Let be a set. A topology on E is a subset of such that:
and .
For all one has .
For all one has .
A pair is a topological space when is a set and is a topology on . Moreover, a subset of is called open, if , and closed, if .
The section on fundamental examples collects many examples of topologies. Note that and both are obvious topologies on any set , but of course the examples in the next section are more substantial. For now, we will work out a few basic properties of topologies.
Let be a (nonempty) set, be a topological space, and a function. Define:
Then is a topological space. One calls the initial topology for or the topology induced by .
By construction, and so . Let . Then by definition, there exists such that for . Thus:
Hence by definition, . Last, let . By definition, each element of is the preimage by of some open set in . We set:
Now if and only if there exists such that . Now, if and only if there exists such that . Hence if and only if there exists such that , which in turn is equivalent to . Hence:
and therefore, we have shown that is a topology on . ∎
Let and be two topological spaces. Let be a function. We say that is continuous on when:
It is immediate that is continuous if and only if for all closed subset , the set is closed as well.
It is immediate by definition that any constant function is continuous. It is also obvious that the identity function on any topological space is continuous.
Let , and be three topological spaces. Let and . If and are continuous, so is .
Let . Then by continuity of . Hence by continuity of . ∎
The category whose objects are topological spaces and morphisms are continuous functions is the category of topological spaces.
An isomorphism in the category of topological spaces is called an homeomorphism.
An homeomorphism is by definition a continuous map which is also a bijection and whose inverse function is also continuous. There are continuous bijections which are not homeomorphisms (see the chapter Fundamental Examples).
The initial topology induced by a function is a subset of the topology on if and only if is continuous. This motivates the following definition.
Let be a set. Let and be two topologies on . We say that is finer than and is coarser than when .
Of course, inclusion induces an order relation on topologies on a given set. A remarkable property is that any nonempty subset of the ordered set of topologies on a given set always admits a greatest lower bound.
Let be a set. Let be a nonempty set of topologies on . Then the set:
is a topology on and it is the greatest lower bound of (where the order between topology is given by inclusion).
We first show that is a topology. By definition, for all , since is a topology on , we have . Hence . Now, let . Let be arbitrary. By definition of , we have and . Therefore, since is a topology. Since was arbitrary in , we conclude that by definition. The same proof can be made for unions. Let . Let be arbitrary. By definition of , we have . Since is a topology, . Hence, as was arbitrary, . So is a topology on . By construction, for all , so is a lower bound for . Assume given a new topology on such that for all in . Let . Then for all we have . Hence by definition . So and thus is the greatest lower bound of . ∎
Let be a set and be a topological space. The smallest topology which makes a function continuous is .
Let be the set of all topologies on such that is continuous. By definition, is a lower bound of . Moreover, . Hence it is the greatest lower bound. ∎
We can use Theorem (1.11) to define other interesting topologies. Note that trivially is a topology, so given any for some set , there is at least one topology contaning . From this:
Let be a set. Let be a subset of . The greatest lower bound of the set:
is the smallest topology on containing . We call it the topology induced by on and we denote it .
We note that since belongs to by construction, it is indeed the smallest topology containing .
In general, we will want a better description of the topology induced by a particular set than the general intersection above. This is not always possible, but the concept of basis allows one to obtain useful descriptions of topologies.
When inducing a topology from a family of subsets of some set , the fact that enjoys the following property greatly simplifies our understanding of .
Let be a set. A topological basis on is a subset of such that:
,
For all such that , we have:
The main purpose for this definition stems from the following theorem:
Let be some set. Let be a topological basis on . Then the topology induced by is:
Denote, for this proof, the set , by , and let us abbreviate by . We wish to prove that . First, note that by construction. By definition, and since is a topology, it is closed under arbitrary unions. Hence . To prove the converse, it is sufficient to show that is a topology. As it contains and is the smallest such topology, this will provide us with the inverse inclusion. By definition, and thus . By assumption, since is a basis, so . As the union of unions of elements in is a union of elements in , is closed under abritrary unions. Now, let be elements of . If then . Assume that and are not disjoints. Then by definition of a basis, for all there exists such that and . So:
and therefore, by definition, . We conclude that the intersection of two arbitary elements in is again in by using the distributivity of the union with respect to the intersection. ∎
The typical usage of this theorem is the following corollary. We shall say that a basis on a set is a basis for a topology on when the smallest topology containing is .
Let be a topological basis for a topology on . A subset of is in if and only if for any there exists such that and .
We showed that any open set for the topology is a union of elements in : hence if for then there exists such that and . Conversely, if is some subset of such that for all there exists such that and then and thus . ∎
As a basic application, we show that:
Let and be two topological spaces. Let be a basis for the topology . Let . Then is continuous on if and only if:
Let and be two topological spaces. Let be a basis for the topology . Let . Then is continuous on if and only if:
We leave to the reader to write the statement when both and have a basis.
Given a topological space , an arbitrary subset of may be neither open nor closed. It is useful to find closest “approximations” for which are either open or closed. The proper notions are as follows.
Let be a topological space. Let . There exists a (necessarily unique) largest open set in contained in and a (necessarily unique) smallest closed set in containing (where the order is given by inclusion).
Let . Then since is a topology, . Moreover, by definition, if then there exists such that and since we conclude . Therefore is by construction the largest open subset of contained in .
The reasonning is similar for the smallest closed set. Namely, let . Since closed sets are the complements of open sets, it follows that arbitrary intersections of closed sets is closed. Thus is a closed set, and it is the smallest containing by construction. ∎
Let be a topological space. Let . The interior of is the largest open subset of contained in denoted by . The closure of is the smallest closed subset of containing and is denoted by .
Thus we always have . Moreover:
Let be a topological space. Let . Then:
If then and ,
,
.
If then . Since is open and a subset of , it is contained in by definition. Similarly, if then . Now, is closed and contains so it must contained the smallest closed subset containing , namely .
Note that so . Similarly with , so . On the other hand, is an open set (as the intersection of two open sets), and since and , it is included in . Hence it is included in the largest open subset contained in , which completes this proof.
The same reasonning can be applied to the last assertion of this proposition. ∎
The main theorem about closures is the following.
Let be a topological space. Let . Then if and only if for all such that we have .
Assume first that . Then and is open by definition; moreover by construction since so . Conversely, let and assume that there exists such that and yet . Then and is closed, so by definition , hence . ∎
In general, one can thus write:
The following is immediate, so we omit the proof.
Let be a topological space and a basis for . Let . Then if and only if for all such that we have .
We can use the main theorem about closure to show that:
Let be a topological space. Let . Then and .
We shall only prove that as both assertions are proved similarly. Note first that so and is closed. Hence . Conversely, let . Let such that . Assume . Then so, as open, , which contradicts . So , and thus as desired. ∎
Let and be topological spaces. Let be given. Then is continuous if and only if for all , .
Assume first that is continuous. Let . Since is continuous, is closed, and by definition of images and preimages,
so as desired.
Let us now assume that for any , we have . Then let be a closed set. Let . By assumption, . So , hence so is closed. Hence is continous. ∎
In the next chapter on fundamental examples, one will encounter many topological spaces where singletons are not open. Note that in those cases, if is constant, then would be empty, and thus we can not expect in general for continuous functions.
Let and be topological spaces. Let be given. Then is continuous if and only if for all , we have .
Assume first that is continuous. Let . Then since is continuous. Since , we conclude that (by maximality of among all open sets in contained in ).
Conversely, assume that for all , . Let . Then by assumption:
so i.e. . Since was arbitrary, is continuous. ∎
Let and be topological spaces. Let be given. Then is continuous if and only if for all we have .
Assume first that is continuous. Let . Then so . Since is continuous, then is closed, and thus by minimality of , we have .
Conversely, assume that for all we have . Let be closed. Then:
so i.e. is closed. Since was an abritrary closed subset of , the function is continuous. ∎
Continuity was phrased by stating the topology induced by the function is coarser than the topology on the domain. We can extend this idea to define topologies. Note that if is endowed with the indiscrete topology then any function, to any topological space, is continuous. Hence the following definition can be made:
Let be a set. Let be a topological space. Let be a nonempty set of functions from to . The smallest topology on such that all the functions in are continuous is the initial topology induced by on . We denote it by .
Let be a set. Let be a topological space. Let be a nonempty set of functions from to . Let:
where
for all -tuple of functions in and . Then is a basis for the initial topology induced by .
Note that for all , we must have by defintion of continuity. Hence, if is any topology on such that all the functions in are continuous, then . In particular, . On the other hand, by construction, every function in is continuous for , so by definition of the initial topology. Hence .
It remains to show that is a basis on . By definition, note that since for any . Now, note that by definition, the intersection of two elements in is still in , so is a trivially a basis. ∎
The main two theorems regarding initial topologies describe its universal property:
Let be a set. Let and be topological spaces. Let be a nonempty set of functions from to . A function is continuous when is endowed with the initial topology if and only if is continuous for all .
If is continuous, then for all the function is continuous, as the composition of two continuous functions.
Conversely, assume that is continuous for all . Let and and set:
Now, since are all continuous by assumption, are all in . Hence . Since was an arbitrary set in a basis for the final topology on , is continuous. ∎
Let be a set. Let and be topological spaces. Let be a nonempty set of functions from to . The initial topology for is the unique topology such that, given any topological space and given any function , then is continuous if and only if is continuous for all .
We showed that the initial topology for has the desired property. Conversely, assume that the given property holds for some topology on . In particular, let be the identity, seen as a function from the topological space into the space . Using the specified universal property, we see that is continuous, so is finer than . Conversely, note that is also continuous for the same reason, so is finer than , so these two topologies agree. This completes the proof of this universal property. ∎
Last, we can obtain a basis for initial topologies in a natural way:
Let be a set, be a topological space, a nonempty set of functions from to . Let be a basis for . Then the set:
is a basis for the initial topology for .
This is a quick computation. ∎
The dual notion of initial topology is final topology:
Let be a set. Let be a set of triplets where is a topological space, and is a function. The final topology on is the smallest topology such that all fuctions such that are continuous.
In the following chapters, we will see that the product topology, the trace topology and the norm topology are examples of initial topologies. The quotient topology is an example of a final topology.
This chapter provides various examples of topological spaces which will be used all along these notes and are often at the core of the subject.
Let be a set. The topology is the indiscrete topology on . The topology is the discrete topology on .
Let be a topological space. All functions from into are continuous. The only continuous functions from to are constant if is T1.
When is given the discrete topology, then for all open subsets of one has open in . On the other hand, if is given the indiscrete topology and is T1 then assume takes two values, and . Then is open, so must be open, and as it is nonempty (it contains ), it is all of , which is absurd. ∎
Let be a linearly ordered set. Let be two symbols not in . Define and extend to by setting for all :
Let be the relation (i.e. ) on . Let . We define the set .
Let be a linearly ordered set. Then the topology induced by the set:
is the order topology on .
The set is a basis for the order topology on .
Since is linearly ordered, so is . It is immediate that if is the largest of and and is the smallest of and . ∎
The default topology on is the order topology.
Let be an ordered set, and . We define .
Let be a topological space. Let . Then the trace topology on induced by is the topology:
It is a trivial exercise to show that the above definition indeed gives a topology on . ∎
Just as easy is the following observation:
Let be a topological space. Let be a basis for . Let . Then the set is a basis for the trace topology on induced by .
Trivial exercise. ∎
The default topology on , and is the trace topology induced by the order topology on . Since for all , we see that the natural topologies form and are in fact the discrete topology.
Let , where is a topological space. Let be the inclusion map. Then the trace topology is the initial topology for , i.e. .
Let be some nonempty set. Let us assume given a family of topological spaces. A basic open set of the cartesian product is a set of the form where is finite and for all , we have .
Let be some nonempty set. Let us assume given a family of topological spaces. The product topology on is the smallest topology containing all the basic open sets.
Let be some nonempty set. Let us assume given a family of topological spaces. The collection of all basic open sets is a basis on the set .
Trivial exercise. ∎
The product topology is not just the basic open sets on the cartesian products: there are many more open sets!
Let be some nonempty set. Let us assume given a family of topological spaces. The product topology on is the initial topology for the the set where is the canonical surjection for all .
Fix . Let . By definition, where for , and . Hence is open in the product topology. As was an arbitrary open subset of , the map is continuous by definition. Hence, as was arbitrary in , the initial topology for is coarser than the product topology.
Conversely, note that the product topology is generated by , so it is coarser than the initial topology for . This concludes this proof. ∎
Let be some nonempty set. Let us assume given a family of topological spaces. Let be the product topology on . Let be a topological space. Then is continuous if and only if is continuous from to for all , where is the canonical surjection on for all .
We simply applied the fundamental property of initial topologies. ∎
The box topology on the cartesian product is the smallest topology containing all possible cartesian products of open sets. It is finer than the product topology in general. Since the product topology is the coarsest topology which makes the canonical projections continuous, it is the preferred one on cartesian products. Of course, both agree on finite products.
The product topology is the default topology on a cartesian product of topological spaces.
Let be a set. A function is a distance on when:
For all , we have if and only if ,
For all we have ,
For all we have .
A pair is a metric space when is a set and a distance on .
The following is often useful:
Let (E,d) be a metric space. Let . Then:
Since we have . Since we have . Hence the proposition holds. ∎
Let be a metric space. Let and . The open ball of center and radius in is the set:
Let be a metric space. The metric topology on induced by is the smallest topology containing all the open balls of .
Let be a metric space. The set of all open balls on is a basis for the metric topology on induced by .
It is enough to show that the set of all open balls is a basis. By definition, . Now, let us be given and for some and . If the intersection of these two balls is empty, we are done; let us assume that there exists . Let be the smallest of and . Let . Then:
so . Similarly, . Hence, as desired. ∎
The following theorem shows that metric topologies are minimal in the sense of making the distance functions continuous.
Let be a metric space. For all , the function is continuous on for the metric topology. Moreover, the metric topology is the smallest topology such that all the functions in the set are continuous.
Fix . It is sufficient to show that the preimage of and by is open in the metric topology of , where is arbitrary. Indeed, these intervals form a basis for the topology of . Let be given. Then by definition, so it is open. Moreover, it shows that the minimal topology making all these maps continuous must indeed contain the metric topology. Now, let such that . Let . Then if for some then:
so . Hence
for all . Therefore, is open, as desired, and our proposition is proven. ∎
The topology on is the trace topology on induced by the usual, i.e. the order topology on .
The metric topology is the default topology on a metric space.
There are more examples of continuous functions between metric spaces. More precisely, a natural category for metric spaces consists of metric spaces and Lipschitz maps as arrows, defined as follows:
Let , be metric spaces. A function is -Lipschitz for if:
Let , be metric spaces. Let be a Lipschitz function. Then the Lipschitz constant of is defined by:
if and only if is constant.
Let , be metric spaces. If is a Lipschitz function, then it is continuous.
Assume is nonconstant (otherwise the result is trivial). Let be the Lipschitz constant for . Let and . Let . Let such that (note that the upper bound is nonzero).
(2.1) | |||||
(2.2) | |||||
(2.3) |
Hence . So is continuous. ∎
The proof of continuity for Lipshitz maps can be simplified: it is a consequence of the squeeze theorem. We refer to the chapter on metric spaces for this.
Using Lipshitz maps as morphisms for a category of metric spaces is natural. Another, more general type of morphisms, would be uniform continuous maps, which are discussed in the compact space chapter.
A potential source for counter-examples, the family of cofinite topologies is easily defined:
Let be a set. Let:
Then is a topology on .
By definition, . Moreover, which is finite, so . Let . If or is empty then so . Otherwise, which is finite, since by definition and are finite. Hence . Last, let . Again, if then . Let us now assume that contains at least one nonempty set . Then:
Since is finite by definition, so is , which is therefore in . This completes our proof. ∎
Limits of sequences is a central tool in topology and this section introduces the natural topology for this concept. The general notion of limit is the subject of the next chapter.
Let be some symbol not found in . We define to be .
The set:
is a topology on .
By definition, so . Moreover which has cardinal so . Let . If either or is a subset of then is a subset of so . Othwiwse, . Yet which is finite as a finite union of finite sets. Hence again.
Last, assume that . Of course, if and only if for some . So, if then by definition. If, on the other hand, , then there exists with finite. Now, so it is finite, and thus again . ∎
The general definition of topology allows for examples where elements of a topological space, seen as a set, can not be distinguished from each others (for instance if the topology is indiscrete). When points can be topologically differentiated, a topology is in some sense separated. There are many axioms, or definitions, of separability, and we will use the most common and intuitive: namely, the notion of Hausdorff spaces. We do however present a few basic notions in this section which are weaker than Hausdorff separation, as such spaces are certainly common in mathematics.
Let be a topological space. We say that is T0 when given any two points with , there exists an open set such that either or .
Thus a space is T0 when there are enough open sets to separate the points, i.e. when the set of all open sets containing one point is not the same as the set of all open sets containing a different point. However, this notion is not symmetric. The following definition add that the separation property should be symmetric:
Let be a topological space. We say that is T1 when given any two points with , there exists two open sets such that and .
A key advantage of T1 separation is:
Let be a topological space. Then is T1 if and only if for all the set is closed.
Assume is T1. Fix . Let . Then there exists such that and . We can thus write:
which shows that is open as desired.
Assume now that all singletons are closed. Let with . Then and , i.e. is T1. ∎
Note that T0 is not enough for the above result.
Let be a topological space. is T1 if and only if for all we have .
Assume is T1. Let . Let . If then there exists such that and . This is a contradiction. So .
Assume now that for all we have . Let such that . Then:
, so there exists such that and . Similarly:
so there exists such that and . Hence is T1. ∎
The indiscrete topology is not T0.
Let be an infinite set (for instance ), endowed with the cofinite topology . By defintion, is closed for all so is T1. We make a useful observation. Let be a continuous map, where the codomain is endowed with the usual order topology. Assume is not constant: then there exists such that . Without loss of generality, we assume . Let . Let and . Then and are open sets in such that and . Since is continuous, and are open in , i.e. are cofinite. Since and are disjoint, we conclude that is in the complement of , which is finite. Since the complement of is finite as well, we conclude that is finite, which is a contradiction. So is a constant.
The cofinite topology on infinite set example shows that T1 still allows for counter intuitive situations. We also saw that we could find two disjoint open sets in containing given distinct points: this stronger property is the separation axiom we will focus our attention to.
Let be a topological space. We say that is a Hausdorff space (or T2) when for any such that there exists such that , and .
If is Hausdorff, then it is T1.
This result holds by definition. ∎
Let be a linearly ordered set and let be the associated order topology on . Then is Hausdorff. Indeed, let with ; without loss of generality we assume that . Then if there exists such that then and , where both intervals are disjoint and open. Otherwise, and ,.
In particular, is Hausdorff.
Let be a Hausdorff space. Then by definition, it is immediate that the trace topology induced on any subset of is also Hausdorff.
In particular, the topology on is Hausdorff.
The discrete topology on any set is always Hausdorff.
In particular, the topology on and is Hausdorff.
Let be a metric space. Then is Hausdorff. Indeed, let such that . Let and note that by definition of a distance. Let . Then:
and by symmetry, if then . Hence .
Let be some family of Hausdorff topological spaces. Then the cartesian product with the product topology is Hausdorff. Indeed, let and such that . By definition, there exists such that . Since is Hausdorff, there exists such that , and . Define the two families:
and
and set and . Then by construction, , and .
The following result is a characterization of Hausdorff separation.
Let be a topological space. Then is Hausdorff if and only if is closed in for the product topology.
Assume is Hausdorff. Let (so !). Then there exists two disjoint open sets and in such that and . If then and which is impossible since and are disjoint. So , which is open in the product topology, contains by definition, and is a subset of . Hence is open.
Conversely, assume is open. Let . Since basic open sets form a basis for the product topology, there exists open in such that and . Now, by definition, so as above, , as desired. So is Hausdorff. ∎
Hausdorff spaces have a nice relation with continuous maps as well.
Let be a topological space. Let be a Hausdorff topological space. Let and be given.
If is continuous, then the kernel of is closed.
If is continuous, then the graph of is closed.
If and are continuous, then the equalizer of is closed.
Since is Hausdorff, the set is closed.
We first prove that the kernel of is closed when is continuous. Let:
The function is continuous since the product topology is the initial topology for the canonical surjections, and is continuous by assumption, so is continuous as the composition of two continuous functions. Similarly, is continuous. Hence, by the universal property of the initial topology, is continuous.
Now, by definition, and thus it is closed.
We now prove that the graph of is closed when is continuous. Let:
By a similar argument as for , the map is continuous since and are continuous for the product topology, and is continuous, so by composition of continuous functions, is continuous. Hence by the universal property of the product topology (seen again as the initial topology for the canonical surjections), is continuous.
Now so the graph of is closed.
Last, we show the equalizer of and is closed when both and are continuous. Let:
By assumption, and are continuous, so by the universal property of the product topology, is continuous. So is closed. ∎
Let and be two topological spaces. Let .
If open (i.e. ) and a surjection, and if the kernel of is closed then is Hausdorff.
If is a continuous, open surjection, then the kernel of is closed if and only if is Hausdorff.
Let with . Since is surjective, there exists such that and . Since , is in the complement of the kernel of , which is open by assumption. Since is a basis for the product topology on , there exists such that and is a subset of the complement of the kernel of . Let and . If then lies in the kernel of and in , which is a contradiction. Hence and are disjoint. Since is open, they are open. By construction, and . So is Hausdorff.
Assume now that is a continuous, open surjection. We just proved that since is an open surjection, if its kernel is closed then is closed. Conversely, if is continuous and is Hausdorff then its kernel is closed. Hence the equivalence stated. ∎
We offer an alternative proof of the first assertion of the previous proposition when is assumed to be open and bijective. We keep the notations used in that proposition.
Assume that is an open bijection and is closed. Since is bijective, it has a right inverse . If is open in then is open, so is continuous. Moreover, letting and since is surjective, we have:
which is therefore closed by assumption, as the preimage of the closed set by the continuous function . Hence is Hausdorff.
An open map may not be continuous. For instance, a nonconstant map from with the indiscrete topology to with the discrete topology is always open but never continuous.
Dealing with continuity can be complicated from the original definition. It is easier to introduce limits and the notion of continuity at a point. We first introduce a piece of vocabulary:
Let be a topological space. Let . The set of open neighborhoods of in is the set of all such that .
Let and be topological spaces. Assume is a Hausdorff space. Let . Let and . Let . We say that has limit at along when:
If then the notion is silly, since we could find one open set containing and so that in the above definition.
Let and be topological spaces. Assume is a Hausdorff space. Let . Let and . Let and . Assume that for all . Then has limit at along if and only if has limit at along .
Simply observe that the definition of limit involves only and . ∎
We thus can define without ambiguity the limit of a partially defined function at a point in the closure of its domain:
Let and be topological spaces. Assume is nonempty and . Then has limit at along if any extension of to has limit at along .
Let and be topological spaces. Assume is a Hausdorff space. Let . Let and . Let . Then if has limit at along then has limit at along .
First, since , we have , so implies that , hence the notion of limits are well-defined. Let . Since , there exists such that . Since we have as desired. ∎
Let and be topological spaces. Assume is a Hausdorff space. Let . Let and . Let . If has a limit at along then this limit is unique and denoted by .
Assume has limit and with at along . Since is Hausdorff, there exists open in such that and . Then by definition of limits, there exists open in so that , and and . So . This is absurd since and so . ∎
Let and be topological spaces. Assume is a Hausdorff space. Let . Let and . Let . If has a limit at along then .
Let . Let such that . Then by definition of limit, there exists such that and . By definition, . Since , we have so . So . ∎
In general, it is difficult to compute , so the result is used as follows: if then .
It is unpractical to check a statement for all open sets, as in general they are difficult to describe. The use of a basis makes things more amenable, and in fact it is the main role of basis.
Let and be topological spaces, with Hausdorff. Assume that we are given a basis for and a basis for . Let , and . Let . Let us denote by the subset of all such that , and we define similarly. Then has limit at along if and only if:
Assume 3.31. Let such that . By definition, there exists such that and . By 3.31 there exists , i.e. an open set in containing , such that . This shows that converges to at along .
Conversely, assume that converges to at along . Let . By definition, is an open set in containing so there exists with such that by definition of limit. There exists such that and as is a basis for . Hence, as desired. ∎
Let and be topological spaces, with T1. Let , . If and if has a limit at along then this limit is .
Assume is the limit of at along and . Then since is T1, there exists such that . By definition of limit, there exists with and . Now, so which is absurd. ∎
Let be a topological space. Let be a Hausdorff topological space. Let . Let . Then has limit at along if and only if has limit at along and along .
Note that implies that . The condition is necessary since and . Conversely, let . By definition of limits along and there exist such that and . Hence . Since , the proof is complete. ∎
Let , and be three topological spaces, where and are Hausdorff. Let and . Let , . Set . If is the limit of at along and is the limit of at along then is the limit of at along .
Note first that , as necessary to make sense of the statement of the theorem. Let with . There exists with such that . Now, there exists such that and . By construction, so . Hence, as desired. ∎
Beware of this theorem. Take , for and , and and . Then:
yet has no limit at along the set of nonzero reals.
Let and be topological spaces. Let . Then is continuous at if:
The notion of continuity at a point looks more familiar when the codomain is Hausdorff.
Let and be topological spaces. Assume is Hausdorff. Let . Then is continuous at if:
.
The main connection between continuity between spaces and continuity at a point is given in the following key theorem.
Let and be two topological spaces. The function is continuous at every if and only if is continuous on .
Assume first that is continuous at every . Let . Let . Then is continuous at so there exists such that and (since ). Consequently, . Hence:
and thus is open. So is continuous on .
Conversely, assume is continuous on . Let . Let such that . Then is open in by continuity of on , and it contains . Let . Then . This concludes the proof. ∎
We refer to the chapter Fundamental Examples for the topology on the one point compactification of .
Let be a set. A sequence in is a function from to .
Sequences are denoted as families. Namely, if is sequence in , then we write for for , and we usually identify with .
By abuse of language, we will also call sequences functions from a subset of , for some . The obvious re-indexing is left implicit.
The definition of limit for functions apply to sequences.
Let be a topological space. Let be a sequence in . Then we say that converges to when the limit of has limit at along in .
We learnt that if two functions agree on some set inside of a topological space , then their limits along agree (including whether they exist!). So the previous definition is understood as follows: choose any extension of a sequence to (i.e. pick some element in ). Then take its limit at in along . Then the existence and, if applicable, the value of this limit is independant of the choice the extension of our sequence. It is the limit of the sequence.
Let be a Hausdorff topological space. A sequence converges to if and only if:
Assume that has limit . Let . By definition of limits and of the topology , there exists a finite set in such that for all . Let be the successor of the greatest element in if is nonempty, or otherwise. Then for all we have as desired.
The converse is obvious. ∎
A subsequence of a sequence in a set is a sequence of the form for some strictly increasing function .
Let be strictly increasing. Then for all we have .
By definition, . Assume that for some we have . Then by assumption, so . The lemma holds by the theorem of induction. ∎
Let be a topological space. If a sequence in converges to then all subsequences of converge to .
Assume that converges to and let be a strictly increasing function. Let . By assumption, there exists such that for all we have . Therefore, since . Hence the subsequence converges to as well. ∎
Let be a topological space. A sequence in converges to if and only if every subsequence of has a subsequence which converges to .
The condition is necessary by the previous result. Let us show it is sufficient. Assume that does not converge to . Then there exists such that for all there exists such that . Set to be the smallest such that . Assume we have constructed for some such that for . Then let . Then , so it has a smallest element . Note that by construction, and . Hence we have constructed a subsequence entirely contained in . By construction, it has no subsequence which converges to . ∎
In this section, the topology on any metric space is meant as the metric topology.
Let be a metric space. A sequence in converges to if and only if:
Since open balls form a basis for the metric topology, the definition of limit for a sequence at infinity (along ) is equivalent to:
which is equivalent by definition to the statement in the proposition. ∎
The following result is often useful.
Let be a metric space. Let be a sequence in and . If there exists a sequence in and such that and for all , then .
Let be given. Since converges to in , there exists such that for all , . Hence, for we have , as desired. ∎
This theorem shows that a decent supply of sequences converging to in is useful. The following results help in these constructions.
Let be increasing and not bounded. Let . Then converges to .
Let be given. By assumption, there exists such that . Since is increasing, for all we have . Hence for all we have , as desired. ∎
The existence of unbounded functions is a consequence of the fact that is Archimedean. Namely, the injection is unbounded precisely because of the Archimedean principle (in fact, the two statements are equivalent). Therefore, converges to , as desired.
Let , and be three sequences in a linearly ordered set , and such that for all we have . If and have limit then so does .
∎
The above result is used in under many names: squeeze theorem, guardsmen theorem, pinching theorem, and more. It is often useful to prove limits.
Of course, is a field, and we will have to address the matter of the continuity of the operations. We shall do this later, however, as we do not need them at this moment.
Let be a metric space. Let . Then if and only if there exists a sequence in converging to .
If there exists a sequence in converging to then .
Conversely, assume . For , there exists . By construction, for all so converges to , which completes our proof. ∎
Limits of subsequences can be characterized in metric spaces as follows:
Let be a metric space. Let be a sequence in . Then is a limit for a subsequence of if and only if:
Assume is the limit of some subsequence of . Let . Then is the limit of the truncated sequence at , so since . Hence the condition is necessary.
Assume now . To ease notations, write for all . Since , the set is not empty. Let be its smallest element.
Assume now we have constructed for some such that for . Since , the set is not empty. Let be its smallest element.
One checks that as desired. ∎
Let , be metric spaces. Let be a map. Let , with , and . Then is has limit at along if and only if for any sequence in which converges to , we have .
Note that since and is metric, there is a sequence of elements of converging to . The theorem on composition of limits shows that the condition is necessary. Let us show that it is sufficient by contraposition. Assume does not converge to at . Hence, there exists such that for all , there exists such that yet . For each , pick such that and . Note that since , we can always choose such an . By the squeeze theorem, we conclude that . On the other hand, does not converge to . This proves our result. ∎
Let , be metric spaces. Let be a map. Let . Then is continuous at if and only if for any sequence which converges to in , we have .
This is a direct application of the result on limits in a metric space. ∎
This chapter introduces filters as a mean to analyze topological spaces. In a sense to be illustrated below, filters provide a generalization of sequences, albeit in a dual manner.
Let be a set. A filter base on is a subset of such that:
is not empty and,
If then there exists such that .
Let be a set. A filter on is a subset of such that:
,
,
If and with then ,
If then .
Hence a filter is a proper nonempty upset closed under finite intersections. Note that the empty set is never in a filter. Note that a filter is a filter basis.
Let be a filter basis on a set . Let:
Then is a filter on , and it is the smallest filter on containing . We call it the filter generated by .
By definition, so is not empty. Moreover, if and such that then, by construction, there exists such that and thus so . If then there exists such that for . Thus . Since is a filter basis, there exists such that so . Hence . Last, if then there must be such that , i.e. which is a contradiction. So . So is a filter.
Let be some filter containing . Let . Then by definition, there exists such that . Sicne and is a filter, we conclude . Hence as desired. ∎
Let , be sets and be a function. If is a filter basis on , then , defined as:
is a filter basis.
By definition, is not empty and does not contain the emptyset. Now, let . By definition, there exists such that and . Now, since is a filter basis, there exists such that . Hence , and by definition, . ∎
Let be constant, equal to , while have more than one element. Let . Then is a filter on . Now, . One checks easily that this is a filter basis, but not a filter (as it would contain which is assumed to contain a set of cardinal 2). Hence, images of filters may not be filters. We could define to be the filter generated by the filter basis , thus defining a map from filters to filters for each map from to .
Let be a topological space. Let . Let be a filter basis on . We say that converges to when:
Equivalently, a filter basis converges to if the set of open neighborhoods of is contained in the filter generated by .
The set is a filter basis which converges to . The elements of the filter generated by are called neighborhoods of .
The following theorem illustrates the connection between filters and sequences.
Let be a topological space. Let be a sequence in . For we define . Then is a filter basis. Moreover, converges to if and only if converges to .
Since for all , it is immediate that is a filter basis. Assume that converges to . Let such that . Then by definition, there exists such that for all we have . Hence . Thus, converges to (since was arbitrary).
Conversely, assume that converges to . Let such that . By definition, there exists such that . Hence for all we have . So converges to (since was arbitrary). ∎
Let be a sequence. Let be a subsequence. The filter generated by the filter basis is a subset of the filter generated by the filter basis . So filters “grow” as we take subsequences. From this observation, we note that by definition, it is now immediate to see that if a sequence converge to some then all of its subsequences do too. The following proposition generalizes this observation.
Let be a topological space. Let and be two filter bases on such that for all there exists such that (we say that is finer than ). If converges to , then converges to .
Let be a limit of . Let such that . There exists such that . By assumption, there exists such that . Hence converges to . ∎
Let be a topological space. If are two filters on and if converges to , then converges to .
Limits of filters may not be unique. We have:
Let be a topological space. The following are equivalent:
The space is Hausdorff,
Every convergent filter basis on has a unique limit.
Assume that is Hausdorff. Let be a filter basis on . Assume it converges to and in . If then, since is Hausdorff, there exists such that , and . By definition of convergence, there exist such that and . By definition of filter basis, there exists such that . This contradicts the fact , as a filter basis, does not contain the empty set. Hence limits are unique if they exist.
Assume now that all filter bases have unique limits in . In particular, for any , the filter basis of open neighborhoods of in , has for unique limit. Let . Suppose that for all , for all , we have . Let:
By assumption, is a nonempty set of elements in not containing the emptyset. Moreover, it is closed by finite intersection (as is). Hence, it is a filter basis. Yet, for any and any , we have so converges to and by definition. This is a contradiction. Hence, for all there exists such that , i.e. is Hausdorff. ∎
When is Hausdorff and is a filter basis converging to then we write: .
As a rule, filters can be used to replace sequences in general topology to obtain results limited to metric spaces. For instance:
Let be a topological space. Let . Let be a filter basis in which converges to in . Then .
Let . By definition of convergence, there exists such that . Now, by assumption, so . So since is arbitrary. ∎
Let be a topological space. Let . The closure of in is the set of all limits in of filters of .
All limits of filters of are in by the previous lemma. Assume . If then . It is then obvious that:
is a filter basis of which converges to . ∎
Let be a set, be a topological space, a filter basis on , and . Then we say that converges to along if converges to .
When is Hausdorff, we shall write: for this limit.
Let and be two topological spaces, and let . Let , . Then has limit at along if and only if for all filter bases of converging to , the filter basis converges to in .
Assume first that converges. Let be a filter basis of converging to . Note that we have shown such a filter basis exists. Now, let such that . Since converges to at along , there exists with such that . Since converges to , there exists such that . Since by assumption, so as desired. Since was arbitrary, converges to as desired.
Conversely, assume that converges to for all filter bases of converging to . Let such that . The filter basis converges to and is a subset of so by assumption, converges to . Thus, there exists such that . Hence, by definition, there exists such that . As is arbitrary in , we conclude that converges to at along . ∎
Let and be topological spaces, and . Then is continuous at if and only if the image by of all filter bases of converging to converge to .
Let be a set. An ultrafilter on is a filter such that for all , we have either or .
Let . Then is an ultrafilter.
Since for no set can we have and in a filter of , this definition suggests the following:
A filter on a set is an ultrafilter if and only if it is maximal among all filters on (for the inclusion).
Assume first that is an ultrafilter on . Let be a filter on such that . Let . Assume . Then , and thus is in , as filters are closed under finite intersections. This is a contradiction. Since is an ultrafilter, . As was arbitrary, .
Conversely, assume that is a maximal filter for inclusion. Assume that it is not an ultrafilter. Let such that . Without loss of generality, assume . Let:
Now, by construction, is not empty. Assume . Then there exists such that . Thus . As is a filter, this would imply which is a contradiction. So does not contain the empty set. By construction, if and then . Moreover, if then there exist such that for . Thus . Since we conclude that . Thus is a filter. By definition, since for all , we have . Since and , we have , yet . This contradicts the maximality of . So is an ultrafilter. ∎
Maximality lends itself to the application of Zorn’s lemma to prove the existence of ultrafilters containing any given filer.
Let be a set, and let be a filter basis on . Then there exists an ultrafilter containing .
Since every filter basis is contained in a filter, we will assume is a filter. Let:
The set is ordered by inclusion, and it is not empty (it contains . Note that is an upper set, and thus a maximal element of is a maximal element of the set of all filters on . Thus, it is enough to show that has a maximal element.
Let be a chain in , i.e. a linearly ordered subset of . Let . If then there exists such that and . Since is linearly ordered, we may assume . Thus , and is a filter, so . The set is not empty (filters never are), does not contain the empty set (filters never do), and is an upper set: if and for some , then there exists such that , and since is a filter, so . Therefore, is a filter. Trivially . Hence, admits an upper bound. Since is an arbitrary chain, Zorn’s lemma implies that admits a maximal element . Thus is a maximal filter, and hence an ultrafilter, containing . ∎
Due to this result, it is possible to rephrase the theorems in the application section of this chapter in terms of ultrafilters only.
A topoligical space is compact if, given any such that , there exists a finite subset of such that .
A subset of such that is called an open covering of .
Let be a topological space. Then the following are equivalent:
is compact,
For any set of closed subsets of whose intersection is empty, there exists a finite subset of whose intersection is empty.
The theorem follows by taking complements. ∎
Let be a compact space. Let be a decreasing sequence of nonempty closed subsets of (where the order on sets in inclusion). Then:
If then since is compact, we can find such that . Yet since the family is decreasing, which is assumed not empty. Our result follows by contraposition. ∎
Let be a topological space. Let be a set of closed subsets of such that given any finite , we have (we say that has the finite intersection property). Then if and only if is compact.
Assume first that is compact and consider a set of closed subsets of . Assume . Since is compact, there exists a finite set such that . This proves the necessity of our theorem by contraposition.
Conversely, assume that for all sets of closed subsets of with the finite intersection property, . Let be a set of closed subsets of such that . Then by our assumption, there exists at least one finite subset such that . ∎
The space is compact if and only if is finite.
It is striaghtforward that if in for all then is an open covering of which has no finite subcovering. So is not compact.
Let be a topological space. Let . Then is a compact subspace of if is a compact space (where is the trace topology on ).
We will say that is compact in when is a compact subspace of .
Let be a topological space. Let . Then is compact if and only if for all such that , there exists and such that:
Assume that is compact in . Let such that . Then . By definition, for all . Hence by compactness of , there exists such that . Hence .
Assume now that any open covering of in admits a finite covering. Let such that . By definition of , for each there exists such that . Thus , so there exists such that . Hence . ∎
Let be a topological space. Let . Then is compact if and only if for any family of closed sets in such that:
there exists a finite subset such that:
Take complements in previous theorem. ∎
Let be a Hausdorff topological space. Then if is compact, then is closed.
Let and . Since is Hausdorff, there exists such that , with and . Let . By construction:
and thus, since is compact, there exists and such that:
Let:
As a finite intersection of open sets, is open. Moreover:
Hence, . Therefore:
and thus is open, so is closed as desired. ∎
Let be a compact space. Then if is closed, then it is compact.
Let such that and let be closed. Then so, if , then . Since is compact, there exists a finite set such that . Now, if , one checks readily that and by construction, with finite. ∎
Let be a compact Hausdorff space. Then is compact if and only if is closed.
Let be a compact space, be a topological space, and be a continuous function. Then is a compact subspace of .
Let such that . Since is continuous, for all , we have . Let . Then and by construction, . Since is compact, there exists and such that . Hence as desired. ∎
Let and be topological spaces. If is continuous, and is compact, then is compact.
Let be a compact space, and be a Hausdorff topological space. Let be a continuous bijection. Then is a homeomorphism.
It suffices to show that is continuous. Let be a closed set. Then is compact since is compact. So is compact in . Since is Hausdorff, is closed. Since is a bijection, . So is continuous. ∎
Let be a metric space. Then the following are equivalent:
is compact in its metric topology.
Bolzanno-Weierstrass Every sequence in admits a convergent subsequence.
First, assume that is compact. Let be a sequence in . Let for all . The sequence is a decreasing sequence of nonempty closed subsets of , which is compact, so . By Theorem (LABEL:AccumulationSeqThm), this implies that admits a convergent subsequence.
Second, assume that the Bolzanno-Weierstass axiom holds for . We first prove a few lemmas.
Let be an open covering of . There exists such that for all , there exists such that the open ball of center and radius is contained in .
Assume that for all , there exists such that for all , we have . Let . There exists such that for all . The sequence admits a convergent subsequence by the Bolzanno-Weierstrass axiom. Let be its limit. Since covers , there exists such that . Since is open, there exists such that . Since converges to , there exists such that for all , we have . Let such that and . Let . Then:
since . Hence, so . This is a contradiction. ∎
For any there exists and such that:
.
Assume that there exists such is not the union of finitely many open balls of radius . Let : since , there exists such that .
Assume we have constructed such that for . Since , there exists , i.e. .
By induction, we have constructued a sequence which admits no convergent subseqence. This contradicts the Bolzanno-Weierstrass axiom. ∎
We can now conclude our proof. Assume is an open covering of . There exists such that any open ball of radius or less is contained in some . Now, there exists such that . Let such that . Then . So is compact, as desired. ∎
A compact metric space is bounbed.
The precompact lemma shows that a compact metric space can be covered with finitely many balls of radius 1. ∎
The definition of compacity by means of open covering is referred to the Heine-Borel axioms of compactity. Thus we have shown that in metric spaces, Heine-Borel and Bolzanno-Weierstrass are equivalent.
We provide two proofs of the following result, both based on your real analysis class.
A subset of is compact if and only if it is closed and bounded.
Since is Hausdorff, compact subsets are closed. Since is metric, compact subsets are bounded. ∎
Assume that is closed and bounded. Let be a sequence in . By the monotone subsequence theorem, there exists a monotone subsequence of . Since is bounded, this subsequence must converge. Hence satisfies Bolzanno-Weierstrass axiom in a metric space and is thus compact. ∎
Let be real numbers. Let be a sequence in . Set .
Assume that for some we have constructed , and such that and . Let . If is infinite, then let be its smallest element. We also let , . Otherwise, we set to be the smallest element of the (necessarily infinite) set , and we let , . One checks that by induction, we have constructed an subsequence and monotone sequences and such that:
is increasing and bounded above by so it converges.
is decreasing and bounded above by so it converges.
We have so .
Thus by the squeeze theorem, .
Hence again we have proven the Bolzanno-Weierstrass axiom for all closed intervals. Now, if is closed and bounded, then it is a closed subset of some closed interval, i.e. of a compact space. Hence it is compact, as needed. ∎
The second proof is longer but can be applied as is to prove that compact subsets of are closed bounded subsets of . Indeed, rather than cut an interval in two, one can repeat the argument, cutting a square in four, a cube in eight, and so forth. This process works well as long as we divide an hypercube into finitely many hypercubes at each stage. This proofs fails for infinite dimensional normed vector spaces, and in fact so does the result: one can show that locally compact normed vector spaces are all finite dimensional.
Let and be metric spaces. A function is uniformly continuous when:
A uniformly continuous function is continuous.
Obvious. ∎
A -Lipshitz function is uniformly continuous: take if (i.e. nonconstant functions), and arbitrary for constants.
The function is not uniformly continuous in . Indeed basic precalculus shows that if and only if .
The following shows that continuity can be strengthened to uniform continuity when working on a compact metric space.
Let be a compact metric space. Let be a metric space. Let be given. Then is continuous if and only if is uniformly continuous on .
If is uniformly continuous then it is continuous. Let us prove the converse. Assume is continuous on . Let . Then for all there exists such that if then . Now, since , and is compact, there exists such that . Let . Let with . There exists such that . Then
Thus:
as desired.
∎
If is uniformly continuous then it is continuous. Let us prove the converse. Assume is continuous but not uniformly continuous. There exists such that for all there exists such that and . For let such that and . Since is compact, there exists a convergent subsequence . Then there exists a convergent subsequence of . Let and . Now . On the other hand, using the continuity of , we have which is a contradiction. ∎
The notion of completeness of metric spaces is not topological: there exist two homeomorphic metric spaces, one complete and one not. A possible choice of morphisms for the category of metric spaces are the uniformly continuous functions, in which case isomorphic metric spaces are either all complete or all not complete.
Now, if and are compact metric spaces, and if they are homeomorphic, then is complete if and only if is complete. This is a very remarkable fact: a purely topological notion (compactness) imposes a metric notion (completeness). This is seen by observing that a Cauchy sequence has at most one accumulation point, and we have seen that every sequence of a compact space has at least one. With the Heine theorem, we see a different reason: an homeomorphism between compact metric spaces is always bi-uniformly continuous, and hence preserve Cauchy sequences.
Let be a topological space. is compact if and only if every ultrafilter in converge.
Assume that is compact. Let be an ultrafilter in . Assume it does not converge. Then for all there exists with such that . As is an ultrafilter, . Now, , and as is compact, there exists a finite subset of such that . Now, which is a contradiction. So converges.
Assume now that all ultrafilters in converge in . Assume is not compact. Let such that , with infinite, and such that no finite subset of covers . Let:
Since no finite subset of covers , the emptyset is not in . By construction, is closed by finite intersections, and thus it is a filter basis. Assume it converges to . Since covers , there exists such that . By definition of convergence, there exists a finite set such that , i.e.:
with finite subset of . This is a contradiction. Hence does not converge. Now, is contained in some ultrafilter , which therefore does not converge at all. ∎
Let be a family of compact topological spaces. Let be the product topology of . Then is compact.
Let be an ultrafilter in . Let . Let be the canonical surjection. Let . Now, let and such that . Since is a filter, and since , we conclude and . Since is a surjection, by definition. It is also immediate that is a nonempty subset of to which the empty set does not belong. Now, the same reasonning shows that if then . Hence, is a filter in . Last, let . Then either or its complement are in , since is an ultrafilter. Since as is a surjection, either or its complemenent belongs to . Thus is an ultrafilter on .
Since is compact, converges to some .
Let us show that converges to in the product topology . Let such that . Let and finite such that for , and , . Write for all and note that , with for all . Also note that . By definition of convergence and , we have , so . Since is closed under finite intersections, we conclude , hence . So converges to as desired. ∎