topology (point-set topology, point-free topology)
see also differential topology, algebraic topology, functional analysis and topological homotopy theory
Basic concepts
fiber space, space attachment
Extra stuff, structure, properties
Kolmogorov space, Hausdorff space, regular space, normal space
sequentially compact, countably compact, locally compact, sigma-compact, paracompact, countably paracompact, strongly compact
Examples
Basic statements
closed subspaces of compact Hausdorff spaces are equivalently compact subspaces
open subspaces of compact Hausdorff spaces are locally compact
compact spaces equivalently have converging subnet of every net
continuous metric space valued function on compact metric space is uniformly continuous
paracompact Hausdorff spaces equivalently admit subordinate partitions of unity
injective proper maps to locally compact spaces are equivalently the closed embeddings
locally compact and second-countable spaces are sigma-compact
Theorems
Analysis Theorems
Uniform spaces were invented by André Weil, to capture a general notion of space for which it makes sense to speak of uniformly continuous maps. Such spaces include (pseudo)metric spaces and topological groups.
A uniform structure or uniformity on a set consists of a collection of global binary relations which allow us to say when one point is “-close” to another . For a metric space, we have for instance relations expressed by the formula , where is a positive number. For a topological group, we have relations , where is a neighborhood of the identity.
The definition described above is based on entourages; it was the original one published in Bourbaki and is still most commonly seen today. There is an equivalent definition based on uniform covers, although it is more difficult to generalize in some ways (e.g. to quasiuniform spaces). The relationship with gauge spaces (defined below) also allows for another definition. The original definition in Bourbaki can be seen as describing a contravariant functor from the category of finite sets to a category of filters of subsets.
A uniform structure, or uniformity, on a set consists of a collection of binary relations (called entourages or vicinities) satisfying some conditions. Write if is related to through ; then the conditions are the following:
In light of axiom (6), it follows that itself is an entourage.
A set equipped with a uniform structure is called a uniform space.
A collection of entourages satisfying (1–5) is a base for a uniformity; a base is precisely what generates a uniformity by taking the upward closure. A collection satisfying (1–3) is a preuniformity. Slightly more generally, we can replace each in the statement of these axioms with a finite intersection of entourages to get the concept of a subbase for a uniformity. A subbase is precisely what generates a base by closing up under finite intersections and precisely what generates a uniformity by closing up under finite intersections and taking the upward closure.
A uniform structure or uniformity could be represented as a family of binary relations on a set indexed by a set of entourages or vicinities, or equivalently by currying, as a ternary relation , where is the set of truth values, such that
for every entourage and ,
for every entourage , there exists an entourage such that for every , , and , if and , then .
For every entourage , there exists an entourage such that for every and , if , then .
There is an entourage and elements and such that .
If there are entourages and , then there is an entourage such that for every element and , if , then and .
If there is an entourage , then there is an entourage such that for all and , if , then .
An equivalent way to characterize a uniform space is by its collection of uniform covers. Here a cover of a set is a collection with union . For covers , we define:
refines , written , if every element of is a subset of some element of .
; this is also a cover.
For , .
.
We now define a covering uniformity on to be a collection of covers, called uniform covers, such that
If is a uniform cover, there exists a uniform cover such that .
There exists a uniform cover; in light of axiom (4), it follows that the cover is a uniform cover.
If are uniform covers, so is some cover that refines . In light of axiom (4), it follows that is a uniform cover.
If is a uniform cover and , then is a uniform cover.
The axioms (2–4) here roughly correspond (respectively) to the axioms (4–6) in the entourage definition. Axiom (1) takes on all of the real work; any collection of covers that satisfies it may be called a subbase (but not corresponding directly to a subbase in the previous definition), and then anything satisfying (1–3) is a base.
If is a uniform space defined in terms of entourages, we give it a covering uniformity by declaring a cover to be uniform if it is refined by for some entourage , where . Note that this does not mean that a uniform cover “consists of -sized sets” but only that it contains a subcover consisting of sets “no smaller than ”.
Conversely, given a covering uniformity, we define a base of entourages to consist of sets of the form for a uniform cover. That is, for each cover , we have a basic entourage such that iff for some . This defines a bijection between entourage uniformities and covering uniformities.
We give these in terms of entourages, but they could also be given directly in terms of uniform covers if desired.
The uniform topology induced by a uniformity is defined by taking the neighborhoods of a point to be sets of the form
where is an entourage. (Recall that a subset is open iff it is a neighborhood of every point that it contains.) We point out that different uniformities may give rise to the same topology (just as different metrics, even uniformly inequivalent ones, may give rise to the same topology).
In classical mathematics, the uniform topology is always regular and indeed completely regular. In constructive mathematics this is not necessarily true, and to reproduce classical results it is often useful or necessary to assume that uniform spaces are regular, completely regular, or satisfy an intermediate condition called uniform regularity.
Given uniform spaces and , a function is said to be uniformly continuous if for every entourage of , is an entourage of . Clearly, uniformly continuous functions are continuous with respect to the corresponding uniform topologies, but the converse is false (although see below the discussion in the case where is compact). There is an obvious concrete category of uniform spaces and uniformly continuous maps.
One feature of uniform space theory which is not available for general topological spaces is the possibility of taking (Cauchy) completions. The relevant definitions are straightforward:
A Cauchy net in a uniform space consists of a directed set and a function such that for every entourage , there exists such that whenever . (One can similarly define a Cauchy filter. This definition makes a uniform space into a Cauchy space.)
A Cauchy net converges to if for every entourage , there exists such that . (This makes sense for arbitrary nets or filters, but it can be proved that any convergent net is Cauchy. This definition makes a uniform space into a convergence space.)
A uniform space is complete if every Cauchy net/filter in converges (not necessarily to a unique point).
A uniform space is Hausdorff or separated if every convergent net/filter converges to a unique point, or equivalently if whenever for every entourage . (This is a purely topological concept.)
Every uniform space admits a Hausdorff completion , i.e., there is a uniformly continuous map (an embedding if is Hausdorff, dense? in any case), characterized by the following universal property:
In short, the category of complete Hausdorff uniform spaces is a reflective subcategory of .
One can also define a (not necessarily Hausdorff) completion of by replacing the image of in by itself, but this does not have as nice properties; in particular, complete uniform spaces do not form a reflective subcategory of .
Every uniform space also has an underlying proximity (defined there), and the resulting functor has a fully faithful right adjoint which identifies proximity spaces with uniform spaces that are totally bounded.
Uniform spaces can also be identified with syntopogenous spaces that are both perfect and symmetric; see syntopogenous space.
Every uniform space has a inequality relation (an irreflexive symmetric relation) where “” means that there exists an entourage such that . If is uniformly regular, then this is an apartness relation, i.e. it is also a comparison. This inequality is tight exactly when the uniform space is Hausdorff.
Uniform spaces and uniform maps form a topological category.
In particular, limits and colimits of uniform spaces are computed similarly to the case of topological spaces: take the limit/colimit of underlying sets and equip it with the final/initial uniform structure.
See Chapter 21 in Joy of Cats for more information.
Every metric space (or more generally any pseudometric space) is a uniform space, with a base of uniformities indexed by positive numbers . (You can even get a countable base, for example by using only those equal to for some integer .) Define to mean that (or if you prefer). Then axiom (2) may be proved by using ; similarly, every metric space is uniformly regular in constructive mathematics, which may be proved by using any positive number less than and applying the comparison law. (The other axioms are easy.) Every quasi(pseudo)metric space is a quasiuniform space in the same way. We can also generalise from metric spaces to gauge spaces; see under Variations below.
Every topological group is a uniform space, with a base of uniformities indexed by neighbourhoods of the identity element, in two ways: left and right. (These two ways agree for abelian groups, of course; they also agree for compact groups, by the general theorem below for uniformities on compact spaces. I wonder if that has anything to do with Haar measure?) In particular, any Banach space or Lie group is a uniform space. Define to mean that (or for the other way). Then axiom (2) may be proved by invoking the continuity of multiplication; constructively, we cannot prove that every topological group is uniformly regular, although this can be proved for the classical examples of Lie groups and topological vector spaces (TVSs).
These are in a way the motivating examples. The theory of uniformly continuous maps was developed first for metric spaces, but it was noticed that, for a metrisable TVS, it could be described entirely in terms of the topology and the addition, which immediately generalised to non-metrisable TVSs. The theory of uniform spaces covers both of these (and their generalisations to pseudometric spaces and topological groups) at once.
We can also form certain uniformities on function spaces. For instance, if is a uniform space and is any set, then the set of functions has a “uniformity of uniform convergence”; this is generated by a base of entourages consisting of, for each entourage of , the generating entourage
This can be defined for any set , but it is best-behaved when has some compactness properties. For instance, in constructive mathematics, to prove uniform regularity of from uniform regularity of we seem to need a condition such as that is a covert set.
A first wave of results concerns separation axioms:
As a matter of fact, a significant theorem is that a topology (on a given set) is the uniform topology for some uniformity if and only if it is completely regular. See also the discussion below on the relation with metric and pseudometric spaces.
For a uniform space, the following separation conditions are equivalent: (the topology distinguishes points), (points are closed), (Hausdorff), (regular Hausdorff), (completely regular Hausdorff). Of course, this follows from the fact that it is completely regular.
The category of uniform spaces admits arbitrary small products (which are preserved by the forgetful functors to Top and to Set). Hence it is not generally true that uniform spaces are normal (so that separated ones would be ), because for instance an uncountable power of the real line (with its usual topology) is not a normal space.
A second wave of results relates uniform spaces to pseudometric spaces (like metric spaces, but dropping the separation axiom that implies ):
By this result, the results mentioned above on completions of uniform spaces may be proved by appeal to similar results for (pseudo)metric spaces.
We also mention the special case of compact spaces:
It follows then that if are uniform spaces with compact Hausdorff and if is continuous, then is uniformly continuous.
Finally, uniform spaces and uniformly continuous maps form a category . This category admits a closed monoidal structure in which the function-space is the set of uniformly continuous maps with the uniformity of uniform convergence (mentioned above). The corresponding tensor product is called the “semi-uniform product”, and is not symmetric, because the evaluation map is not generally uniformly continuous. See Isbell, Chapter III, particularly Theorem 26 p46
On the other hand, also has a different closed symmetric monoidal structure in which the internal-hom is the space of uniformly continuous functions with the uniformity of pointwise convergence, i.e. the subspace uniformity induced by the product uniformity on . See Isbell, exercise III.7, p53
Some authors insist that a uniform space must be separated; this can be arranged directly in the definition (that is without reference to the uniform topology or to the concept of convergent net/filter) by adding the following axiom, a sort of converse axiom (1):
This makes the discussion of completions slightly simpler.
If the symmetry axiom (3) in the entourage definition is dropped, then the result is a quasiuniform space. Quasiuniform spaces are related to quasi(pseudo)metrics in the same way as uniform spaces are related to (psuedo)metrics. Perhaps surprisingly, every topological space is quasiuniformisable. (It is rather triciker to define quasiuniform spaces in terms of covers, but technically possible using covers by pairs of sets; see Gantner and Steinlage.)
A set with a preuniformity or subbase (only axioms 1-3) is a preuniform space.
A gauge space consists of a set and a collection of pseudometrics on ; one usually requires to be a filter. A gauge space defines a uniform space (necessarily uniformly regular) by taking one basic entourage for each pseudometric in and each positive number ; conversely, every uniform space arises in this way, with the pseudometrics in the gauge being those that are uniformly continuous as maps on the product space. However, gauge spaces form a category with a stricter notion of morphism, in which the categories (of metric spaces and short maps) and (of uniform spaces and uniformly continuous maps) are both full subcategories. A quasigauge space consists of a set and a collection of quasipseudometrics; every quasiuniform space arises from a quasigauge space.
In weak foundations of mathematics, the theorems above may not be provable. In particular, the theorem that every uniform space arises from a gauge space is equivalent (internal to an arbitrary topos with a natural numbers object) to dependent choice (plus excluded middle if you don't require the uniform space to be uniformly regular). If the concept is to be applied to analysis, then it may be best to define a uniform space as a gauge space satisfying a saturation condition.
There is also a “pointless” notion of uniform space, called a uniform locale.
The really critical axioms are (1–3): a collection of binary relations which satisfies those three axioms is a subbase (although not every subbase takes this form), from we can generate a uniform structure straightforwardly. Indeed, (4–6) simply state that the entourages form a filter, so generating a uniformity from a base or subbase is simply the usual generation of a filter from a base or subbase.
We draw particular attention to axiom (2), which may be called an “” principle. It generalizes a principle familiar from analysis in metric spaces, where one establishes by showing there exists such that and , and applying the triangle inequality. The utility of this principle for metric spaces, extrapolated in this way, gives uniform spaces much of their power.
For full power in constructive mathematics, we also need uniform regularity, which may similarly be called a “something less than ” principle. (That is, for any there is an such that any two points are either -close or -far; classically we may take to be , but constructively it's better to think of .) This can actually be combined with axiom (2) into a single statement, as you might expect since , but that makes the intuition less clear.
Axiom (1) is a nullary version of axiom (2); together they prove that, given any entourage and any integer , there exists an entourage whose -fold composite is contained in . The symmetry axiom (3) then allows one to take the opposite of at any point in the composite as well.
Altogether, these may be seen as axiomatising the notion of approximate equivalence. If is an approximate equivalence relation, then we might expect it to be
One could stop there, but this is not a very useful notion of approximation. Instead we generalise to a family of approximate equivalence relations and impose the principle to allow them to be used. This is nearly the definition of uniform space; in particular, the axiom (1) states precisely that each entourage is reflexive. The symmetry axiom (3) in the standard definition is weaker than requiring each individual entourage to be symmetric, but that is not an essential change; every uniformity has a base consisting of its symmetric entourages. The final three axioms have already been explained as a closure condition; they force equivalent uniformities on a given set (in the sense that the identity function on the set is uniformly continuous either way) to be equal.
The idea that a uniformity is an “approximate equivalence relation” can be made precise as follows. A preorder is the same as a category enriched over the poset of truth values, and it is an equivalence relation if and only if this category is symmetric. In fancier language, a preorder is a monoid (or monad) in the bicategory . A quasi-uniform space can then be identified with a monoid in the bicategory , whose hom-categories are the categories of pro-objects in the hom-categories of , aka filters. Of course, it is a uniform space just when it is also symmetric. See also prometric space.
In all these cases, in order to recover the correct notion of morphism abstractly, we must consider monoids in a double category or equipment rather than merely a bicategory.
With a uniform structure on a set associate a contravariant functor from the category of finite sets to a category of filters of subsets as follows. Call a subset of big iff there is an entourage such that provided for each . For a metric space, this is the filter of -neighbourhoods of the diagonal. This defines a contravariant functor from the category of finite sets to the category of filters of subsets. Axiom 1 says that there is a unique big subset in . Axiom 2 says that the filter on is the pullback of the obvious maps . To give a uniformly continuous map (morphism) of uniform structures is the same as to give a natural transformation . Indeed, such a natural transformation is determined by the function , and continuity of mean uniform continuity.
With a filter on associate the functor as follows: is . A filter on is Cauchy iff the obvious map is well-defined.
complete uniform space and uniform completion
Generalized uniform structures
André Weil, Sur les espaces à structure uniforme et sur la topologie générale, Actualités Sci. Ind. 551, Paris, (1937).
John Kelley, General Topology, GTM 27, 1955.
Nicolas Bourbaki, Uniform Structures, Chapter II in: General topology, Elements of Mathematics, Springer (1971, 1990, 1995) [doi:10.1007/978-3-642-61701-0]
Eric Schechter, Handbook of Analysis and its Foundations (1996)
Warren Page, Topological Uniform Structures, Dover
I.M. James, Topologies and Uniformities, Springer
Norman Howes, Modern analysis and topology, Springer
P. Samuel, Ultrafilters and compactifications of uniform spaces, Trans. Amer. Math. Soc. 64 (1948) 100–132
J. R. Isbell, Uniform spaces, Math. Surveys 12, Amer. Math. Soc. 1964
T. E. Gantner and R. C. Steinlage, Characterizations of quasi-uniformities, J. London Math Soc (2) 5 (1972) (defines quasi-uniformities using covers)
Last revised on May 24, 2023 at 12:13:56. See the history of this page for a list of all contributions to it.