normed space is compact for the weak topology Alaoglu s theorem The Cantor set is compact In fact every compact metric space is a continuous image of the Cantor set Consider the set K of all functions f R displaystyle mathbb R 0 1 from the real number line to the closed unit interval and define a topology on K so that a sequence f n displaystyle f n in K converges towards f K if and only if f n x displaystyle f n x converges towards f x for all real numbers x There is only one such topology it is called the topology of pointwise convergence or the product topology Then K is a compact topological space this follows from the Tychonoff theorem Consider the set K of all functions f 0 1 0 1 satisfying the Lipschitz condition f x f y x y for all x y 0 1 Consider on K the metric induced by the uniform distance d f g sup x 0 1 f x g x displaystyle d f g sup x in 0 1 f x g x Then by Arzela Ascoli theorem the space K is compact The spectrum of any bounded linear operator on a Banach space is a nonempty compact subset of the complex numbers C displaystyle mathbb C Conversely any compact subset of C displaystyle mathbb C arises in this manner as the spectrum of some bounded linear operator For instance a diagonal operator on the Hilbert space ℓ 2 displaystyle ell 2 may have any compact nonempty subset of C displaystyle mathbb C as spectrum Algebraic examples Edit Compact groups such as an orthogonal group are compact while groups such as a general linear group are not Since the p adic integers are homeomorphic to the Cantor set they form a compact set The spectrum of any commutative ring with the Zariski topology that is the set of all prime ideals is compact but never Hausdorff except in trivial cases In algebraic geometry such topological spaces are examples of quasi compact schemes quasi referring to the non Hausdorff nature of the topology The spectrum of a Boolean algebra is compact a fact which is part of the Stone representation theorem Stone spaces compact totally disconnected Hausdorff spaces form the abstract framework in which these spectra are studied Such spaces are also useful in the study of profinite groups The structure space of a commutative unital Banach algebra is a compact Hausdorff space The Hilbert cube is compact again a consequence of Tychonoff s theorem A profinite group e g Galois group is compact See also EditCompactly generated space Compactness theorem Eberlein compactum Exhaustion by compact sets Lindelof space Metacompact space Noetherian topological space Orthocompact space Paracompact space Precompact set also called totally bounded Relatively compact subspace Totally boundedNotes Edit Let X a b N displaystyle mathbb N U a N displaystyle mathbb N and V b N displaystyle mathbb N Endow X with the topology generated by the following basic open sets every subset of N displaystyle mathbb N is open the only open sets containing a are X and U and the only open sets containing b are X and V Then U and V are both compact subsets but their intersection which is N displaystyle mathbb N is not compact Note that both U and V are compact open subsets neither one of which is closed Let X a b and endow X with the topology X a Then a is a compact set but it is not closed Let X be the set of non negative integers We endow X with the particular point topology by defining a subset U X to be open if and only if 0 U Then S 0 is compact the closure of S is all of X but X is not compact since the collection of open subsets 0 x x X does not have a finite subcover References Edit Compactness Encyclopaedia Britannica mathematics Retrieved 2019 11 25 via britannica com Engelking Ryszard 1977 General Topology Warsaw PL PWN p 266 a b Sequential compactness www groups mcs st andrews ac uk MT 4522 course lectures Retrieved 2019 11 25 Kline 1990 pp 952 953 Boyer amp Merzbach 1991 p 561 Kline 1990 Chapter 46 2 Frechet M 1904 Generalisation d un theorem de Weierstrass Analyse Mathematique Weisstein Eric W Compact Space mathworld wolfram com Retrieved 2019 11 25 Howes 1995 pp xxvi xxviii Kelley 1955 p 163 Bourbaki 2007 10 2 Theorem 1 Corollary 1 Bourbaki 2007 9 1 Definition 1 Arkhangel skii amp Fedorchuk 1990 Theorem 5 3 7 Willard 1970 Theorem 30 7 Gillman amp Jerison 1976 5 6 Robinson 1996 Theorem 4 1 13 Arkhangel skii amp Fedorchuk 1990 Theorem 5 2 3 Closed set in a compact space is compact at PlanetMath Closed subsets of a compact set are compact at PlanetMath Arkhangel skii amp Fedorchuk 1990 Theorem 5 2 2 See also Compactness is preserved under a continuous map at PlanetMath Arkhangel skii amp Fedorchuk 1990 Corollary 5 2 1 Steen amp Seebach 1995 p 67Bibliography EditAlexandrov Pavel Urysohn Pavel 1929 Memoire sur les espaces topologiques compacts Koninklijke Nederlandse Akademie van Wetenschappen te Amsterdam Proceedings of the Section of Mathematical Sciences 14 Arkhangel skii A V Fedorchuk V V 1990 The basic concepts and constructions of general topology In Arkhangel skii A V Pontrjagin L S eds General Topology I Encyclopedia of the Mathematical Sciences Vol 17 Springer ISBN 978 0 387 18178 3 Arkhangel skii A V 2001 1994 Compact space Encyclopedia of Mathematics EMS Press Bolzano Bernard 1817 Rein analytischer Beweis des Lehrsatzes dass zwischen je zwey Werthen die ein entgegengesetzes Resultat gewahren wenigstens eine reele Wurzel der Gleichung liege Wilhelm Engelmann Purely analytic proof of the theorem that between any two values which give results of opposite sign there lies at least one real root of the equation Borel Emile 1895 Sur quelques points de la theorie des fonctions Annales Scientifiques de l Ecole Normale Superieure 3 12 9 55 doi 10 24033 asens 406 JFM 26 0429 03 Bourbaki Nicolas 2007 Topologie generale Chapitres 1 a 4 Berlin Springer doi 10 1007 978 3 540 33982 3 ISBN 978 3 540 33982 3 Boyer Carl B 1959 The history of the calculus and its conceptual development New York Dover Publications MR 0124178 Boyer Carl Benjamin Merzbach Uta C 1991 A History of Mathematics 2nd ed John Wiley amp Sons ISBN 978 0 471 54397 8 Arzela Cesare 1895 Sulle funzioni di linee Mem Accad Sci Ist Bologna Cl Sci Fis Mat 5 5 55 74 Arzela Cesare 1882 1883 Un osservazione intorno alle serie di funzioni Rend Dell Accad R Delle Sci dell Istituto di Bologna 142 159 Ascoli G 1883 1884 Le curve limiti di una varieta data di curve Atti della R Accad Dei Lincei Memorie della Cl Sci Fis Mat Nat 18 3 521 586 Frechet Maurice 1906 Sur quelques points du calcul fonctionnel Rendiconti del Circolo Matematico di Palermo 22 1 1 72 doi 10 1007 BF03018603 hdl 10338 dmlcz 100655 S2CID 123251660 Gillman Leonard Jerison Meyer 1976 Rings of continuous functions Springer Verlag Howes Norman R 23 June 1995 Modern Analysis and Topology Graduate Texts in Mathematics New York Springer Verlag Science amp Business Media ISBN 978 0 387 97986 1 OCLC 31969970 OL 1272666M Kelley John 1955 General topology Graduate Texts in Mathematics Vol 27 Springer Verlag Kline Morris 1990 1972 Mathematical thought from ancient to modern times 3rd ed Oxford University Press ISBN 978 0 19 506136 9 Lebesgue Henri 1904 Lecons sur l integration et la recherche des fonctions primitives Gauthier Villars Robinson Abraham 1996 Non standard analysis Princeton University Press ISBN 978 0 691 04490 3 MR 0205854 Scarborough C T Stone A H 1966 Products of nearly compact spaces PDF Transactions of the American Mathematical Society 124 1 131 147 doi 10 2307 1994440 JSTOR 1994440 Archived PDF from the original on 2017 08 16 Steen Lynn Arthur Seebach J Arthur Jr 1995 1978 Counterexamples in Topology Dover Publications reprint of 1978 ed Berlin New York Springer Verlag ISBN 978 0 486 68735 3 MR 0507446 Willard Stephen 1970 General Topology Dover publications ISBN 0 486 43479 6 External links EditCountably compact at PlanetMath Sundstrom Manya Raman 2010 A pedagogical history of compactness arXiv 1006 4131v1 math HO This article incorporates material from Examples of compact spaces on PlanetMath which is licensed under the Creative Commons Attribution Share Alike License Retrieved from https en wikipedia org w index php title Compact space amp oldid 1128520092, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.