Notes
Introduction
This chapter generalises the concepts of connectedness, which is depended on by a basic theorem regarding continuous functions called the 'intermediate value theorem, as well as compactness, which sees application in establishing important results such as extreme value theorem and uniform continuity, from real numbers which is central in mathematical analysis to arbitrary topological spaces.
Connectedness
Connected Spaces
Definition. Let be a topological space. A separation of is pair of disjoint nonempty open subsets of whose union is . The space is is said to be connected if there does not exist a separation of .
-
If is connected, then so is any space homeomorphic to , as is a topological property.
-
Another way of phrasing the definition of connectedness is the following:
A space is connected if and only if the only subsets of that are both open and closed in are the empty set and itself.
This is because if is a nonempty proper subset of that is both open and closed in , then the sets and form a separation of , since they are open, disjoint, nonempty, and have as their union. Conversely, if and form a separation of , then is nonempty and different from , and it is both open and closed in .
[Lemma 23.1] (Subspace connectedness). If is a subspace of , a separation of is a pair of disjoint nonempty sets and whose union is , neither of which contains a limit point of the other. The space is connected if there exists no separation of .
Constructing new connected spaces from old.
[Lemma 23.2] If the sets and form a separation of , and if is a connected subspace of , then lies entirely within either or .
[Theorem 23.3] The union of a collection of connected subspaces of that have a point in common (i.e. the intersection of the collection is nonempty) is connected.
[Theorem 23.4] Let be a connected subspace of . If , then is also connected.
Equivalently: if is formed by adjoining to the connected subspace some or all of tis limit points, then is connected.
[Theorem 23.5] (Continuous functions preserve connectedness). The image of a connected space under a continuous map is connected.
Let be a continuous map where is connected. We want to prove that is connected.
Since the map obtained from by restricting its range to the space is also continuous (see Theorem 18.2 in Chapter 2), it suffices to consider the case of a continuous surjective map .
Assume that there exists two disjoint nonempty open subsets and of such that , then and are disjoint sets whose union is . Since is continuous, and are open in ; they are also nonempty since is surjective. Thus, they form a separation of , contradicting the connectedness of .
[Theorem 23.6] A finite Cartesian product of connected spaces is connected.
Does Theorem 23.6 extend to arbitrary products of connected space? The answer is: it depends on the topology used for the product.
In fact, an arbitrary product of connected space is connected in the product topology. A proof of this result is a solution to Exercise 23.10.
Connected Subspaces of the Real Line
Definition. A simply ordered (a.k.a. strictly totally ordered) set having more than one element is called a linear continuum if the following hold:
- has the least upper bound property.
- If , there exists such that (i.e. the order on is dense).
[Theorem 24.1] If is a linear continuum in the order topology, then is connected, and so are intervals and rays in .
Since is a linear continuum in order topology, we immediately see that:
[Corollary 24.2] The real line is connected and so are intervals and rays in .
One may also ask: does the converse of Theorem 24.1 hold? The answer is in fact yes, and a proof of this is a solution to Exercise 24.4.
[Theorem 24.3] (Intermediate Value Theorem). Let be a continuous map, where is a connected space and is an ordered set in the order topology. If and are two points of and if is a point of lying between and , then there exists a point of such that .
Definition. (Path-connectedness). Given points and of the space , a path in from to is a continuous map of some closed interval in the real line into , such that and . A space is said to be path connected if every pair of points of can be joined by a path in .
The converse of the definition above does not hold; a connected space need not be path connected.
[Example 6]. The orderred square (recall that it is defined as , where , endowed with the dictionary order) is connected but not path connected.
Since it is a linear continuum (as a subset of ), it is connected. We show that is not path connected as follows:
Let and (note that they are respectively the smallest and largest element of ). Assume for the sake of contradiction that there is a path joining and . It follows that the image set must contain every point of by the intermediate value theorem (recall that a path is a continuous map).
Therefore, for each , the set
is a nonempty subset of ; by continuity, it is open in . Choose, for each , a rational number belonging to . Since the sets are disjoint, the map is an injective mapping of into . This contradicts the fact that the inverval is uncountable.
The uncountability of comes from the fact that if there were a bijective mapping from the open interval (note that this is just but 0 and 1 are excluded) to the natural numbers (which is the definition of being countable) so that each element of it can be indexed as where , then one can come up with an element in given by where , which is not equal to any of the 's, contradicting the existence of said bijective mapping.
There is also a topological proof for the uncountability of , which is discussed in Section 27 (Compact Subspaces of the Real Line), specifically Theorem 27.7.
Components and Local Connectedness
Breaking up into smaller pieces.
The notion of local connectedness refers to a natural process of breaking up an arbitrary (topological) space into connected components.
Definition. ([Connected] components). Given , define an equivalence relation on by setting if there is a connected subspace of containing both and . The equivalence classes are then called the components (or the 'connected components') of .
[Theorem 25.1] The components of are connected disjoint subspaces of whose union is , such that each nonempty connected subspace of intersects only one of them.
This comes from the fact that the components form equivalence classes in , which induces a partition of .
Definition. (Path components). We define another equivalence relation on the space by defining if there is a path in from to . The equivalence classes are called the path components of .
[Theorem 25.2] The path components of are path-connected disjoint subspaces of whose union is , such that each nonempty path-connected subspace of intersects only one of them.
Definition. (Local connectedness). A space is said to be locally connected at if for every neighbourhood of , there is a connected neighbourhood of contained in . If is a locally connected at each of its points, it is said simply to be locally connected.
Similarly, a space is said to be locally path connected at if for every neighbourhood of , there is a path-connected neighbourhood of contained in . If is locally path connected at each of its points, then it is said to be locally path connected.
[Theorem 25.3] A space is locally connected if and only if for every open set of , each component of is open in .
[Theorem 25.4] A space is locally path connected if and only if for every open set of , each path component of is open in .
[Theorem 25.5] (Relation between path components and components). If is a topological space, each path component of lies in a component of . If is locally path connected, then the components and the path components of are the same.
Compactness
According to the Wikipedia article on compact spaces, the notion of compactness seeks to generalise the notion of closed and bounded subsets of Euclidean space.
Compactness is also key in extending local properties to global properties (i.e. properties that apply to the whole space).
Definition. A collection of subsets of a space is said to cover , or to be a covering of , if the union of the elements of is equal to . It is called an open covering of if its elements are open subsets of .
Definition. A space is said to be compact if every open covering of contains a finite subcover (i.e. a finite subcollection that also covers ).
[Lemma 26.1] Let be a subspace of , then is compact if and only if every covering of by sets open in contains a finite subcollection covering .
[Theorem 26.2] Every closed subspace of a compact space is compact.
[Theorem 26.3] Every compact subspace of a Hausdorff space is closed.
[Lemma 26.4] If is a compact subspace of the Hausdorff space and is not in , then there exists disjoint open sets and of containing and , respectively.
[Theorem 26.5] The image of a compact space under a continuous map is compact.
Let be continuous and be compact. Let be a covering of the set by sets open in . It follows that
is a collection of sets covering ; since is continuous, these sets are open. Since is compact, there are finitely many of them, say,
that cover , so the sets cover .
[Theorem 26.6] Let be a bijective continuous function. If is compact and is Hausdorff, then is a homeomorphism.
[Theorem 26.7] The product of finitely many compact spaces is compact.
A natural question arises from Theorem 26.7: Is the product of infinitely many compact spaces compact?. The answer is in fact 'yes.' It is called the Tychonoff's theorem, which is further discussed in Chapter 5 of this book.
[Lemma 26.8] (The tube lemma). Consider the product space , where is compact. If is an open set of containing the slice of , then contains some tube about , where is a neighbourhood of in .
Definition. A tube about is the set , where is a neighbourhood of in such that (as defined above) contains the entire set .
Definition. A collection of subsets of is said to have the finite intersection property if for every finite subcollection of , the intersection is nonempty.
[Theorem 26.9] Let be a topological space. Then is compact if and only if for every collection of closed sets in having the finite intersection property, the intersection of all the elements of is nonempty.
A special case of Theorem 26.9 is the Cantor's intersection theorem: given a nested sequence of closed sets in a compact space , if each of the sets is nonempty, then the collection automatically has the finite intersection property, so the intersection is nonempty.
Compact Subspaces of the Real Line
[Theorem 27.1] Let be a simply ordered set having the least upper bound property. In the order topology, each closed interval in is compact.
[Corollary 27.2] Every closed interval in is compact.
[Theorem 27.3] A subspace of is compact if and only if it is closed and is bounded in the Euclidean metric or the square metric .
Theorem 27.3 is also known as the Heine-Borel theorem in real analysis.
[Theorem 27.4] (Extreme value theorem, generalised). Let be continuous, where is an ordered set in the order topology. If is compact, then there exists points and in such that for every .
Definition. Let be a metric space; let be a nonempty subset of . For each , we define the distance from to by the equation
For fixed , the function is a continuous function of . Indeed, given , we have the inequalities
for each , which implies that
The same inequality holds with and interchanged. The continuity of the function follows from the continuity of and Theorem 21.1 (by setting ).
[Lemma 27.5] (The Lebesgue number lemma). Let be an open covering of the metric space . If is compact, there is a such that for each subset of having diameter less than , there exists an element of containing it.
The number is called a Lebesgue number for the covering .
Definition. A function from the metric space to the metric space is said to be uniformly continuous if given , there is a such that for every pair of points of , we have .
[Theorem 27.6] (Uniform continuity theorem). Let be a continuous map of the compact metric space to the metric space , then is uniformly continuous.
Theorem 27.6 is closely related to the continuous extension theorem in real analysis.
Definition. If is a space, a point of is called an isolated point of if the one-point set is open in .
[Theorem 27.7] Let be a nonempty compact Hausdorff space. If has no isolated points, then is uncountable.
Step 1. We show first that for any nonempty open set of and any point in , there exists a nonempty open set contained in such that .
Choose a point such that ; this is possible if because is not an isolated point of and it is possible if since . Now choose disjoint open sets and about and , respectively (i.e. and ; this is possible since is Hausdorff), then the set is the desired open set; it is contained in , it is nonempty because it contains , and its closure does not contain .
Step 2. We show that given , the function is not surjective, which implies that is uncountable.
Let . Apply Step 1 to the nonempty open set to choose a nonempty open set such that . In general, given open and nonempty, choose to be a nonempty open set such that and .
Consider the nested sequence of nonempty closed sets of . Since is compact, by Theorem 26.9, there is a point . Now cannot equal for any , since belongs to and does not.
[Corollary 27.8] (from Theorem 27.7) Every closed interval in is uncountable.
Every closed interval in is nonempty, compact and Hausdorff. It also has no isolated points, since if we consider as a subspace of , the intersection of with every nonempty open set of must contain more than one point.
Limit Point Compactness
Definition. A space is said to be limit point compact if every infinite subset of has a limit point.
[Theorem 28.1] Compactness implies limit point compactness, but the converse is not true.
Definition. (Subsequence and sequential compactness). Let be a topological space. If is a sequence of points of , and if is an increasing sequence of positive integers, then the sequence defined by setting is called a subsequence of the sequence . The space is said to be sequentially compact if every sequence of poitns of has a convergent subsequence.
The notion of sequential compactness is closely related to the Bolzano-Weierstrass theorem.
[Theorem 28.2] Let be a metrizable space, then is compact iff is limit point compact iff is sequentially compact.
Local Compactness
Definition. A space is said to be locally compact at if there is some compact subspace of that contains a neighbourhood of . If is locally compact at each of its points, is said simply to be locally compact.
Note that a compact space is automatically locally compact.
[Theorem 29.1] Let be a space, then is locally compact Hausdorff if and only if there exists a space satisfying the following conditions:
- is a subspace of .
- The set consists of a single point.
- is a compact Hausdorff space.
In fact, such can be constructed by adjoining a point not in to , and defining its topology to consist of all sets that are open in and all sets of the form where is a compact subspace of .
If and are two spaces satisfying the conditions above, then there is a homeomorphism of with which equals the identity map on . (Another way of phrasing this is that the existence of such space is unique up to a homeomorphism that maps identically.)
Definition. If is a compact Hausdorff space and is a proper subspace of whose closure equals , then is said to be a compactification of . If is a single point, then is called the one-point compactification of (a.k.a. Alexandroff compactification).
From Theorem 29.1, we see that has a one-point compactification if and only if is a locally compact Hausdorff space that is not itself compact, since in this case the point of is a limit point of , so that .
If itself happens to be compact, then the space obtained from Theorem 29.1 is obtained from by adjoining a single isolated point.
[Theorem 29.2] (Equivalent formulation of local compactness for Hausdorff spaces) Let be a Hausdorff space, then is locally compact if and only if given in , and given a neighbourhood of , there is a neighbourhood of such that is compact and .
[Corollary 29.3] Let be locally compact Hausdorff and be a subspace of . If is closed in or open in , then is locally compact.
[Corollary 29.4] A space is homeomorphic to an open subspace of a compact Hausdorff space if and only if is locally compact Hausdorff.
Suppose that is locally compact Hausdorff. If itself is compact, then itself is an open subspace of a compact Hausdorff space that is homeomorphic to , as desired. If is not compact, then has a one-point compactification given by Theorem 29.1, such that is an open subspace of , which is as desired because again, is indeed homeomorphic to itself.
Suppose that is homeomorphic to an open subspace of a compact Hausdorff space. If is compact, then must be locally compact and the assumed homeomorphism implies that is also Hausdorff due to Exercise 17.12, so we are done. If is not compact, then the one-point compactification given by Theorem 29.1 has , which is Hausdorff, as its open subspace. It is also locally compact by Corollary 29.3.