Base (topology)

From Wikipedia, the free encyclopedia

In mathematics, a base or basis for the topology τ of a topological space (X, τ) is a family B of open subsets of X such that every open set of the topology is equal to a union of some sub-family of B[1][2][3][4][5] (this sub-family is allowed to be infinite, finite, or even empty[note 1]). For example, the set of all open intervals in the real number line is a basis for the Euclidean topology on because every open interval is an open set, and also every open subset of can be written as a union of some family of open intervals.

Bases are ubiquitous throughout topology. The sets in a base for a topology, which are called basic open sets, are often easier to describe and use than arbitrary open sets.[6] Many important topological definitions such as continuity and convergence can be checked using only basic open sets instead of arbitrary open sets. Some topologies have a base of open sets with specific useful properties that may make checking such topological definitions easier.

Not all families of subsets form a base for a topology. For example, because X is always an open subset of every topology on X, if a family B of subsets is to be a base for a topology on X then it must cover X, which by definition means that the union of all sets in B must be equal to X. If X has more than one point then there exist families of subsets of X that do not cover X and consequently, they can not form a basis for any topology on X. A family B of subsets of X that does form a basis for some topology on X is called a base for a topology on X,[1][2][3] in which case this necessarily unique topology, call it τ, is said to be generated by B and B is consequently a basis for the topology τ. Such families of sets are frequently used to define topologies. A weaker notion related to bases is that of a subbasis for a topology. Bases for topologies are closely related to neighborhood bases.

Definition and basic properties[]

A base for a topology on X is a collection B of subsets of X satisfying the following properties:

  1. The elements of B cover X, i.e., every element of X belongs to some element in B.
  2. Given elements B1, B2 of B, for every x in B1 ∩ B2 there is an element B3 in B containing x and such that B3 is a subset of B1 ∩ B2.

An equivalent property is: any finite intersection[note 2] of elements of B can be written as a union of elements of B. These two conditions are exactly what is needed to ensure that the set of all unions of subsets of B is a topology on X.

If a collection B of subsets of X fails to satisfy these properties, then it is not a base for any topology on X. (It is a subbase, however, as is any collection of subsets of X.) Conversely, if B satisfies these properties, then there is a unique topology on X for which B is a base; it is called the topology generated by B. (This topology is the intersection of all topologies on X containing B.) This is a very common way of defining topologies. A sufficient but not necessary condition for B to generate a topology on X is that B is closed under intersections; then we can always take B3 = I above.

For example, the collection of all open intervals in the real line forms a base for a topology on the real line because the intersection of any two open intervals is itself an open interval or empty. In fact they are a base for the standard topology on the real numbers.

However, a base is not unique. Many different bases, even of different sizes, may generate the same topology. For example, the open intervals with rational endpoints are also a base for the standard real topology, as are the open intervals with irrational endpoints, but these two sets are completely disjoint and both properly contained in the base of all open intervals. In contrast to a basis of a vector space in linear algebra, a base need not be maximal; indeed, the only maximal base is the topology itself. In fact, any open set generated by a base may be safely added to the base without changing the topology. The smallest possible cardinality of a base is called the weight of the topological space.

An example of a collection of open sets which is not a base is the set S of all semi-infinite intervals of the forms (−∞, a) and (a, ∞), where a is a real number. Then S is not a base for any topology on R. To show this, suppose it were. Then, for example, (−∞, 1) and (0, ∞) would be in the topology generated by S, being unions of a single base element, and so their intersection (0,1) would be as well. But (0, 1) clearly cannot be written as a union of elements of S. Using the alternate definition, the second property fails, since no base element can "fit" inside this intersection.

Given a base for a topology, in order to prove convergence of a net or sequence it is sufficient to prove that it is eventually in every set in the base which contains the putative limit.

Examples[]

The set Γ of all open intervals in form a basis for the Euclidean topology on . Every topology τ on a set X is a basis for itself (that is, τ is a basis for τ). Because of this, if a theorem's hypotheses assumes that a topology τ has some basis Γ, then this theorem can be applied using Γ := τ.

A non-empty family of subsets of a set X that is closed under finite intersections of two or more sets, which is called a π-system on X, is necessarily a base for a topology on X if and only if it covers X. By definition, every σ-algebra, every filter (and so in particular, every neighborhood filter), and every topology is a covering π-system and so also a base for a topology. In fact, if Γ is a filter on X then { ∅ } ∪ Γ is a topology on X and Γ is a basis for it. A base for a topology does not have to be closed under finite intersections and many aren't. But nevertheless, many topologies are defined by bases that are also closed under finite intersections. For example, each of the following families of subset of is closed under finite intersections and so each forms a basis for some topology on :

  • The set Γ of all bounded open intervals in generates the usual Euclidean topology on .
  • The set Σ of all bounded closed intervals in generates the discrete topology on and so the Euclidean topology is a subset of this topology. This is despite the fact that Γ is not a subset Σ. Consequently, the topology generated by Γ, which is the Euclidean topology on , is coarser than the topology generated by Σ. In fact, it is strictly coarser because Σ contains non-empty compact sets which are never open in the Euclidean topology.
  • The set Γ of all intervals in Γ such that both endpoints of the interval are rational numbers generates the same topology as Γ. This remains true if each instance of the symbol Γ is replaced by Σ.
  • Σ = { [r, ∞) : r ∈ ℝ } generates a topology that is strictly coarser than the topology generated by Σ. No element of Σ is open in the Euclidean topology on .
  • Γ = { (r, ∞) : r ∈ ℝ } generates a topology that is strictly coarser than both the Euclidean topology and the topology generated by Σ. The sets Σ and Γ are disjoint, but nevertheless Γ is a subset of the topology generated by Σ.

Objects defined in terms of bases[]

  • The order topology is usually defined as the topology generated by a collection of open-interval-like sets.
  • The metric topology is usually defined as the topology generated by a collection of open balls.
  • A second-countable space is one that has a countable base.
  • The discrete topology has the singletons as a base.
  • The profinite topology on a group is defined by taking the collection of all normal subgroups of finite index as a basis of open neighborhoods of the identity.

The Zariski topology on the spectrum of a ring has a base consisting of open sets that have specific useful properties. For the usual basis of this topology, every finite intersection of basis elements is a basis element. Therefore bases are sometimes required to be stable by finite intersection.[citation needed]

  • The Zariski topology of is the topology that has the algebraic sets as closed sets. It has a basis formed by the set complements of algebraic hypersurfaces.
  • The Zariski topology of the spectrum of a ring (the set of the prime ideals) has a basis such that each element consists of all prime ideals that do not contain a given element of the ring.

Theorems[]

  • For each point x in an open set U, there is a base element containing x and contained in U.
  • A topology T2 is finer than a topology T1 if and only if for each x and each base element B of T1 containing x, there is a base element of T2 containing x and contained in B.
  • If are bases for the topologies then the set product is a base for the product topology In the case of an infinite product, this still applies, except that all but finitely many of the base elements must be the entire space.
  • Let B be a base for X and let Y be a subspace of X. Then if we intersect each element of B with Y, the resulting collection of sets is a base for the subspace Y.
  • If a function maps every base element of X into an open set of Y, it is an open map. Similarly, if every preimage of a base element of Y is open in X, then f is continuous.
  • A collection of subsets of X is a topology on X if and only if it generates itself.
  • B is a basis for a topological space X if and only if the subcollection of elements of B which contain x form a local base at x, for any point x of X.

Base for the closed sets[]

Closed sets are equally adept at describing the topology of a space. There is, therefore, a dual notion of a base for the closed sets of a topological space. Given a topological space a family of closed sets forms a base for the closed sets if and only if for each closed set and each point not in there exists an element of containing but not containing A family is a base for the closed sets of if and only if its dual in denoted by [7] is a base of open sets of that is, if and only if the family of complements of members of is a base for the open sets of

Let be a base for the closed sets of Then

  1. For each the union is the intersection of some subfamily of (that is, for any not in there is some containing and not containing ).

Any collection of subsets of a set satisfying these properties forms a base for the closed sets of a topology on The closed sets of this topology are precisely the intersections of members of

In some cases it is more convenient to use a base for the closed sets rather than the open ones. For example, a space is completely regular if and only if the zero sets form a base for the closed sets. Given any topological space the zero sets form the base for the closed sets of some topology on This topology will be the finest completely regular topology on coarser than the original one. In a similar vein, the Zariski topology on An is defined by taking the zero sets of polynomial functions as a base for the closed sets.

Weight and character[]

We shall work with notions established in (Engelking 1977, p. 12, pp. 127-128).

Fix X a topological space. Here, a network is a family of sets, for which, for all points x and open neighbourhoods U containing x, there exists B in for which Note that, unlike a basis, the sets in a network need not be open.

We define the weight, w(X), as the minimum cardinality of a basis; we define the network weight, nw(X), as the minimum cardinality of a network; the character of a point, as the minimum cardinality of a neighbourhood basis for x in X; and the character of X to be

The point of computing the character and weight is to be able to tell what sort of bases and local bases can exist. We have the following facts:

  • nw(X) ≤ w(X).
  • if X is discrete, then w(X) = nw(X) = |X|.
  • if X is Hausdorff, then nw(X) is finite if and only if X is finite discrete.
  • if B is a basis of X then there is a basis of size
  • if N a neighbourhood basis for x in X then there is a neighbourhood basis of size
  • if is a continuous surjection, then nw(Y) ≤ w(X). (Simply consider the Y-network for each basis B of X.)
  • if is Hausdorff, then there exists a weaker Hausdorff topology so that So a fortiori, if X is also compact, then such topologies coincide and hence we have, combined with the first fact, nw(X) = w(X).
  • if a continuous surjective map from a compact metrisable space to an Hausdorff space, then Y is compact metrisable.

The last fact follows from f(X) being compact Hausdorff, and hence (since compact metrisable spaces are necessarily second countable); as well as the fact that compact Hausdorff spaces are metrisable exactly in case they are second countable. (An application of this, for instance, is that every path in an Hausdorff space is compact metrisable.)

Increasing chains of open sets[]

Using the above notation, suppose that w(X) ≤ κ some infinite cardinal. Then there does not exist a strictly increasing sequence of open sets (equivalently strictly decreasing sequence of closed sets) of length ≥ κ+.

To see this (without the axiom of choice), fix

as a basis of open sets. And suppose per contra, that
were a strictly increasing sequence of open sets. This means

For

we may use the basis to find some Uγ with x in UγVα. In this way we may well-define a map, f : κ+κ mapping each α to the least γ for which UγVα and meets

This map is injective, otherwise there would be α < β with f(α) = f(β) = γ, which would further imply UγVα but also meets

which is a contradiction. But this would go to show that κ+κ, a contradiction.

See also[]

Notes[]

  1. ^ By a standard convention, the empty set, which is always open, is the union of the empty collection.
  2. ^ We are using a convention that the empty intersection of subsets of X is considered finite and is equal to X.

References[]

  1. ^ Jump up to: a b Bourbaki 1989, pp. 18–21.
  2. ^ Jump up to: a b Dugundji 1966, pp. 62–68.
  3. ^ Jump up to: a b Willard 2004, pp. 37–40.
  4. ^ Merrifield, Richard E.; Simmons, Howard E. (1989). Topological Methods in Chemistry. New York: John Wiley & Sons. p. 16. ISBN 0-471-83817-9. Retrieved 27 July 2012. Definition. A collection B of open subsets of a topological space (X,T) is called a basis for T if every open set can be expressed as a union of members of B.
  5. ^ Armstrong, M. A. (1983). Basic Topology. Springer. p. 30. ISBN 0-387-90839-0. Retrieved 13 June 2013. Suppose we have a topology on a set X, and a collection of open sets such that every open set is a union of members of Such a family of open sets is said to generate or define this topology. Then is called a base for the topology...
  6. ^ Adams & Franzosa 2009, pp. 46–56.
  7. ^ Narici & Beckenstein 2011, pp. 2–7.

Bibliography[]

  • Adams, Colin; (2009). Introduction to Topology: Pure and Applied. New Delhi: Pearson Education. ISBN 978-81-317-2692-1. OCLC 789880519.
  • Arkhangel’skij, A.V.; Ponomarev, V.I. (1984). Fundamentals of general topology: problems and exercises. Mathematics and Its Applications. 13. Translated from the Russian by V. K. Jain. Dordrecht: D. Reidel Publishing. Zbl 0568.54001.
  • Bourbaki, Nicolas (1989) [1966]. General Topology: Chapters 1–4 [Topologie Générale]. Éléments de mathématique. Berlin New York: Springer Science & Business Media. ISBN 978-3-540-64241-1. OCLC 18588129.
  • Bourbaki, Nicolas (1989) [1967]. General Topology 2: Chapters 5–10 [Topologie Générale]. Éléments de mathématique. 4. Berlin New York: Springer Science & Business Media. ISBN 978-3-540-64563-4. OCLC 246032063.
  • Dixmier, Jacques (1984). General Topology. Undergraduate Texts in Mathematics. Translated by Berberian, S. K. New York: Springer-Verlag. ISBN 978-0-387-90972-1. OCLC 10277303.
  • ; Mynard, Frederic (2016). Convergence Foundations Of Topology. New Jersey: World Scientific Publishing Company. ISBN 978-981-4571-52-4. OCLC 945169917.
  • Dugundji, James (1966). Topology. Boston: Allyn and Bacon. ISBN 978-0-697-06889-7. OCLC 395340485.
  • Engelking, Ryszard (1977). General Topology. Monografie Matematyczne. 60. Warsaw: PWN. Zbl 0373.54002.
  • (1983). Introduction to General Topology. New York: John Wiley and Sons Ltd. ISBN 978-0-85226-444-7. OCLC 9218750.
  • James Munkres (1975) Topology: a First Course. Prentice-Hall.
  • Munkres, James R. (2000). Topology (Second ed.). Upper Saddle River, NJ: Prentice Hall, Inc. ISBN 978-0-13-181629-9. OCLC 42683260.
  • ; (2011). Topological Vector Spaces. Pure and applied mathematics (Second ed.). Boca Raton, FL: CRC Press. ISBN 978-1584888666. OCLC 144216834.
  • Schechter, Eric (1996). Handbook of Analysis and Its Foundations. San Diego, CA: Academic Press. ISBN 978-0-12-622760-4. OCLC 175294365.
  • Schubert, Horst (1968). Topology. London: Macdonald & Co. ISBN 978-0-356-02077-8. OCLC 463753.
  • Willard, Stephen (2004) [1970]. General Topology. (First ed.). Mineola, N.Y.: Dover Publications. ISBN 978-0-486-43479-7. OCLC 115240.
  • Willard, Stephen (1970) General Topology. Addison-Wesley. Reprinted 2004, Dover Publications.
Retrieved from ""