Skip to main content

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 XX be a topological space. A separation of XX is pair U,VU, V of disjoint nonempty open subsets of XX whose union is XX. The space is XX is said to be connected if there does not exist a separation of XX.

info
  • If XX is connected, then so is any space homeomorphic to XX, as XX is a topological property.

  • Another way of phrasing the definition of connectedness is the following:

    A space XX is connected if and only if the only subsets of XX that are both open and closed in XX are the empty set and XX itself.

    This is because if AA is a nonempty proper subset of XX that is both open and closed in XX, then the sets U=AU=A and V=X\AV=X \backslash A form a separation of XX, since they are open, disjoint, nonempty, and have XX as their union. Conversely, if UU and VV form a separation of XX, then UU is nonempty and different from XX, and it is both open and closed in XX.

[Lemma 23.1] (Subspace connectedness). If YY is a subspace of XX, a separation of YY is a pair of disjoint nonempty sets AA and BB whose union is YY, neither of which contains a limit point of the other. The space YY is connected if there exists no separation of YY.

Constructing new connected spaces from old.

[Lemma 23.2] If the sets CC and DD form a separation of XX, and if YY is a connected subspace of XX, then YY lies entirely within either CC or DD.

[Theorem 23.3] The union of a collection of connected subspaces of XX that have a point in common (i.e. the intersection of the collection is nonempty) is connected.

[Theorem 23.4] Let AA be a connected subspace of XX. If ABAA \subseteq B \subseteq \overline{A}, then BB is also connected.

Equivalently: if BB is formed by adjoining to the connected subspace AA some or all of tis limit points, then BB is connected.

[Theorem 23.5] (Continuous functions preserve connectedness). The image of a connected space under a continuous map is connected.

Let f:XYf : X \to Y be a continuous map where XX is connected. We want to prove that f(X)f(X) is connected.

Since the map obtained from ff by restricting its range to the space f(X)f(X) is also continuous (see Theorem 18.2 in Chapter 2), it suffices to consider the case of a continuous surjective map g:Xf(X)g : X \to f(X).

Assume that there exists two disjoint nonempty open subsets AA and BB of YY such that f(X)=ABf(X)=A \cup B, then g1(A)g^{-1}(A) and g1(B)g^{-1}(B) are disjoint sets whose union is XX. Since gg is continuous, g1(A)g^{-1}(A) and g1(B)g^{-1}(B) are open in XX; they are also nonempty since gg is surjective. Thus, they form a separation of XX, contradicting the connectedness of XX.

[Theorem 23.6] A finite Cartesian product of connected spaces is connected.

info

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 LL having more than one element is called a linear continuum if the following hold:

  1. LL has the least upper bound property.
  2. If x<yx < y, there exists zz such that x<z<yx < z < y (i.e. the order on LL is dense).

[Theorem 24.1] If LL is a linear continuum in the order topology, then LL is connected, and so are intervals and rays in LL.

info

Since R\mathbb{R} is a linear continuum in order topology, we immediately see that:

[Corollary 24.2] The real line R\mathbb{R} is connected and so are intervals and rays in R\mathbb{R}.

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 f:XYf : X \to Y be a continuous map, where XX is a connected space and YY is an ordered set in the order topology. If aa and bb are two points of XX and if rr is a point of YY lying between f(a)f(a) and f(b)f(b), then there exists a point cc of XX such that f(c)=rf(c)=r.

Definition. (Path-connectedness). Given points xx and yy of the space XX, a path in XX from xx to yy is a continuous map f:[a,b]Xf : [a,b] \to X of some closed interval in the real line into XX, such that f(a)=xf(a)=x and f(b)=yf(b)=y. A space XX is said to be path connected if every pair of points of XX can be joined by a path in XX.

warning

The converse of the definition above does not hold; a connected space need not be path connected.

[Example 6]. The orderred square Io2I_o^2 (recall that it is defined as I×IR2I \times I \subset \mathbb{R}^2, where I:=[0,1]I := [0,1], endowed with the dictionary order) is connected but not path connected.

Since it is a linear continuum (as a subset of R2\mathbb{R}^2), it is connected. We show that Io2I_o^2 is not path connected as follows:

Let p=(0,0)p = (0,0) and q=(1,1)q = (1,1) (note that they are respectively the smallest and largest element of Io2I_o^2). Assume for the sake of contradiction that there is a path f:[a,b]Io2f : [a,b] \to I_o^2 joining pp and qq. It follows that the image set f([a,b])f([a,b]) must contain every point (x,y)(x,y) of Io2I_o^2 by the intermediate value theorem (recall that a path is a continuous map).

Therefore, for each xIx \in I, the set

Ux=f1({x}×(0,1))U_x = f^{-1}(\{x\} \times (0,1))

is a nonempty subset of [a,b][a,b]; by continuity, it is open in [a,b][a,b]. Choose, for each xIx \in I, a rational number qxq_x belonging to UxU_x. Since the sets UxU_x are disjoint, the map xqxx \mapsto q_x is an injective mapping of II into Q\mathbb{Q}. This contradicts the fact that the inverval II is uncountable.

note

The uncountability of II comes from the fact that if there were a bijective mapping from the open interval (0,1)(0,1) (note that this is just II 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 xi=0.di1di2di3x_i=0.d_{i1}d_{i2}d_{i3}\cdots where iNi \in \mathbb{N}, then one can come up with an element in II given by x=0.a1a2a3x=0.a_1 a_2 a_3 \cdots where ai=dii+1(mod10)a_i=d_{ii}+1 \pmod{10}, which is not equal to any of the xix_i's, contradicting the existence of said bijective mapping.

There is also a topological proof for the uncountability of II, 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 XX, define an equivalence relation on XX by setting xyx \sim y if there is a connected subspace of XX containing both xx and yy. The equivalence classes are then called the components (or the 'connected components') of XX.

[Theorem 25.1] The components of XX are connected disjoint subspaces of XX whose union is XX, such that each nonempty connected subspace of XX intersects only one of them.

This comes from the fact that the components form equivalence classes in XX, which induces a partition of XX.

Definition. (Path components). We define another equivalence relation on the space XX by defining xyx \sim y if there is a path in XX from xx to yy. The equivalence classes are called the path components of XX.

[Theorem 25.2] The path components of XX are path-connected disjoint subspaces of XX whose union is XX, such that each nonempty path-connected subspace of XX intersects only one of them.

Definition. (Local connectedness). A space XX is said to be locally connected at xx if for every neighbourhood UU of xx, there is a connected neighbourhood VV of xx contained in UU. If XX is a locally connected at each of its points, it is said simply to be locally connected.

Similarly, a space XX is said to be locally path connected at xx if for every neighbourhood UU of xx, there is a path-connected neighbourhood VV of xx contained in UU. If XX is locally path connected at each of its points, then it is said to be locally path connected.

[Theorem 25.3] A space XX is locally connected if and only if for every open set UU of XX, each component of UU is open in XX.

[Theorem 25.4] A space XX is locally path connected if and only if for every open set UU of XX, each path component of UU is open in XX.

[Theorem 25.5] (Relation between path components and components). If XX is a topological space, each path component of XX lies in a component of XX. If XX is locally path connected, then the components and the path components of XX 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 A\mathcal{A} of subsets of a space XX is said to cover XX, or to be a covering of XX, if the union of the elements of AA is equal to XX. It is called an open covering of XX if its elements are open subsets of XX.

Definition. A space XX is said to be compact if every open covering A\mathcal{A} of XX contains a finite subcover (i.e. a finite subcollection that also covers XX).

[Lemma 26.1] Let YY be a subspace of XX, then YY is compact if and only if every covering of YY by sets open in XX contains a finite subcollection covering YY.

[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 YY is a compact subspace of the Hausdorff space XX and x0x_0 is not in YY, then there exists disjoint open sets UU and VV of XX containing x0x_0 and YY, respectively.

[Theorem 26.5] The image of a compact space under a continuous map is compact.

Let f:XYf : X \to Y be continuous and XX be compact. Let A\mathcal{A} be a covering of the set f(X)f(X) by sets open in YY. It follows that

{f1(A):AA}\{f^{-1}(A) : A \in \mathcal{A}\}

is a collection of sets covering XX; since ff is continuous, these sets are open. Since XX is compact, there are finitely many of them, say,

f1(A1),,f1(An),f^{-1}(A_1),\ldots, f^{-1}(A_n),

that cover XX, so the sets A1,,AnA_1,\ldots,A_n cover f(X)f(X).

[Theorem 26.6] Let f:XYf : X \to Y be a bijective continuous function. If XX is compact and YY is Hausdorff, then ff is a homeomorphism.

[Theorem 26.7] The product of finitely many compact spaces is compact.

info

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 X×YX \times Y, where YY is compact. If NN is an open set of X×YX \times Y containing the slice x0×Yx_0 \times Y of X×YX \times Y, then NN contains some tube W×YW \times Y about x0×Yx_0 \times Y, where WW is a neighbourhood of x0x_0 in XX.

info

Definition. A tube about x0×Yx_0 \times Y is the set W×YW \times Y, where WW is a neighbourhood WW of x0x_0 in XX such that NN (as defined above) contains the entire set W×YW \times Y.

Definition. A collection C\mathcal{C} of subsets of XX is said to have the finite intersection property if for every finite subcollection {C1,,Cn}\{C_1,\ldots,C_n\} of C\mathcal{C}, the intersection C1CnC_1 \cap \cdots \cap C_n is nonempty.

[Theorem 26.9] Let XX be a topological space. Then XX is compact if and only if for every collection C\mathcal{C} of closed sets in XX having the finite intersection property, the intersection CCC\bigcap_{C \in \mathcal{C}}C of all the elements of C\mathcal{C} is nonempty.

info

A special case of Theorem 26.9 is the Cantor's intersection theorem: given a nested sequence C1C2CnCn+1C_1 \supseteq C_2 \supseteq \cdots \supseteq C_n \supseteq C_{n+1} \supseteq \cdots of closed sets in a compact space XX, if each of the sets CnC_n is nonempty, then the collection C={Cn}nZ+\mathcal{C} = \{C_n\}_{n \in \mathbb{Z}_+} automatically has the finite intersection property, so the intersection nZ+Cn\bigcap_{n \in \mathbb{Z}_+} C_n is nonempty.

Compact Subspaces of the Real Line

[Theorem 27.1] Let XX be a simply ordered set having the least upper bound property. In the order topology, each closed interval in XX is compact.

[Corollary 27.2] Every closed interval in R\mathbb{R} is compact.

[Theorem 27.3] A subspace AA of Rn\mathbb{R}^n is compact if and only if it is closed and is bounded in the Euclidean metric dd or the square metric ρ\rho.

note

Theorem 27.3 is also known as the Heine-Borel theorem in real analysis.

[Theorem 27.4] (Extreme value theorem, generalised). Let f:XYf : X \to Y be continuous, where YY is an ordered set in the order topology. If XX is compact, then there exists points cc and dd in XX such that f(c)f(x)f(d)f(c) \leqslant f(x) \leqslant f(d) for every xXx \in X.

Definition. Let (X,d)(X,d) be a metric space; let AA be a nonempty subset of XX. For each xXx \in X, we define the distance from xx to AA by the equation

d(x,A)=inf{d(x,a):aA}d(x,A)=\inf\{d(x,a) : a \in A\}
info

For fixed AA, the function d(x,A)d(x,A) is a continuous function of xx. Indeed, given x,yXx, y \in X, we have the inequalities

d(x,A)d(x,a)d(x,y)+d(y,a)d(x,A) \leqslant d(x,a) \leqslant d(x,y)+d(y,a)

for each aAa \in A , which implies that

d(x,A)d(x,y)infaAd(y,a)=d(y,A)    d(x,A)d(y,A)d(x,y).d(x,A)-d(x,y) \leqslant \inf_{a \in A}d(y,a)=d(y,A) \implies d(x,A)-d(y,A) \leqslant d(x,y).

The same inequality holds with xx and yy interchanged. The continuity of the function d(x,A)d(x,A) follows from the continuity of d(x,y)d(x,y) and Theorem 21.1 (by setting δ=ε\delta=\varepsilon).

[Lemma 27.5] (The Lebesgue number lemma). Let A\mathcal{A} be an open covering of the metric space (X,d)(X,d). If XX is compact, there is a δ>0\delta>0 such that for each subset of XX having diameter less than δ\delta, there exists an element of A\mathcal{A} containing it.

The number δ\delta is called a Lebesgue number for the covering A\mathcal{A}.

Definition. A function ff from the metric space (X,dX)(X,d_X) to the metric space (Y,dY)(Y,d_Y) is said to be uniformly continuous if given ε>0\varepsilon>0, there is a δ>0\delta>0 such that for every pair of points x0,x1x_0,x_1 of XX, we have dX(x0,x1)<δ    dY(f(x0),f(x1))<εd_X(x_0,x_1)<\delta \implies d_Y(f(x_0),f(x_1)) < \varepsilon.

[Theorem 27.6] (Uniform continuity theorem). Let f:XYf : X \to Y be a continuous map of the compact metric space (X,dX)(X,d_X) to the metric space (Y,dY) (Y,d_Y), then ff is uniformly continuous.

note

Theorem 27.6 is closely related to the continuous extension theorem in real analysis.

Definition. If XX is a space, a point xx of XX is called an isolated point of XX if the one-point set {x}\{x\} is open in XX.

[Theorem 27.7] Let XX be a nonempty compact Hausdorff space. If XX has no isolated points, then XX is uncountable.

Step 1. We show first that for any nonempty open set UU of XX and any point xx in XX, there exists a nonempty open set VV contained in UU such that xVx \in \overline{V}.

Choose a point yUy \in U such that yxy \neq x; this is possible if xUx \in U because xx is not an isolated point of XX and it is possible if xUx \notin U since UU \neq \emptyset. Now choose disjoint open sets W1W_1 and W2W_2 about xx and yy, respectively (i.e. xW1x \in W_1 and yW2y \in W_2; this is possible since XX is Hausdorff), then the set V=W2UV=W_2 \cap U is the desired open set; it is contained in UU, it is nonempty because it contains yy, and its closure does not contain xx.

Step 2. We show that given f:Z+Xf : \mathbb{Z}_+ \to X , the function ff is not surjective, which implies that XX is uncountable.

Let xn=f(n)x_n = f(n). Apply Step 1 to the nonempty open set U=XU = X to choose a nonempty open set V1XV_1 \subseteq X such that x1V1x_1 \notin \overline{V_1}. In general, given Vn1V_{n-1} open and nonempty, choose VnV_n to be a nonempty open set such that VnVn1V_n \subseteq V_{n-1} and xnVnx_n \notin \overline{V_n}.

Consider the nested sequence V1V2\overline{V_1} \supseteq \overline{V_2} \supseteq \cdots of nonempty closed sets of XX. Since XX is compact, by Theorem 26.9, there is a point xVnx \in \bigcap \overline{V_n}. Now xx cannot equal xnx_n for any nn, since xx belongs to Vn\overline{V_n} and xnx_n does not.

info

[Corollary 27.8] (from Theorem 27.7) Every closed interval in R\mathbb{R} is uncountable.

Every closed interval II in R\mathbb{R} is nonempty, compact and Hausdorff. It also has no isolated points, since if we consider II as a subspace of R\mathbb{R}, the intersection of II with every nonempty open set of R\mathbb{R} must contain more than one point.

Limit Point Compactness

Definition. A space XX is said to be limit point compact if every infinite subset of XX has a limit point.

[Theorem 28.1] Compactness implies limit point compactness, but the converse is not true.

Definition. (Subsequence and sequential compactness). Let XX be a topological space. If (xn)(x_n) is a sequence of points of XX, and if 0<n1<n2<<ni<0 < n_1 < n_2 < \cdots < n_i < \cdots is an increasing sequence of positive integers, then the sequence (yi)(y_i) defined by setting yi=xniy_i = x_{n_i} is called a subsequence of the sequence (xn)(x_n). The space XX is said to be sequentially compact if every sequence of poitns of XX has a convergent subsequence.

note

The notion of sequential compactness is closely related to the Bolzano-Weierstrass theorem.

[Theorem 28.2] Let XX be a metrizable space, then XX is compact iff XX is limit point compact iff XX is sequentially compact.

Local Compactness

Definition. A space XX is said to be locally compact at xx if there is some compact subspace CC of XX that contains a neighbourhood of xx. If XX is locally compact at each of its points, XX is said simply to be locally compact.

Note that a compact space is automatically locally compact.

[Theorem 29.1] Let XX be a space, then XX is locally compact Hausdorff if and only if there exists a space YY satisfying the following conditions:

  1. XX is a subspace of YY.
  2. The set Y\XY \backslash X consists of a single point.
  3. YY is a compact Hausdorff space.

In fact, such YY can be constructed by adjoining a point not in XX to XX, and defining its topology to consist of all sets UU that are open in XX and all sets of the form Y\CY \backslash C where CC is a compact subspace of XX.

If YY and YY' are two spaces satisfying the conditions above, then there is a homeomorphism of YY with YY' which equals the identity map on XX. (Another way of phrasing this is that the existence of such space is unique up to a homeomorphism that maps XX identically.)

Definition. If YY is a compact Hausdorff space and XX is a proper subspace of YY whose closure equals YY, then YY is said to be a compactification of XX. If Y\XY \backslash X is a single point, then YY is called the one-point compactification of XX (a.k.a. Alexandroff compactification).

info

From Theorem 29.1, we see that XX has a one-point compactification YY if and only if XX is a locally compact Hausdorff space that is not itself compact, since in this case the point of Y\XY \backslash X is a limit point of XX, so that X=Y\overline{X}=Y.

If XX itself happens to be compact, then the space YY obtained from Theorem 29.1 is obtained from XX by adjoining a single isolated point.

[Theorem 29.2] (Equivalent formulation of local compactness for Hausdorff spaces) Let XX be a Hausdorff space, then XX is locally compact if and only if given xx in XX, and given a neighbourhood UU of xx, there is a neighbourhood VV of xx such that V\overline{V} is compact and VU\overline{V} \subseteq U.

[Corollary 29.3] Let XX be locally compact Hausdorff and AA be a subspace of XX. If AA is closed in XX or open in XX, then AA is locally compact.

[Corollary 29.4] A space XX is homeomorphic to an open subspace of a compact Hausdorff space if and only if XX is locally compact Hausdorff.

Suppose that XX is locally compact Hausdorff. If XX itself is compact, then XX itself is an open subspace of a compact Hausdorff space that is homeomorphic to XX, as desired. If XX is not compact, then XX has a one-point compactification YY given by Theorem 29.1, such that XX is an open subspace of YY, which is as desired because again, XX is indeed homeomorphic to itself.

Suppose that XX is homeomorphic to an open subspace of a compact Hausdorff space. If XX is compact, then XX must be locally compact and the assumed homeomorphism implies that XX is also Hausdorff due to Exercise 17.12, so we are done. If XX is not compact, then the one-point compactification YY given by Theorem 29.1 has XX, which is Hausdorff, as its open subspace. It is also locally compact by Corollary 29.3.