Skip to main content

Notes

Topology and Topological Spaces: A Primer

The study of topological spaces is a generalisation of the concepts of metric spaces, as well as open and closed sets, both of which are important in real analysis. Informally, a topological space is a set that has a topology, which "decides" which subset is open or not.

Definition. A topology T\mathcal{T} on a set XX is a collection of subsets of XX (in other words, TP(X)\mathcal{T} \subseteq \mathcal{P}(X)) which has the following properties:

  1. The empty set \emptyset and XX itself are in T\mathcal{T}.
  2. Any arbitrary union of subcollections of T\mathcal{T} is in T\mathcal{T}.
  3. Any finite intersection of subcollections of T\mathcal{T} is in T\mathcal{T}.

We can then say that a subset UU of XX is an open set if XX is in T\mathcal{T}. One can compare this with the metric-space definition of open sets, which is the definition often adapted in real analysis courses and can allow us to derive the topological definition of open sets in XX, as described above.

One may then ask: what if it is the other way round, i.e. is it possible to produce a metric space from a topological space? In fact, this is a question that attracts so much interest that mathematicians assign a special name to those that have this property: metrizable spaces, which we shall learn later in this chapter.

This may look obvious to some, but for the sake of reminder: it is possible for a set XX to have multiple topologies. There is also a not-so-obvious one: different sets can have the same topology(!).

Concepts for Topologies

Fineness and Coarseness

Definition. Given (any) two topologies T,T\mathcal{T}, \mathcal{T'} of a set XX, we say that T\mathcal{T'} is finer than T\mathcal{T} if TT\boxed{\mathcal{T'} \supseteq \mathcal{T}}. We say T\mathcal{T'} is strictly finer than T\mathcal{T} if TT\boxed{\mathcal{T'} \supset \mathcal{T}}.

Conversely, if TT\boxed{\mathcal{T'} \subseteq \mathcal{T}} (resp. TT\boxed{\mathcal{T'} \subset \mathcal{T}}), then we say that T\mathcal{T'} is coarser (resp. strictly coarser) than T\mathcal{T}.

We say T\mathcal{T} and T\mathcal{T'} are comparable if either TT\mathcal{T'} \supseteq \mathcal{T} or TT\mathcal{T'} \subseteq \mathcal{T}, i.e. if it is possible to say if one is finer or coarser than the other.

Informally, a topology is finer if it "contains more detail" or "has a finer scale" than the other topology.

A visualisation of topological coarseness and fineness using ruler markings.

A visualisation of topological coarseness and fineness using ruler markings.

Note that if T\mathcal{T} and T\mathcal{T}' are coarser/finer than each other, then they are the same topology.

Basis (for a Topology)

Basic building blocks of a topology.

Definition. Let XX be a set. A basis for a topology T\mathcal{T} on XX is a collection of subsets of XX (which are called basis elements), denoted here as B\mathcal{B}, such that for each xXx \in X:-

  1. There is at least one basis element BB that contains xx, i.e. xBx \in B.
  2. If xB1B2x \in B_1 \cap B_2, where B1B_1 and B2B_2 are (two) basis elements, then there is a (third) basis element B3B_3 such that xB3x \in B_3 and B3B_3 is contained by the intersection of B1B_1 and B2B_2, i.e. B3B1B2B_3 \subseteq B_1 \cap B_2. (Note: this is immediately true if B1=B2B_1 = B_2: the desired B3B_3 will just be B1=B2 B_1 = B_2 itself.)

The topology T\mathcal{T} generated by the basis B\mathcal{B} is then defined as: if a subset UU of XX is open (i.e. UTU \in \mathcal{T}), then for each xUx \in U, there is a basis element BB B \in \mathcal{B} such that xBx \in B and BUB \subseteq U. From the definition, we can see that the basis elements themselves are in the topology generated from them.

Why does this definition of basis-generated topology do give us a topology?

We need to check that a topology T\mathcal{T} generated by a basis B\mathcal{B} satisfies the basic definition of topology.

If U=U = \emptyset, then UTU \in \mathcal{T} vacuously because we cannot find any xUx \in U to check the condition of openness. If U=XU = X, then since every basis element BBB \in \mathcal{B} satisfies BXB \subseteq X, as well as the first condition for a basis, it also satisfies this condition of openness and is thus in T\mathcal{T}.

To check that any arbitrary union of T\mathcal{T}, i.e. U=αJUαU = \bigcup_{\alpha \in J}U_\alpha, where {Uα}αJ\{U_\alpha\}_{\alpha \in J} is an indexed family of sets in T\mathcal{T}, is still in T\mathcal{T}, we see that since each UαU_\alpha is open, by definition, there is a basis element BB such that xBUαx \in B \subseteq U_\alpha. Note that UαUU_\alpha \subseteq U for any αJ\alpha \in J, so we have BUB \subseteq U as well. This shows that UU is open and hence in T\mathcal{T}.

To check that any finite intersection of T\mathcal{T}, i.e. U=i=1nUiU = \bigcap_{i=1}^{n}U_i, where U1,,UnTU_1,\ldots, U_n \in \mathcal{T} and nNn \in \mathbb{N}, is in T\mathcal{T}, we can do so by induction. We first settle the base case: for n=1n = 1, U1U_1 itself is already in T\mathcal{T}, so it immediately holds.

When n=2n = 2 so that we have U1U2U_1 \cap U_2, given xU1U2x \in U_1 \cap U_2, by definition, we can choose basis elements B1B_1 and B2B_2 such that xB1x \in B_1 (resp. B2B_2) and B1U1B_1 \subseteq U_1 (resp. B2U2B_2 \subseteq U_2). By the definition of a basis, we can obtain a basis element B3B_3 such that xB3x \in B_3 and B3B1B2B_3 \subseteq B_1 \cap B_2. Since B1B2U1U2B_1 \cap B_2 \subseteq U_1 \cap U_2, this gives us B3U1U2B_3 \subseteq U_1 \cap U_2, so U1U2TU_1 \cap U_2 \in \mathcal{T}, by definition.

This argument above can be used to prove the inductive case. For a finite intersection U1UnU_1 \cap \cdots \cap U_n of elements of T\mathcal{T}, we first suppose that U1Un1U_1 \cap \cdots \cap U_{n-1} does belong to T\mathcal{T}. Now, note that

(U1Un)=(U1Un1)Un.(U_1 \cap \cdots \cap U_n) = (U_1 \cap \cdots \cap U_{n-1}) \cap U_n.

Therefore, by induction hypothesis and applying the argument above, we can conclude that any finite intersection of elements in T\mathcal{T} belongs to T\mathcal{T}.

We have then checked that all three conditions of a topology are satisifed, so T\mathcal{T} indeed gives us a topology.

If XX is any set, the collection of all one-point subsets (singletons) of XX is a basis for the discrete topology on XX. Why?

Let XX be a set and B\mathcal{B} be the collection of all one-point subsets of XX. Since there is a one-to-one correspondence, specifically x{x}x \mapsto \{x\}, between XX and B\mathcal{B}, this allows us to conclude that for any xXx \in X, there exists at least one basis element BBB \in \mathcal{B} such that xBx \in B, giving us the first condition for a basis. Due to this one-to-one correspondence as well, there is no elements in XX that belongs to two different basis elements, so the second condition for a basis holds vacuously.

Any arbitrary union of the one-point subsets BBB \in \mathcal{B} will give us P(X)\{}\mathcal{P}(X) \backslash \{\emptyset\}, i.e. the collection of all subsets of XX except for the empty set, and this indeed includes XX itself. Any finite intersection of BB will then give us the last piece of the puzzle: the empty set \emptyset.

This finally gives us the complete P(X)\mathcal{P}(X), i.e. the discrete topology on XX.

[Lemma 13.1] T\mathcal{T} as a topology on a set XX that is generated by a basis B\mathcal{B} can also be described as the collection of all unions of elements of B\mathcal{B}. Why?

For any given collection of elements of B\mathcal{B}, since they are also elements of T\mathcal{T} and T\mathcal{T} is a topology, their union is also in T\mathcal{T}.

Conversely, for any UTU \in \mathcal{T}, we can find for each xUx \in U a basis element BxBB_x \in \mathcal{B} such that xBxUx \in B_x \subseteq U, so U=xUBxU = \bigcup_{x \in U}B_x, and thus UU equals a union of elements of B\mathcal{B}.

warning

This means that any open set UU in XX can be expressed a union of basis elements, which sounds analogous to bases in linear algebra in which vectors are expressed as a linear combination of basis vectors. However, such expression needs not be unique, which makes it different from bases in linear algebra, which are unique.

In reverse, we can also obtain a basis for a given topology, which is a frequently cited fact in the study of this topic.

[Lemma 13.2] Let XX be a topological space. Suppose that C\mathcal{C} is a collection of open sets of XX such that for each open set UU of XX and each xx in UU, there is an element CC of C\mathcal{C} such that xCUx \in C \subseteq U, then C\mathcal{C} is a basis for the topology of XX. Why?

We need to check that C\mathcal{C} satisifies the defining conditions of a (topological) basis. Since XX is a topological space and thus is an open set itself, by the hypothesis for C\mathcal{C}, for each xUx \in U, there is an element CC of C\mathcal{C} such that xCXx \in C \subseteq X, so the first condition is satisfied.

To check the second condition, suppose that xC1C2x \in C_1 \cap C_2, where C1,C2CC_1,C_2 \in \mathcal{C}. Since C1C_1 and C2C_2 are open, so is C1C2C_1 \cap C_2. Therefore, there exists by hypothesis an element C3C_3 in C\mathcal{C} such that xC3C1C2x \in C_3 \subseteq C_1 \cap C_2.

Let T\mathcal{T} be the collection of open sets of XX; we also need to show that the topology T\mathcal{T}' generated by C\mathcal{C} is indeed equal to the topology T\mathcal{T}. First, note that if UU belongs to T\mathcal{T} and if xUx \in U, then there is by hypothesis an element CC of C\mathcal{C} such that xCUx \in C \subseteq U, so UU belongs to the (generated) topology T\mathcal{T}'. Conversely, if WW belongs to the (generated) topology T\mathcal{T}', then WW equals a union of elements of C\mathcal{C}, by Lemma 13.1. Since each element of C\mathcal{C} belongs to T\mathcal{T} and T\mathcal{T} is a topology, WW also belongs to T\mathcal{T}.

[Lemma 13.3] Let B\mathcal{B} and B\mathcal{B}' be bases for the topologies T\mathcal{T} and T\mathcal{T}', respectively, on XX, then T\mathcal{T}' is finer than T\mathcal{T} if and only if for each xXx \in X and each basis element BBB \in \mathcal{B} containing xx, there is a basis element BBB' \in \mathcal{B}' such that xBBx \in B' \subseteq B.

Proof of Lemma 13.3.

(    )(\impliedby). Given an element UU of T\mathcal{T}, we wish to show that UTU \in \mathcal{T}'. Let xUx \in U. Since B\mathcal{B} generates T\mathcal{T}, there is an element BBB \in \mathcal{B} such that xBUx \in B \subseteq U. By hypothesis, there exists an element BBB' \in \mathcal{B}' such that xBBx \in B' \subseteq B, then xBUx \in B' \subseteq U, so UTU \in \mathcal{T}' by definition.

(    )(\implies). Let xXx \in X and BBB \in \mathcal{B}, with xBx \in B. Now BB belongs to T\mathcal{T} by definition and TT\mathcal{T} \subseteq \mathcal{T}' by hypothesis (being finer). Therefore, BTB \in \mathcal{T}'. Since T\mathcal{T}' is generated by B\mathcal{B}', there is an element BBB' \in \mathcal{B}' such that xBBx \in B' \subseteq B.

note

Using Lemma 13.3, we can see that the collection of all circular regions in the plane generates the same topology as the collection of all rectangular regions.

Definition. A subbasis S\mathcal{S} for a topology on XX is a collection of subsets of XX whose union equals XX. The topology generated by the subbasis S\mathcal{S} is defined to be the collection T\mathcal{T} of all (arbitrary) unions of finite intersections of elements of S\mathcal{S}.

Sanity check of the definition of subbasis-generated topology.

It suffices to show that the collection B\mathcal{B} of all finite intersections of elements of S\mathcal{S} is a basis for T\mathcal{T}, as we can then apply Lemma 13.1 to conclude that the collection T\mathcal{T} of all unions of elements of B\mathcal{B} is indeed a topology as desired.

Given xXx \in X, it belongs to an element of S\mathcal{S} by definition of subbasis, so it also belongs to an element of B\mathcal{B}; this shows the first condition for a basis. To check the second condition (that involves intersection), consider the following two elements of B\mathcal{B},

B1:=S1SmandB2:=S1Sn.B_1 := S_1 \cap \cdots \cap S_m \quad \text{and} \quad B_2 := S_1' \cap \cdots \cap S_n'.

Their intersection

B1B2=(S1Sm)(S1Sn)B_1 \cap B_2 = (S_1 \cap \cdots \cap S_m) \cap (S_1' \cap \cdots \cap S_n')

is also a finite intersection of elements of S\mathcal{S}, so it belongs to B\mathcal{B}.

note

Munkres' definition of subbasis here is different from that given in the corresponding Wikipedia article. In Wikipedia's definition, the topology generation is considered given, rather than a consequence.

It is also worth noting that in Rudin's Functional Analysis, a subbasis is defined to be a subcollection BB of an established topology such that the collection of open sets consisting of all finite intersections of elements of BB forms a basis for the topology. This is something that requires proof in Munkres, but is taken for granted in Rudin.

tip

Don't think of the use of the prefix 'sub' in the word 'subbasis' here in the same way as the way as 'subset' does, but more like the word 'subpar': it doesn't mean that it is contained by a larger (topological) basis as a collection of basis elements; instead, think of it as 'something that is not quite a basis yet but serves a foundation to it.'

Important Topologies

There are several main topologies covered in this chapter. Some of them are so important that they deserve a dedicated subsection, whereas some others that are not so much are just included as an item of a bullet list here. Here, we denote a set as XX, and its equipped topology, unless otherwise stated, as T\mathcal{T}.

  • Discrete topology: the collection of all subsets of XX.
  • Indiscrete/trivial topology: the other end of the extreme, i.e. the collection of only \emptyset and XX itself.
  • Finite complement topology (a.k.a. cofinite topology), Tf\mathcal{T}_f: the collection of all subsets UU of XX such that X\UX \backslash U is either finite or XX itself (so that U=U = \emptyset).
  • Countable complement topology, Tc\mathcal{T}_c: the collection of all subsets UU of XX such that X\UX \backslash U is either countable or XX itself (so that U=U = \emptyset).
  • Standard topology (on the real line), R\mathbb{R}: the topology generated by the collection of all open intervals in the real line, (a,b)={x:a<x<b}(a,b)=\{x : a < x < b\}. This is assumed to be the default topology for R\mathbb{R}, unless otherwise stated.
  • Lower limit topology, Rl\mathbb{R}_l: the topology generated by the collection of all half-open intervals of the form [a,b)={x:ax<b}[a,b) = \{x : a \leqslant x < b\}, where a<ba < b.
  • K-topology, RK\mathbb{R}_K: the topology generated by the collection of all open intervals (a,b)(a,b), along with all sets of the form (a,b)\K(a,b) \backslash K, where K={1n:nZ+}K = \left\{\frac{1}{n} : n \in \mathbb{Z}_+ \right\}.

[Lemma 13.4] The topologies of Rl\mathbb{R}_l and RK\mathbb{R}_K are strictly finer than the standard topology on R\mathbb{R}, but are not comparable with one another.

Proof of Lemma 13.4.

Let T\mathcal{T}, T\mathcal{T}' and T\mathcal{T}'' be the standard R\mathbb{R}-topology, lower limit topology and K-topology respectively. Lemma 13.3 will be applied in the proof here.

Given a basis element (a,b)(a,b) for T\mathcal{T} and a point xx inside (a,b)(a,b), the basis element [x,b)[x,b) for T\mathcal{T}' contains xx and lies in (a,b)(a,b). On the other hand, given the basis element [x,d)[x,d) for T\mathcal{T}', there is no open interval (a,b)(a,b) that contains xx and lies in [x,d)[x,d). Thus T\mathcal{T}' is strictly finer than T\mathcal{T}.

Similarly, given a basis element (a,b)(a,b) for T\mathcal{T} and a point xx of (a,b)(a,b), this same interval is a basis element for T\mathcal{T}'' that contains xx. On the other hand, consider the basis element B=(1,1)\KB = (-1,1) \backslash K for T\mathcal{T}'' and the point 00 of B\mathcal{B}, there is no open interval that contains 00 and lies in BB (any open interval that contains 00 must contain 1n\frac{1}{n} for some nZ+n \in \mathbb{Z}_+). Thus, T\mathcal{T}'' is strictly finer than T\mathcal{T}.

The proof for the non-comparability between Rl\mathbb{R}_l and RK\mathbb{R}_K is a solution to Exercise 13.6.

Order Topology

An order topology is the generalisation of the topologies on R\mathbb{R} that are generated by intervals to any set with a simple order (a.k.a. strict total order).

Definition. Let XX be a set with a simple order relation and assume that XX has more than one element. Let B\mathcal{B} be the collection of all sets of the following types:

  1. All open intervals (a,b)(a,b) in XX.
  2. All intervals of the form [a0,b)[a_0,b), where a0a_0 is the smallest element (if any) of XX.
  3. All intervals of the form (a,b0](a,b_0], where b0b_0 is the largest element (if any) of XX.

The collection B\mathcal{B} is then a basis for a topology on XX, which is called the order topology.

note

If XX has no smallest element, there are no sets of the second type. If XX has no largest element, there are no sets of the third type.

Check that the collection B\mathcal{B} is a basis.

  1. Indeed, every element xx of XX lies in at least one element of B\mathcal{B}: the smallest element (if any) is in all sets of the second type, the largest element (if any) is in all sets of the third type and every other element is in a set of the first type.

  2. The intersection of any two sets of the preceding types is again a set of one of these types, or is empty (if the two sets are disjoint).

    Assuming the two sets are not disjoint,

    • the intersection of two open intervals (a,b)(a,b) and (a,b)(a',b') is (max{a,a},min{b,b})(\max\{a,a'\}, \min\{b,b'\}), an open interval;
    • the intersection of two intervals (a,b)(a,b) and [a0,b)[a_0, b') is (a,min{b,b})(a,\min\{b,b'\});
    • the intersection of two intervals (a,b)(a,b) and (a,b0](a',b_0] is (max{a,a},b)(\max\{a,a'\},b);
    • the intersection of two intervals [a0,b)[a_0,b) and (a,b0](a,b_0] is (a,b)(a,b).
    • the intersection of two intervals [a0,b)[a_0,b) and [a0,b)[a_0,b') is [a0,min{b,b})[a_0,\min\{b,b'\});
    • the intersection of two intervals (a,b0](a,b_0] and (a,b0](a',b_0] is (max{a,a},b0](\max\{a,a'\}, b_0].

Definition. If XX is an ordered set, and aa is an element of XX, there are four subsets of XX that are called rays determined by aa. They are the following:

(a,+)={x:x>a}(a,+\infty) = \{x : x > a\} (,a)={x:x<a}(-\infty,a) = \{x : x < a\} [a,+)={x:xa}[a,+\infty) = \{x : x \geqslant a\} (,a]={x:xa}(-\infty,a] = \{x : x \leqslant a\}

Sets of the first two types are called open rays, and sets of the last two types are called closed rays.

:::

Subspace Topology (a.k.a. Induced Topology)1

Definition. Let XX be a topological space with a topology T\mathcal{T}. If YY is a subset of XX, then the collection

TY={YU:UT}\mathcal{T}_Y = \{ Y \cap U : U \in \mathcal{T} \}

which are all intersections of open sets of XX with YY, is a topology on YY, and we call it the subspace topology. With this topology, YY is called a subspace of XX.

[Lemma 16.1] If B\mathcal{B} is a basis for the topology of X X then the collection

BY={BY:BB}\mathcal{B}_Y = \{B \cap Y : B \in \mathcal{B}\}

is a basis for the subspace topology on YY.

[Lemma 16.2] Let YY be a subspace of XX. If UU is open in YY and YY is open in XX, then UU is open in XX.

[Theorem 16.3] If AA is a subspace of XX and BB is a subspace of YY, then the product topology on A×BA \times B is the same as the topology A×BA \times B inherits as a subspace of X×YX \times Y.

Given an ordered set XX, let us say that a subset YY of XX is convex in XX if for each pair of points a<ba < b of YY, the entire interval (a,b)(a,b) of points of XX lies in YY.

(Note that this implies that intervals and rays in XX are convex in XX.)

[Theorem 16.4] Let XX be an ordered set in the order topology; let YY be a subset of XX that is convex in XX. Then the order topology on YY is the same as the topology YY inherits as a subspace of XX.

warning

Note that without the restriction that YY needs to be convex in XX, Theorem 16.4 is not true in general.

A counterexample is given as follows: let I=[0,1]I = [0,1]. The dictionary order on I×II \times I is just the restriction to I×II \times I of the dictionary order on the plane R×R\mathbb{R} \times \mathbb{R}. However, the dictionary order topology on I×II \times I is NOT the same as the subspace topology on I×II \times I obtained from the dictionary topology on R×R\mathbb{R} \times \mathbb{R}! For example, the set U:={12}×(12,1]U := \left\{\frac{1}{2}\right\} \times \left(\frac{1}{2}, 1\right] is open in I×II \times I in the subspace topology, as it is the intersection of I×II \times I with V:=((12,12),(12,50))V := \left(\left(\frac{1}{2},\frac{1}{2}\right), \left(\frac{1}{2}, 50\right)\right) and VV is open in R×R\mathbb{R} \times \mathbb{R}, but UU is not open in the order topology, as any basis element for the order topology on I×II \times I that contains UU must contain points of I×II \times I such that the first coordinate is between 12\frac{1}{2} and 11.

Indeed, I×II \times I is not convex because (0.5,50)(0.5, 50) lies in the interval ((0.5,0.5),(1,0.5))((0.5, 0.5), (1, 0.5)), where (0.5,0.5),(1,0.5)I×I(0.5, 0.5), (1, 0.5) \in I \times I, but (0.5,50)I×I(0.5, 50) \notin I \times I.

(The set I×II \times I in the dictionary order topology is called the ordered square, denoted by Io2I_o^2.)

info

To avoid ambiguity, whenever XX is an ordered set in the order topology and YY is a subset of XX, we shall assume that YY is given the subspace topology unless specifically stated otherwise.

Product Topology (+ Box Topology)

Forming new topologies from old ones.

Product Topology on X×YX \times Y

Definition. Let XX and YY be topological spaces. The product topology on X×YX \times Y is the topology generated by the collection B\mathcal{B} of all sets of the form U×VU \times V as the basis, where UU is an open subset of XX and VV is an open subset of YY.

Check that the collection B\mathcal{B} is a basis.

  1. Indeed, X×YX \times Y is a basis element (XX and YY themselves open), and they contain all of the elements of X×YX \times Y.

  2. The intersection of any two basis elements U1×V1U_1 \times V_1 and U2×V2U_2 \times V_2 is another basis element, since (U1×V1)(U2×V2)=(U1U2)×(V1V2)(U_1 \times V_1) \cap (U_2 \times V_2) = (U_1 \cap U_2) \times (V_1 \cap V_2).

warning

Note that the collection B\mathcal{B} itself is not a topology on X×YX \times Y, but it is open in X×YX \times Y.

[Theorem 15.1] If B\mathcal{B} is a basis for the topology of XX and C\mathcal{C} is a basis for the topology of YY, then the collection

D={B×C:BB and CC}\mathcal{D} = \{ B \times C : B \in \mathcal{B} \text{ and } C \in \mathcal{C}\}

is a basis for the topology of X×YX \times Y.

Definition. Define the projections π1:X×YX\pi_1 : X \times Y \to X by π1(x,y)=x\pi_1(x,y)=x and π2:X×YY\pi_2 : X \times Y \to Y by π2(x,y)=y\pi_2(x,y)=y.

It follows that π11(U)\pi_1^{-1}(U) where UU is open in XX is U×YU \times Y and π21(V)\pi_2^{-1}(V) where VV is open in YY is X×VX \times V, so π11(U)π21(V)=U×V\pi_1^{-1}(U) \cap \pi_2^{-1}(V) = U \times V.

[Theorem 15.2] The collection

S={π11(U):U open in X}{π21(V):V open in Y}\mathcal{S} = \{\pi_1^{-1}(U) : U \text{ open in } X\} \cup \{ \pi_2^{-1}(V) : V \text{ open in } Y\}

is a subbasis for the product topology on X×YX \times Y.

Product Topology in General + Box Topology

There are two possible ways to define a product topology on more general Cartesian products with finite or infinite dimensions, i.e. Cartesian products in the form of X1××XnX_1 \times \cdots \times X_n and X1×X2×X_1 \times X_2 \times \cdots, where each XiX_i is a topological space, which are the following:

  1. Box topology: topology generated by taking as basis all sets of the form U1××UnU_1 \times \cdots \times U_n in the first case, and of the form U1×U2×U_1 \times U_2 \times \cdots in the second case, where UiU_i is an open set of XiX_i for each ii.
  2. Product topology: topology generated by taking as a subbasis all sets of the form πi1(Ui)\pi_i^{-1}(U_i) (recall that πi\pi_i represents the projection function), where ii is any index and UiU_i is an open set of XiX_i.

How do these topologies differ? Let BB be a basis element for the second topology, then it is a finite intersection of subbasis elements π11(Ui)\pi_1^{-1}(U_i), say for i=i1,,iki = i_1, \ldots, i_k, then a point x\mathbf{x} belongs to BB if and only if πi(x)\pi_i(\mathbf{x}) belongs to UiU_i for i=i1,,iki=i_1,\ldots,i_k; there is no restriction on πi(x)\pi_i(\mathbf{x}) for other values of ii.

It follows that the two topologies above agree for the finite Cartesian product but differ for the infinite product.

Now, we shall introduce a more general notion of Cartesian products, which are formed by a family of sets indexed by an arbitrary set, instead of a subset of Z+\mathbb{Z}_+.

Definition. (Tuples with arbitrary index). Let JJ be an index set. Given a set XX, we define a JJ-tuple of elements of XX to be a function x:JX\mathbf{x} : J \to X. If α\alpha is an element of JJ, we often denote the value of x\mathbf{x} at α\alpha by xαx_\alpha rather than x(α)\mathbf{x}(\alpha); we call the the α\alpha -th coordinate of x\mathbf{x}. We also often denote the function x\mathbf{x} itself by the symbol (xα)αJ(x_\alpha)_{\alpha \in J}, which is as close as we can come to a 'tuple notation' for an arbitrary index set JJ. We denote the set of all JJ-tuple of elements of XX by XJX^J.

Definition. (Cartesian products with arbitrary index). Let {Aα}αJ\{A_\alpha\}_{\alpha \in J} be an indexed family of sets; let X=αJAαX = \bigcup_{\alpha \in J}A_\alpha. The Cartesian product of this indexed family, denoted by αJAα\prod_{\alpha \in J}A_\alpha, is defined to be the set of all JJ-tuples (xα)αJ(x_\alpha)_{\alpha \in J} of elements of XX such that xαAαx_\alpha \in A_\alpha for each αJ\alpha \in J, i.e. it is the set of all functions x:JαJAα\mathbf{x} : J \to \bigcup_{\alpha \in J}A_\alpha such that x(α)Aα\mathbf{x}(\alpha) \in A_\alpha for each αJ\alpha \in J.

note

If such Cartesian product is formed by an indexed family such that all its component sets are non-empty, the assertion that this Cartesian product is non-empty, i.e. there exists such function x\mathbf{x}, is equivalent to the Axiom of Choice.

A proof of this equivalence is a solution to Exercise 19.9.

Definition. (Box topology for arbitrarily indexed Cartesian products). Let {Xα}αJ\{X_\alpha\}_{\alpha \in J} be an indexed family of topological spaces. Let us take as a basis for a topology on the product space αJXα\prod_{\alpha \in J}X_\alpha the collection of all sets of the form αJUα\prod_{\alpha \in J}U_\alpha, where UαU_\alpha is open in XαX_\alpha, for each αJ\alpha \in J. The topology generated by this basis is called the box topology.

Definition. (Projection mapping for arbitrarily indexed Cartesian products). Generalising the subbasis formulation of the definition, we call the function πβ:αJXαXβ\pi_\beta : \prod_{\alpha \in J}X_\alpha \to X_\beta that assigns to each element of the product space its β\beta-th coordinate, i.e. πβ((xα)αJ)=xβ\pi_\beta((x_\alpha)_{\alpha \in J})=x_\beta the projection mapping associated with the index β\beta.

Definition. (Product topology and product spaces for arbitrarily indexed Cartesian products). Let Sβ\mathcal{S}_\beta denote the collection Sβ={πβ1(Uβ):Uβ open in Xβ}\mathcal{S}_\beta=\{ \pi_\beta^{-1}(U_\beta) : U_\beta \text{ open in } X_\beta\}, and let S\mathcal{S} denote the union of these collections, i.e. S=βJSβ\mathcal{S}=\bigcup_{\beta \in J}\mathcal{S}_\beta. The topology generated by the subbasis S\mathcal{S} is called the product topology. In this topology αJXα\prod_{\alpha \in J}X_\alpha is called a product space.

[Theorem 19.1] (Comparision of the box and product topologies). The box topology on Xα\prod X_\alpha has as basis all sets of the form Uα\prod U_\alpha, where UαU_\alpha is open in XαX_\alpha for each α\alpha. The product topology on Xα\prod X_\alpha has as basis all sets of the form Uα\prod U_\alpha, where UαU_\alpha is open in XαX_\alpha for each α\alpha and UαU_\alpha equals XαX_\alpha except for finitely many values of α\alpha.

Consider the basis B\mathcal{B} that S\mathcal{S} (as defined above) generates. The collection B\mathcal{B} consists of all finite intersections of elements of S\mathcal{S}. We consider two cases.

Case I: the elements in the intersection belong to the same one of the sets Sβ\mathcal{S}_\beta. We then do not get anything new, because

πβ1(Uβ)πβ1(Vβ)=πβ1(UβVβ)=πβ1(UβVβ)\pi_\beta^{-1}(U_\beta) \cap \pi_\beta^{-1}(V_\beta) = \pi_\beta^{-1}(U_\beta \cap V_\beta)=\pi_\beta^{-1}(U_\beta \cap V_\beta)

is still an element of Sβ\mathcal{S}_\beta. The same applies for finitely many elements of Sβ\mathcal{S}_\beta.

Case II: the elements in the intersection come from different sets in Sβ\mathcal{S}_\beta. The typical element of the basis B\mathcal{B} can then be described as follows: Let β1,,βn\beta_1, \ldots, \beta_n be a finite set of distinct indices from the index set JJ, and let UβiU_{\beta_i} be an open set in XβiX_{\beta_i} for i=1,,ni=1,\ldots,n, then

B=πβ11(Uβ1)πβ21(Uβ2)πβn1(Uβn)B=\pi_{\beta_1}^{-1}(U_{\beta_1}) \cap \pi_{\beta_2}^{-1}(U_{\beta_2}) \cap \cdots \cap \pi_{\beta_n}^{-1}(U_{\beta_n})

is also the typical element of B\mathcal{B}.

Now, a point x=(xα)\mathbf{x}=(x_\alpha) is in BB if and only if its β1\beta_1-th coordinate is in Uβ1U_{\beta_1}, its β2\beta_2-th coordinate is in Uβ2U_{\beta_2}, and so on. There is no restriction whatsoever on the α\alpha-th coordinate of x\mathbf{x} if α\alpha is not one of the indices β1,,βn\beta_1,\ldots,\beta_n. Hence, we can write BB as the product

B=αJUαB=\prod_{\alpha \in J}U_\alpha

where UαU_\alpha denotes the entire space XαX_\alpha if αβ1,,βn\alpha \neq \beta_1,\ldots,\beta_n.

info

Since a number of key theorems about finite products will also hold for arbitrary products if we use the product topology, but note if we use the box topology, whenever we consider the product Xα\prod X_\alpha, we shall assume it is given the product topology unless specifically stated otherwise.

  • [Theorem 19.2] (Basis inheritance). Suppose the topology on each space XαX_\alpha is given by a basis Bα\mathcal{B}_\alpha, then the collection of all sets of the form αJBα\prod_{\alpha \in J}B_\alpha, where BαBαB_\alpha \in \mathcal{B}_\alpha for each α\alpha, will serve as a basis for the box topology on αJXα\prod_{\alpha \in J} X_\alpha.

    The collection of all sets of the same form, where BαBαB_\alpha \in \mathcal{B}_\alpha for finitely many indices α\alpha and Bα=XαB_\alpha = X_\alpha for all the remaining indices, will serve as a basis for the product topology αJXα\prod_{\alpha \in J}X_\alpha.

  • [Theorem 19.3] (Subspace inheritance). Let AαA_\alpha be a subspace of XαX_\alpha, for each αJ\alpha \in J, then Aα\prod A_\alpha is a subspace of Xα\prod X_\alpha if both products are given the box topology, or if both products are given the product topology.

  • [Theorem 19.4] (Hausdorff inheritance) If each space XαX_\alpha is Hausdorff, then Xα\prod X_\alpha is a Hausdorff space in both the box and product topologies.

  • [Theorem 19.5] (Closure inheritance). Let {Xα}\{X_\alpha\} be an indexed family of spaces; let AαXαA_\alpha \subseteq X_\alpha for each α\alpha. If Xα\prod X_\alpha is given either the product or the box topology, then

Aα=Aα.\prod \overline{A_\alpha}=\overline{\prod A_\alpha}.

The proofs of Theorems 19.2, 19.3 and 19.4 are given in a solution to Exercises 19.1, 19.2 and 19.3 respectively.

Proof of Theorem 19.5.

Let x=(xα)\mathbf{x}=(x_\alpha) be a point of Aα\prod \overline{A_\alpha}. We show that xAα\mathbf{x} \in \overline{\prod A_\alpha}. Let U=UαU = \prod U_\alpha be a basis element for either the box or product topology that contains x\mathbf{x}. Since xαAαx_\alpha \in \overline{A_\alpha}, we can chooise a point yαUαAαy_\alpha \in U_\alpha \cap A_\alpha for each α\alpha (see Theorem 17.5 below). Then y=(yα)\mathbf{y}=(y_\alpha) belongs to both UU and Aα\prod A_\alpha. Since UU is arbitrary, it follows that x\mathbf{x} belongs to the closure of Aα\prod A_\alpha.

Conversely, suppose x=(xα)\mathbf{x}=(x_\alpha) lies in the closure of Aα\prod A_\alpha, in either topology. We show that for any given index β\beta, we have xβAβx_\beta \in \overline{A_\beta}. Let Vβ V_\beta be an arbitrary open set of XβX_\beta containing xβx_\beta. Since πβ1(Vβ)\pi_\beta^{-1}(V_\beta) is open in Xα\prod X_\alpha in either topology, it contains a point y=(yα)\mathbf{y}=(y_\alpha) of Aα\prod A_\alpha. Then yβy_\beta belongs to VβAβV_\beta \cap A_\beta. It follows that xβAβx_\beta \in \overline{A_\beta}.

[Theorem 19.6] (Continuity inheritance). Let f:AαJXαf: A \to \prod_{\alpha \in J} X_\alpha be given by the equation

f(a)=(fα(a))αJ,f(a)=(f_\alpha(a))_{\alpha \in J},

where fα:AXαf_\alpha : A \to X_\alpha for each α\alpha. Let Xα\prod X_\alpha have the product topology. Then the function ff is continuous if and only if each function fαf_\alpha is continuous.

warning

Theorem 19.6 fails if we apply the box topology. A counterexample is given as follows:

Consider Rω\mathbb{R}^{\omega}, the countably infinite product of R\mathbb{R} with itself. Recall that Rω=nZ+R\mathbb{R}^{\omega}=\prod_{n \in \mathbb{Z}_+} \mathbb{R}.

Define a function f:RRωf : \mathbb{R} \to \mathbb{R}^{\omega} by the equation

f(t)=(t,t,t,);f(t)=(t,t,t,\ldots);

the nn-th coordinate function of ff is the function fn(t)=tf_n(t)=t, i.e. fn=πnff_n=\pi_n \circ f. Each of the coordinate functions fn:RRf_n : \mathbb{R} \to \mathbb{R} is continuous, but ff is not continuous if Rω\mathbb{R}^{\omega} is given the box topology.

Consider, for example, the basis element nZ+(1n,1n)\prod_{n \in \mathbb{Z}_+}\left(-\frac{1}{n},\frac{1}{n}\right) for the box topology (which is certainly an open subset of Rω\mathbb{R}^{\omega}). We assert that f1(B)f^{-1}(B) is not open in R\mathbb{R}. If f1(B)f^{-1}(B) were open in R\mathbb{R}, it would contain some interval (δ,δ)(-\delta,\delta) (δ>0\delta > 0) about the point 00 (because (0,0,0,)B(0,0,0,\ldots) \in B). This would mean that f((δ,δ))Bf((-\delta,\delta)) \subseteq B, so that, by applying πn\pi_n to both sides of the inclusion, we have

fn((δ,δ))=(δ,δ)(1n,1n)f_n((-\delta,\delta)) = (-\delta,\delta) \subseteq \left(-\frac{1}{n},\frac{1}{n}\right)

for all nn, which is a contradiction as it is impossible.

Metric Topology

Definition. A metric on a set XX is a function d:X×XRd : X \times X \to \mathbb{R} having the following properties:

  1. (Positive definite) d(x,y)0d(x,y) \geqslant 0 for all x,yXx,y \in X; equality holds if and only if x=yx = y.
  2. (Symmetry) d(x,y)=d(y,x)d(x,y)=d(y,x) for all x,yXx,y \in X.
  3. (Triangle inequality) d(x,y)+d(y,z)d(x,z)d(x,y) + d(y,z) \geqslant d(x,z), for all x,y,zXx, y, z \in X.

The number d(x,y)d(x,y), where dd is a metric on XX, is often called the distance between xx and yy in the metric dd.

Given ε>0\varepsilon > 0, the set Bd(x,ε):={y:d(x,y)<ε}B_d(x, \varepsilon) := \{ y : d(x,y)<\varepsilon\} is called the ε\varepsilon-ball centred at xx.. When no confusion arises, the subscript dd can be omitted.

Definition. If dd is a metric on the set XX, then the collection of all ε\varepsilon-balls Bd(x,ε)B_d(x,\varepsilon), for xXx \in X and ε>0\varepsilon>0, is a basis for the topology on XX, called the metric topology induced by dd.

The definition of the metric topology can be rephrased as follows:

A set UU is open in the metric topology induced by dd if and only if for each yUy \in U, there is a δ>0\delta>0 such that Bd(y,δ)UB_d(y, \delta) \subseteq U.

Definition. If XX is a topological space, XX is said to be metrizable if there exists a metric dd on the set XX that induces the topology of XX. A metric space is a metrizable space XX together with a specific metric dd that gives the topology of XX.

Many important topological spaces are metrizable, but some are not.

note

For example, R\mathbb{R} with the order topology is metrizable. Indeed, the Euclidean metric d(x,y):=xyd(x,y) := |x-y| is the metric that induces the topology. To see this, note that each basis element (a,b)(a,b) for the order topology is a basis element for the metric topology; indeed, (a,b)=B(x,ε)(a,b)=B(x,\varepsilon) where x=(a+b)/2x = (a+b)/2 and ε=(ba)/2\varepsilon=(b-a)/2. Conversely, each ε\varepsilon-ball B(x,ε)B(x,\varepsilon) equals an open interval: the interval (xε,x+ε)(x-\varepsilon,x+\varepsilon).

A discrete topology is also metrizable. It can be induced by the discrete metric, defined by d(x,y)={1if xy,0if x=yd(x,y)=\begin{cases} 1 & \text{if } x \neq y, \\ 0 & \text{if } x=y \end{cases}.

Definition. Let XX be a metric space with metric dd. A subset AA of XX is said to be bounded if there is some number MM such that d(a1,a2)Md(a_1,a_2) \leqslant M for every pair a1,a2a_1,a_2 of points of AA.

If AA is bounded and non-empty, the diameter of AA is defined to be the number diam A=sup{d(a1,a2):a1,a2A}\text{diam } A = \sup\{d(a_1,a_2) : a_1,a_2 \in A\}.

[Theorem 20.1] Let XX be a metric space with metric dd. Define dˉ:X×XR\bar{d} : X \times X \to \mathbb{R} by the equation dˉ(x,y)=min{d(x,y),1}\bar{d}(x,y)=\min\{d(x,y),1\}, then dˉ\bar{d} is a metric that induces the same topology as dd.

The metric dˉ\bar{d} is called the standard bounded metric corresponding to dd.

Definition. Given x=(x1,,xn)\mathbf{x} = (x_1,\ldots,x_n) in Rn\mathbb{R}^n, we define the norm of x\mathbf{x} by the equation

x=(x12++xn2)12;\Vert\mathbf{x}\Vert=(x_1^2+\cdots+x_n^2)^{\frac{1}{2}};

we define the Euclidean metric dd on Rn\mathbb{R}^n by the equation

d(x,y)=xy=[(x1y1)2++(xnyn)2]12.d(\mathbf{x},\mathbf{y})=\Vert\mathbf{x}-\mathbf{y}\Vert=[(x_1-y_1)^2+\cdots+(x_n-y_n)^2]^{\frac{1}{2}}.

We define the square metric ρ\rho by the equation

ρ(x,y)=max{x1y1,,xnyn}.\rho(\mathbf{x},\mathbf{y})=\max\{|x_1-y_1|,\ldots,|x_n-y_n|\}.

[Lemma 20.2] Let dd and dd' be two metrics on the set XX; let T\mathcal{T} and T\mathcal{T}' be the topologies they induce, respectively. Then T\mathcal{T}' is finer than T\mathcal{T} if and only if for each xx in XX and each ε>0\varepsilon>0, there exists δ>0\delta>0 such that Bd(x,δ)Bd(x,ε)B_{d'}(x,\delta) \subseteq B_d(x,\varepsilon).

[Theorem 20.3] The topologies on Rn\mathbb{R}^n induced by the Euclidean metric dd and the square metric ρ\rho are the same as the product topology on Rn\mathbb{R}^n.

Let x=(x1,,xn)\mathbf{x}=(x_1,\ldots,x_n) and y=(y1,,yn)\mathbf{y}=(y_1,\ldots,y_n) be two points of Rn\mathbb{R}^n. By algebraic manipulation, we obtain that

ρ(x,y)d(x,y)nρ(x,y).\rho(\mathbf{x},\mathbf{y}) \leqslant d(\mathbf{x},\mathbf{y}) \leqslant \sqrt{n}\rho(\mathbf{x},\mathbf{y}).
note

In other words, the metrics dd and ρ\rho are (topologically) equivalent. In fact, as shown above, these two particular metrics are strongly equivalent, which implies topological equivalence.

The first inequality shows that Bd(x,ε)Bρ(x,ε)B_d(\mathbf{x},\varepsilon) \subseteq B_\rho(\mathbf{x},\varepsilon) whereas the second equality shows that Bρ(x,ε/n)Bd(x,ε)B_\rho(\mathbf{x},\varepsilon/n) \subseteq B_d(\mathbf{x},\varepsilon).

Now, we show that the product topology is the same as that given by the metric ρ\rho. First, let

B=(a1,b1)××(an,bn)B=(a_1,b_1) \times \cdots \times (a_n,b_n)

be a basis element for the product topology, and let x=(x1,,xn)\mathbf{x}=(x_1,\ldots,x_n) be and element of BB. For each ii, there is an εi\varepsilon_i such that

(xiεi,xi+εi)(ai,bi);(x_i-\varepsilon_i, x_i+\varepsilon_i) \subseteq (a_i,b_i);

choose ε=min{ε1,,εn}\varepsilon = \min\{\varepsilon_1,\ldots,\varepsilon_n\}. Then Bρ(x,ε)BB_\rho(\mathbf{x},\varepsilon) \subseteq B, since if yBρ(x,ε)\mathbf{y} \in B_\rho(\mathbf{x},\varepsilon), then the maximum of yixi|y_i-x_i| is less than ε\varepsilon, which implies that xiε<yi<xi+εx_i-\varepsilon < y_i < x_i+\varepsilon for each ii; since εεi\varepsilon \leqslant \varepsilon_i for each ii, it follows that yi(xiεi,xi+εi)(ai,bi)y_i \in (x_i-\varepsilon_i,x_i+\varepsilon_i) \subseteq (a_i,b_i) for each ii, so yB\mathbf{y} \in B. As a result, the ρ\rho-topology is finer than the product topology.

Conversely, let Bρ(x,ε)B_\rho(\mathbf{x},\varepsilon) be a basis element for the ρ\rho-topology. Given the element yBρ(x,ε)\mathbf{y} \in B_\rho(\mathbf{x},\varepsilon), we need to find a basis element BB for the product topology such that yBBρ(x,ε)\mathbf{y} \in B \subseteq B_\rho(\mathbf{x},\varepsilon), and this is immediate: we can just take B=Bρ(x,ε)B=B_\rho(\mathbf{x},\varepsilon), which itself is a basis element for the product topology.

Now we consider the infinite Cartesian product Rω\mathbb{R}^\omega. It is natural to try generalising the metrics dd and ρ\rho to this space in the following ways:

d(x,y)=[i=1(xiyi)2]12(1)d(x,y)=\left[\sum_{i=1}^{\infty}(x_i-y_i)^2\right]^{\frac{1}{2}} \tag{1} ρ(x,y)=supnZ+{xnyn}(2)\rho(x,y)=\sup_{n \in \mathbb{Z}_+}\{|x_n-y_n|\} \tag{2}

but both of the equations above do not always make sense, though equation (1) does define a metric on a certain important subset of Rω\mathbb{R}^\omega, whereas equation (2) makes sense when the metric is replaced by the standard bounded metric.

Definition. Given an index set JJ, and given points x=(xα)αJ\mathbf{x}=(x_\alpha)_{\alpha \in J} and y=(yα)αJ\mathbf{y}=(y_\alpha)_{\alpha \in J} of RJ\mathbb{R}^J, let us define a metric ρˉ\bar{\rho} on RJ\mathbb{R}^J by the equation

ρˉ(x,y)=sup{dˉ(xα,yα):αJ},\bar{\rho}(\mathbf{x},\mathbf{y})=\sup\{\bar{d}(x_\alpha,y_\alpha) : \alpha \in J \},

where dˉ\bar{d} is the standard bounded metric on R\mathbb{R}. One can check that ρˉ\bar{\rho} is indeed a metric; it is called the uniform metric on RJ\mathbb{R}^J, and the topology it induces is called the uniform topology.

[Theorem 20.4] The uniform topology on RJ\mathbb{R}^J is finer than the product topology and coarser than the box topology; these theree topologies are all different if JJ is infinite.

[Theorem 20.5] Let dˉ(a,b)=min{ab,1}\bar{d}(a,b)=\min\{|a-b|,1\} be the standard bounded metric on R\mathbb{R}. If x\mathbf{x} and y\mathbf{y} are two points of Rω\mathbb{R}^\omega, define

D(x,y)=sup{dˉ(xi,yi)i}.D(\mathbf{x},\mathbf{y})=\sup\left\{\frac{\bar{d}(x_i,y_i)}{i}\right\}.

Then DD is a metric that induces the product topology on Rω\mathbb{R}^\omega.

note

Theorem 20.5 shows that if JJ is infinite, then RJ\mathbb{R}^J is metrizable if and only if JJ is countable and RJ\mathbb{R}^J has the product topology.

Now, we consider the relation of the metric topology to other topological concepts:

  • Subspaces or metric spaces behave as expected: if AA is a subspace of the topological space XX and dd is a metric for XX, then the restriction of dd to A×AA \times A is a metric for the topology of AA. A proof of this is shown in Exercise 21.1.
  • There is not much to be said regarding order topologies; some are metrizable, some are not.
  • The Hausdorff axiom is satisfied by every metric topology. Indeed, if xx and yy are distinct points of the metric space (X,d)(X,d), then we let ε=d(x,y)2\varepsilon=\frac{d(x,y)}{2}; the triangle inequality implies that Bd(x,ε)B_d(x,\varepsilon) and Bd(y,ε)B_d(y, \varepsilon) are disjoint.
  • For product topologies, Theorem 20.5 shows that countable products of metrizable spaces are metrizable. A proof of this is given in Exercise 21.3.
  • For continuous functions, the familiar 'ε\varepsilon-δ\delta definition' of continuity carries over to general metric spaces, and so does the sequential criterion. We can also see that we can construct continuous functions through 'arithmetic operations' or by applying uniform limit theorem.

[Theorem 20.1] Let f:XYf : X \to Y; let XX and YY be metrizable with metrics dXd_X and dYd_Y, respectively. Then continuity of ff is equivalent to the requirement that given xXx \in X and given ε>0\varepsilon>0, there exists δ>0\delta>0 such that

dX(x,y)    dY(f(x),f(y))<ε.d_X(x,y) \implies d_Y(f(x),f(y))<\varepsilon.

[Lemma 21.2] (The sequence lemma). Let XX be a topological space; let AXA \subseteq X. If there is a sequence of points of AA converging to xx, then xAx \in \overline{A}; the converse holds if XX is metrizable.

Suppose that xnxx_n \to x, where xnAx_n \in A, then every neighbourhood UU of xx contains a point of AA, so xAx \in \overline{A} by Theorem 17.5.

Conversely, suppose that xx is metrizable and xAx \in \overline{A}. Let dd be a metric for the topology of XX. For each positive integer nn, take the neighbourhood Bd(x,1/n)B_d(x,1/n) centred at xx and of radius 1/n1/n. Choose xnx_n to be a point of its intersection with AA. We assert that the sequence xnx_n converges to xx: note that any open set UU containing xx contains an ε\varepsilon-ball Bd(x,ε)B_d(x,\varepsilon) centred at xx, so if we choose NN so that 1/N<ε1/N<\varepsilon (possible by Archimedean property), then UU contains xix_i for all iNi \geqslant N.

[Theorem 21.3] (Sequential criterion of continuity). Let f:XYf : X \to Y. If the function ff is continuous, then for every convergent sequence xnxx_n \to x in XX, the sequence f(xn)f(x_n) converges to f(x)f(x). The converse holds if XX is metrizable.

Suppose that ff is continuous. Given xnxx_n \to x, we wish to show that f(xn)f(x)f(x_n) \to f(x). Let VV be a neighbourhood of f(x)f(x), then f1(V)f^{-1}(V) is a neighbourhood of xx, so there is an NN such that xnf1(V)x_n \in f^{-1}(V) for nNn \geqslant N, thus f(xn)Vf(x_n) \in V for nNn \geqslant N.

Conversely, assume that the convergent sequence condition is satisfied. Let AA be a subset of XX; we show that f(A)f(A)f(\overline{A}) \subseteq \overline{f(A)} (which is equivalent to continuity by Theorem 18.1). If xAx \in \overline{A}, then there is a sequence xnx_n of points of AA converging to xx (by Lemma 21.2). By assumption, the sequence f(xn)f(x_n) converges to f(x)f(x). Since f(xn)f(A)f(x_n) \in f(A), Lemma 21.2 implies that f(x)f(A)f(x) \in \overline{f(A)} (Note that the metrizability of YY is not needed.) Hence f(A)f(A)f(\overline{A}) \subseteq \overline{f(A)}, as desired.

info

Incidentally, in the proof of Lemma 21.2 and Theorem 21.3, we only required the countable collection Bd(x,1/n)B_d(x,1/n) of balls about xx instead of using the full strength of metrizability. This leads us to make a new definition.

A space XX is said to have a countable basis at the point xx if there is a countable collection {Un}nZ+\{U_n\}_{n \in \mathbb{Z}_+} of neighbourhoods of xx such that any neighbourhood UU of xx contains at least one of the sets UnU_n. A space XX that has a countable basis at each of its points is said to satisfy the first countability axiom.

A metrizable space always satisfies the first countability axiom, but the converse is not true.

[Lemma 21.4] The addition, subtraction and multiplication operations are continuous functions from R×R\mathbb{R} \times \mathbb{R} into R\mathbb{R}; the quotient operation is a continuous function from R×(R\{0})\mathbb{R} \times (\mathbb{R} \backslash \{0\}) into R\mathbb{R}.

A proof of this lemma is a solution to Exercise 21.12.

[Theorem 21.5] ('Arithmetic operations' on continuous functions). If XX is a topological space, and if f,g:XRf,g : X \to \mathbb{R} are continuous functions, then f+gf + g, fgf - g, and fgf \cdot g are continuous. If g(x)0g(x) \neq 0 for all xx, then f/gf/g is continuous.

The map h:XR×Rh : X \to \mathbb{R} \times \mathbb{R} defined by

h(x)=(f(x),g(x))h(x) = (f(x),g(x))

is continuous, by Theorem 18.4 (see below). The function f+gf + g equals the composite of hh and the addition operation +:R×RR+ : \mathbb{R} \times \mathbb{R} \to \mathbb{R}, which we know is continuous by Lemma 21.4, thus f+gf+g is continuous. Similar arguments apply to fgf-g, fgf \cdot g and f/gf/g.

Definition. Let fn:XYf_n : X \to Y be a sequence of functions from the set XX to the metric space YY. Let dd be the metric for YY. We say that the sequence (fn)(f_n) converges uniformly to the function f:XYf : X \to Y if given ε>0\varepsilon > 0, there exists an integer NN such that d(fn(x),f(x))<εd(f_n(x),f(x))<\varepsilon for all n>Nn > N and all xx in XX.

note

Contrast this uniform convergence with pointwise convergence, for which the integer NN for the corresponding ε\varepsilon may vary depending on the choice of xXx \in X.

Exercise 21.6 provides an example of a sequence of functions that is pointwise convergent but not uniformly convergent.

[Theorem 21.6] (Uniform limit theorem). Let fn:XYf_n : X \to Y be a sequence of continuous fucntions from the topological space XX to the metric space YY. If (fn)(f_n) converges uniformly to ff, then ff is continuous.

info

Examples of spaces that are not metrizable:

  • Rω\mathbb{R}^\omega in the box topology.
  • An uncountable product of R\mathbb{R} with itself.

Quotient Topology

Definition. Let XX and YY be topological spaces; let p:XYp : X \to Y be a surjective map. The map pp is said to be a quotient map provided a subset UU of YY is open in YY if and only if p1(U)p^{-1}(U) is open in XX.

note

Note that the implication is two-sided compared to the continuity definition, which indicates that this condition is stronger than continuity. An equivalent condition is that a subset AA of YY is closed in YY if and only if p1(A)p^{-1}(A) is closed in XX; such equivalence follows from the fact that f1(Y\B)=f1(Y)\f1(B)=X\f1(B)f^{-1}(Y \backslash B)= f^{-1}(Y) \backslash f^{-1}(B) = X \backslash f^{-1}(B).

We can also decsribe quotient maps in the following way: We say that a subset CC of XX is saturated (with respect to the surjective map p:XYp : X \to Y) if CC contains every set p1({y})p^{-1}(\{y\}) that it intersects (i.e. for those p1({y})p^{-1}(\{y\}) that has a non-empty intersection with CC, we have that p1(y)p^{-1}(y) is a subset of CC). Hence CC is saturated if it equals the complete inverse image of a subset of YY. Saying that pp is a quotient map is equivalent to saying that pp is continuous and pp maps saturated open sets of XX to open sets of YY (or saturated closed sets of XX to closed sets of YY).

Two special kinds of quotient maps are the open maps and the closed maps.

[A map f:XYf : X \to Y is said to be an open map if for each open set UU of XX, the set f(U)f(U) is open in YY. It is said to be a closed map if for each closed set AA of XX, the set f(A)f(A) is closed in YY.]

It follows from the definition that if p:XYp : X \to Y is a surjective continuous map that is either open or closed, then pp is a quotient map.

There are quotient maps that are neither open nor closed. See Exercise 22.3 for a proof.

Definition. If XX is a space and AA is a set and if p:XAp : X \to A is a surjective map, then there exists exactly one topology T\mathcal{T} on AA relative to which pp is a quotient map; it is called the quotient topology induced by pp, and is defined by letting it consist of those subsets UU of AA such that p1(U)p^{-1}(U) is open in XX.

Definition. Let XX be a topological space, and let XX^* be a partition of XX into disjoint subsets whose union is XX. Let p:XXp : X \to X^* be the surjective map that carries each point of XX to the element of XX^* containing it. In the quotient topology induced by pp, the space XX^* is called a quotient space of XX.

[Theorem 22.1] Let p:XYp : X \to Y be a quotient map; let AA be a subspace of XX that is saturated with respect to pp; let q:Ap(A)q : A \to p(A) be the map obtained by restricting pp.

  1. If AA is either open or closed in XX, then qq is a quotient map.
  2. If pp is either an open map or a closed map, then qq is a quotient map.

[Theorem 22.2] Let p:XYp : X \to Y be a quotient map. Let ZZ be a space and let g:XZg : X \to Z be a map that is constant on each set p1({y})p^{-1}(\{y\}), for yYy \in Y. Then gg induces a map f:YZf : Y \to Z such that fp=gf \circ p = g. THe induced map ff is continuous if and only if gg is continuous; ff is a quotient map if and only if gg is a quotient map.

[Corollary 22.3] Let g:XZg : X \to Z be a surjective continuous map. Let XX^* be the following collection of subsets of XX:

X={g1({z}):zZ}.X^* = \{g^{-1}(\{z\}) : z \in Z\}.

Give XX^* the quotient topology.

(a) The map gg induces a bijective continuous map f:XZf : X^* \to Z, which is a homeomorphism if and only if gg is a quotient map.

Corollary 22.3(a) is also known as the universal property of quotient spaces.

(b) If ZZ is Hausdorff, so is XX^*.

warning

(Example 7).

The product of two quotient maps need not be a quotient map. Here is an example (that is nicer).

Let X=RX = \mathbb{R} and let XX^* be the quotient space obtained from XX by identifying the subset Z+\mathbb{Z}_+ to a point bb; let p:XXp : X \to X^* be the quotient map. Let Q\mathbb{Q} be the subspace of R\mathbb{R} consisting of the rational numbers; let i:QQi : \mathbb{Q} \to \mathbb{Q} be the identity map. We show that

p×i:X×QX×Qp \times i : X \times \mathbb{Q} \to X^* \times \mathbb{Q}

is not a quotient map.

For each nn, let cn=2/nc_n=\sqrt{2}/n, and consider the straight lines in R2\mathbb{R}^2 with gradients 1 and -1, respectively, through the point (n,cn)(n,c_n). Let UnU_n consist of all points X×QX \times \mathbb{Q} that lie above both of these lines or beneath both of them, and also between the vertical lines x=n14x=n-\frac{1}{4} and x=n+14x= n+\frac{1}{4}. Then UnU_n is open in X×QX \times \mathbb{Q}, it contains the set {n}×Q\{n\} \times \mathbb{Q} because cnc_n is not rational.

Let UU be the union of the sets UnU_n, then UU is open in X×QX \times \mathbb{Q}. It is saturated with respect to p×ip \times i because it contains the entire set Z+×{q}\mathbb{Z}_+ \times \{q\} for each qQq \in \mathbb{Q}. We assume that U=(p×i)(U)U'=(p \times i)(U) is open in X×QX^* \times \mathbb{Q} and reach a contradiction.

Since UU contains, in particular, the set Z+×{0}\mathbb{Z}_+ \times \{0\}, the set UU' contains the point (b,0)(b,0). Hence UU' contains an open set of the form W×IδW \times I_\delta, where WW is a neighbourhood of bb in XX^* and Iδ:={yQ:y<δ}I_\delta := \{y \in \mathbb{Q} : |y|<\delta\}. It follows that

p1(W)×IδU.p^{-1}(W) \times I_\delta \subseteq U.

Choose nn large enough so that cn<δc_n < \delta, then since p1(W)p^{-1}(W) is open in XX and contains Z+\mathbb{Z}_+, we can choose ε<14\varepsilon < \frac{1}{4} so that the interval (nε,n+ε)(n-\varepsilon,n+\varepsilon) is contained in p1(W)p^{-1}(W), then UU contains the subset V=(nε,n+ε)×IδV = (n-\varepsilon,n+\varepsilon) \times I_\delta of X×QX \times \mathbb{Q}. However, there are many points (x,y)(x,y) of VV that do not lie in UU (e.g. the points (x,y)(x,y) where x=n+12εx = n+\frac{1}{2}\varepsilon and yy is a rational number with ycn<12ε|y-c_n|<\frac{1}{2}\varepsilon.)!

Closed Sets and Limit Points

Closed Sets, Interior, Closure, Neighbourhood

Definition. A subset AA of a topological space XX is said to be closed if the set X\AX \backslash A** is open.

[Theorem 17.1] Let XX be a topological space, then the following hold:

  1. \emptyset and XX are closed.
  2. Arbitrary intersections of closed sets are closed.
  3. Finite unions of closed sets are closed.

The proof for 1. follows from the (topological) definition of openness and the facts that =X\X\emptyset = X \backslash X and X=X\X = X \backslash \emptyset. The proofs for 2. and 3. follow from the definition of a closed set and de Morgan's law.

[Theorem 17.2] Let YY be a subspace of XX, then a set AA is closed in YY if and only if it is equal to the intersection of a closed set of XX with YY.

Suppose that A=CYA = C \cap Y, where CC is closed in XX. It follows that X\CX \backslash C is open, then by definition of the subspace topology, (X\C)Y(X \backslash C) \cap Y is open. Observe that (X\C)Y=Y\A(X \backslash C) \cap Y = Y \backslash A, so Y\AY \backslash A is open and thus AA is closed in YY by definition.

Conversely, suppose that AA is closed in YY, then Y\AY \backslash A is open in YY, so by definition of subspace topology again it is equal to YUY \cap U where UU is open in XX, which means that X\UX \backslash U is closed in XX. Note that A=Y(X\U)A = Y \cap (X \backslash U), so AA is equal to the intersection of a closed set of XX with YY.

[Theorem 17.3] Let YY be a subspace of XX. If AA is closed in YY and YY is closed in XX, then AA is closed in XX.

A proof of this is a solution to Exercise 17.2.

Definition. Let AA be a subset of a topological space XX. The interior of AA, denoted by intA\mathrm{int} A, is defined as the union of all open sets contained in AA. The closure of AA, denoted by A\overline{A}, is defined as the intersection of all closed sets containing AA. In fact, we have the following relationship:

intAAA.\mathrm{int} A \subseteq A \subseteq \overline{A}.

Another important fact is that AA is open if and only if A=intAA=\mathrm{int} A, whereas AA is closed if and only if A=AA=\overline{A}.

tip

Informally, one can imagine that intA\mathrm{int} A is constructed 'bottom-up from inside AA' whereas A\overline{A} is constructed 'top-down from outside AA'.

[Theorem 17.4] Let YY be a subspace of XX; let AA be a subset of YY; let A\overline{A} denote the closure of AA in XX. Then the closure of AA in YY equals AY\overline{A} \cap Y.

[Theorem 17.5] Let AA be a subset of the topological space XX.

a. Then xAx \in \overline{A} if and only if every open set UU containing xx intersects AA (i.e. the intersection is non-empty).

b. Supposing the topology of XX is given by a basis, then xAx \in \overline{A} if and only if every basis element BB containing xx intersects AA.

For (a), it suffices to prove the equivalent statement below:

xA    there exists an open set U containing x that does not intersect A.x \notin \overline{A} \iff \text{there exists an open set } U \text{ containing } x \text{ that does not intersect } A.

If xx is not in A\overline{A}, then the set U=X\AU = X \backslash \overline{A} is an open set UU containing xx that does not intersect AA, as desired. Conversely, if there exists an open set UU containing xx which does not intersect AA, then X\UX \backslash U is a closed set containing AA. By the definition of the closure A\overline{A}, the set X\UX \backslash U must contain A\overline{A} since it is closed. Therefore, xx cannot be in A\overline{A}.

As for (b), it immediately follows from (a) due to the facts that every basis element BB containing xx is an open set and that every open set UU containing xx contains a basis element that contains xx.

Note: we say that UU is a neighbourhood of xx if(f) UU is an open set containing xx.

Limit Point

Definition. Let AA be a subset of a topological space XX and xx be a point of XX. We say that xx is a limit point (or 'cluster point', or 'accumulation point') of AA if every neighbourhood of xx intersects AA in some point other than xx itself. This can be rephrased as: if xx is a limit point, then V(x)(A\{x})V(x) \cap (A \backslash \{x\}) \neq \emptyset, where V(x)V(x) denotes an (arbitrary) neighbourhood of xx, or that xx belongs to the closure of A\{x}A \backslash \{x\}.

note

If a topological space is endowed with a metric, then the neighbourhood notation can be written as Vδ(x)V_\delta(x), where δ>0\delta>0 is a (positive) real number, and the arbitrariness of a neighbourhood of xx is represented by the arbitrariness of the values of δ\delta.

This is in fact the definition often employed in a calculus or real analysis course.

[Theorem 17.6] Let AA be a subset of a topological space XX, then the closure of AA is equal to the union of AA and all of its limit (accumulation) points, i.e. if we denote the set of all limit points of AA by AA', then A=AA\overline{A} = A \cup A'.

If xx is in AA', then every neighbourhood of xx intersects AA (in a point different from xx). Thus, by Theorem 17.5, xx belongs to A\overline{A}. Hence AAA' \subseteq \overline{A}. Since by definition AAA \subseteq \overline{A}, it follows that AAAA \cup A' \subseteq \overline{A}.

To demonstrate the reverse inclusion, we let xx be a point of A\overline{A} and show that xAAx \in A \cup A'. If xx happens to lie in AA, then xAAx \in A \cup A'. If xx does not lie in AA, since xAx \in \overline{A}, every neighbourhood UU of xx intersects AA (Theorem 17.5); as xAx \notin A, the set UU must intersect AA in a point different from xx, so xAx \in A' and thus xAAx \in A \cup A', as desired.

An immediately corollary [Corollary 17.7] from here is that: a subset of a topological space is closed if and only if it contains all limit points. This follows from the fact that AA is closed if and only if A=AA=\overline{A}.

Hausdorff space

Motivation

In an arbitrary topological space, a sequence xnnN\langle x_n \rangle_{n \in \mathbb{N}} of points of the space XX converges to the point xx of XX if, corresponding to each neighbourhood UU of xx, there exists a postive integer NN such that xnUx_n \in U for all nNn \geqslant N.

warning

Notice the lack of a metric and ε>0\varepsilon>0 in the definition of convergence of a sequence in an arbitrary topological space. This implies that unlike in R\mathbb{R} and R2\mathbb{R}^2, there may be multiple points of convergence for a single sequence.

In addition, for arbitrary topological spaces, a one-point set may not be closed, unlike the spaces R\mathbb{R} and R2\mathbb{R}^2. The one-point set {b}\{b\} is not closed in the three-point set S={a,b,c}S = \{a,b,c\} equipped with the topology S={,S,{a,b},{b},{b,c}}\mathcal{S} = \{\emptyset, S, \{a,b\},\{b\},\{b,c\}\}.

These relatively counterintuitive phenomena lead to the notion of Hausdorff spaces, as mathematicians seek to rule them out in the study of topology.

Main Notes

Definition. A topological space XX is called a Hausdorff space (or, 'is Hausdorff') if for each pair x1,x2Xx_1, x_2 \in X such that they are distinct points, i.e. x1x2x_1 \neq x_2, there exist neighbourhoods U1U_1 of x1x_1 and U2U_2 of x2x_2 such that they are disjoint.

[Theorem 17.8] Every finite point set in a Hausdorff space XX is closed.

As the closedness of a set is preserved under finite unions, it suffices to show that every one-point set {x0}\{x_0\} is closed. If xXx \in X and xx0x \neq x_0, then xx and x0x_0 have disjoint neighbourhoods UU and VV by the definition of a Hausdorff space. Since UU does not intersect {x0}\{x_0\}, the point xx cannot belong to the closure of the set {x0}\{x_0\}. Since the choice of xx here is arbitrary, it follows that the closure of the set {x0}\{x_0\} is {x0}\{x_0\} itself, implying that it is closed.

note

The condition for finite point sets to be closed is in fact weaker than the Hausdorff condition. For example, R\mathbb{R} in the finite complement topology is not a Hausdorff space, but it is a space in which finite point sets are closed.

There is a name for the condition that the finite point sets be closed: the T1T_1 axiom. (The reason of this strange terminology will be explained in Chapter 4.)

(Note: R\mathbb{R} in the finite complement topology satisfying the T1T_1 axiom follows immediately from the definition of the topology itself.)

Why is R\mathbb{R} in the finite complement topology not a Hausdorff space?

Suppose that x1,x2Rx_1, x_2 \in \mathbb{R} where x1x2x_1 \neq x_2. Assume that there exists a neighbourhood UU of x1x_1 and VV of x2x_2 such that UV=U \cap V = \emptyset. This implies that R\(UV)=(R\U)(R\V)=R\mathbb{R} \backslash (U \cap V) = (\mathbb{R} \backslash U) \cup (\mathbb{R} \backslash V) = \mathbb{R}, which is impossible because both R\U\mathbb{R} \backslash U and R\V\mathbb{R} \backslash V are supposed to be finite by the definition of the finite complement topology.

[Theorem 17.9] Let XX be a space satisfying the T1T_1 axiom and AA be a subset of XX, then the point xx is a limit point of AA if and only if every neighbourhood of xx contains infinitely many points of AA.

If every neighbourhood of xx intersects AA in infinitely many points, it certainly intersects AA in some points other than xx itself, so xx is a limit point of AA.

Conversely, suppose that xx is a limit point of AA, and assume that some neighbourhood UU of xx intersects AA in only finitely many points, then UU also intersects A\{x}A \backslash \{x\} in finitely many points, say, x1,,xmx_1,\ldots,x_m. Since {x1,,xm}\{x_1,\ldots,x_m\} is a finite point set, it is closed under the T1T_1 axiom, so X\{x1,,xm}X \backslash \{x_1,\ldots,x_m\} is open. This means that

U(X\{x1,.xm})U \cap (X \backslash \{x_1,\ldots.x_m\})

is a neighbourhood of xx that does not intersect A\{x}A \backslash \{x\} at all, contradicting the assumption that xx is a limit point of AA.

As remarked above, the Hausdorff axiom eliminates the possibility of a sequence in a topological space XX converging to more than one point in XX.

[Theorem 17.10] If XX is a Hausdorff space, then a sequence of points of XX converges to at most one point of XX.

Suppose that xnx_n is a sequence of points of XX that converges to xx, then there is a positve integer NN that depends on each neighbourhood UU of xx such that xnUx_n \in U whenever nNn \geqslant N. If yxy \neq x, let UU and VV be disjoint neighbourhoods of xx and yy, respectively. Since UU contains xnx_n for all but finitely many values of nn, the set VV cannot. Therefore, xnx_n cannot converge to yy.

In view of Theorem 17.10, if the sequence xnx_n of points of the Hausdorff space XX converges to the point xx of XX, we can write xnxx_n \to x, and say that xx is the limit of the sequence xnx_n.

[Theorem 17.11] Every simply ordered set is a Hausdorff space in the order topology. The product of two Hausdorff spaces is a Hausdorff space. A subspace of a Hausdorff space is a Hausdorff space.

The proofs of the following results are solutions to Exercises 17.10, 17.11 and 17.12 respectively.

Continuous Functions

Let us consider continuous functions beyond real and complex numbers, by generalising them in any topological space.

Definition. Let XX and YY be topological spaces. A function f:XYf : X \to Y is said to be continuous if for each open subset VV of YY, the set f1(V)f^{-1}(V) is an open subset of XX.

Note that if VV does not intersect the image set f(X)f(X) of ff, then f1(V)f^{-1}(V) is empty.

note

This definition is in fact the Global Continuity Theorem, which is considered a consequence if the continuity of a function is defined using the notion of metric spaces (a.k.a. the 'ε\varepsilon-δ\delta definition').

In fact, in real-valued functions of a real variable, i.e. functions in the form of f:RRf : \mathbb{R} \to \mathbb{R}, the ε\varepsilon-δ\delta definition and the definition above are equivalent. To prove that the open set definition implies the ε\varepsilon-δ\delta definition, it suffices to show that for each open ε\varepsilon-neighbourhood VV of f(x0)Rf(x_0) \in \mathbb{R}, one can always find a suitable δ\delta-neighbourhood UU of x0Rx_0 \in \mathbb{R} such that f(U)Vf(U) \subseteq V. As for the converse, see Exercise 18.1.

Detailed proof that the open set definition implies the ε\varepsilon-δ\delta definition.

Let f:RRf : \mathbb{R} \to \mathbb{R} be a continuous function. For any x0x_0 in Rn\mathbb{R}^n and ε>0\varepsilon > 0, the interval V=f(x0)ε,f(x0)+εV=f(x_0)-\varepsilon, f(x_0)+\varepsilon is a neighbourhood of f(x0)f(x_0) in the codomain R\mathbb{R}. Thus, f1(V)f^{-1}(V) is an open set in the domain R\mathbb{R}. As f1(V)f^{-1}(V) contains the point x0x_0, it contains some basis element (a,b)(a,b) about x0x_0.

Choose δ\delta to be the smaller of the two numbers x0ax_0-a and bx0b-x_0, then if xx0<δ|x-x_0|<\delta, the point xx must be in (a,b)(a,b), so that f(x)Vf(x) \in V, and thus f(x)f(x0)<ε|f(x)-f(x_0)|< \varepsilon, as desired.

note

The arguments above can be modified to prove the same for real multivariate functions, though this may require the notion of box topology or (general) product topology to define the standard topology on Rn\mathbb{R}^n. Nevertheless, it is shown that the topology on Rn\mathbb{R}^n generated by (open) hyperrectangles and nn-balls are the same.

However, this ε\varepsilon-δ\delta definition, as well as the notion of sequential continuity do not generalise to arbitrary topological spaces, but they serve important roles in the study of metric spaces.

If the topology of the range space (codomain) YY is generated by a basis B\mathcal{B}, then to prove continuity it suffices to check that the inverse image of each basis element is open. Why?

Any open set VV of YY can be written as a union of basis elements, i.e.

V=αJBα.V = \bigcup_{\alpha \in J}B_\alpha.

Thus, since inverse images preserve unions, we have

f1(V)=αJf1(Bα).f^{-1}(V) = \bigcup_{\alpha \in J}f^{-1}(B_\alpha).

We can then see that f1(V)f^{-1}(V) is open if each set f1(Bα)f^{-1}(B_\alpha) is open.

note

The same applies for a subbasis, as an arbitrary basis element BB for YY can be expressed as a finite intersection S1SnS_1 \cap \cdots \cap S_n of subbasis elements. Since f1(B)=i=1nf1(Si)f^{-1}(B) = \bigcap_{i=1}^{n}f^{-1}(S_i), the inverse image of every basis element is open if the subbasis element is open.

[Theorem 18.1] Let XX and YY be topological spaces; let f:XYf : X \to Y, then the following are equivalent:

  1. ff is continuous.
  2. For every subset AA of XX, we have f(A)f(A)f(\overline{A}) \subseteq \overline{f(A)}.
  3. For every closed set BB of YY, the set f1(B)f^{-1}(B) is closed in XX.
  4. For each xXx \in X and each neighbourhood VV of f(x)f(x), there is a neighbourhood UU of xx such that f(U)Vf(U) \subseteq V.

If the condition 4. holds for the point xx of XX, we say that ff is continuous at the point xx.

(1)     \implies (2). Assume that ff is continuous. Let AA be a subset of XX. We show that if xAx \in \overline{A}, then f(x)f(A)f(x) \in \overline{f(A)}. Since ff is continuous, if VV is a neighbourhood of f(x)f(x), then f1(V)f^{-1}(V) is an open set of XX containing xx, so it must intersect AA in some point yy by Theorem 17.5. It follows that VV intersects f(A)f(A) in the point f(y)f(y), so that f(x)f(A)f(x) \in \overline{f(A)} as desired.

Homeomorphisms

Definition. Let XX and YY be topological spaces and f:XYf : X \to Y be a bijection. If both the function ff and the inverse function f1:YXf^{-1} : Y \to X are continuous, then ff is called a homeomorphism.

Homeomorphisms are remarkable in the sense that they act like isomorphisms in abstract algebra: they give bijective correspondences between the collections of open sets that preserve the topological structure between two spaces, as well as topological properties (i.e. properties that are entirely expressed in terms of open sets).

If f:XYf : X \to Y is an injective continuous map, ZZ is the image set f(X)f(X), considered as a subspace of YY, and f:XZf' : X \to Z is a homeomorphism, we say that f:XYf : X \to Y is a (topological) imbedding (or 'embedding') of XX in YY.

[Theorem 18.2] (Rules for constructing continuous functions). Let XX, YY and ZZ be topological spaces.

  1. (Constant function) If f:XYf : X \to Y maps all of XX into the single point y0y_0 of YY, then ff is continuous.
  2. (Inclusion) If AA is a subspace of XX, the inclusion function j:AXj : A \to X is continuous.
  3. (Composites) If f:XYf : X \to Y and g:YZg : Y \to Z are continuous, then the map gf:XZg \circ f : X \to Z is continuous.
  4. (Domain restriction) If f:XYf : X \to Y is continuous, and if AA is a subspace of XX, then the restricted function fA:AYf|_A : A \to Y is continuous.
  5. (Range restriction or expansion) Let f:XYf : X \to Y be continuous. If ZZ is a subspace of YY containing the image set f(X)f(X), then the function g:XZg : X \to Z obtained by restricting the range of ff is continuous. If ZZ is a space with YY being a subspace, then the function h:XZh : X \to Z obtained by expanding the range of ff is continuous.
  6. (Local formulation of continuity) If XX can be written as the union of open sets UαU_\alpha such that fUαf|_{U_\alpha} is continuous for each α\alpha, then the map f:XYf : X \to Y is continuous.
note

In analysis, taking sums, difference, products or quotients of continuous real-valued functions yields us a continuous function as well. This can be generalised to functions with domains being an arbitrary topological space XX, but the range still being R\mathbb{R}.

[Theorem 18.3] (Pasting Lemma). Let X=ABX = A \cup B, where AA and BB are both closed in XX. Let f:AYf : A \to Y and g:BYg : B \to Y be continuous. If f(x)=g(x)f(x)=g(x) for every xABx \in A \cap B, then ff and gg combine to give a continuous function h:XYh : X \to Y, defined by h(x)={f(x)if xA,g(x)if xBh(x)=\begin{cases} f(x) & \text{if } x \in A, \\ g(x) & \text{if } x \in B \end{cases}.

(Note that the requirement that f(x)=g(x)f(x)=g(x) for every xABx \in A \cap B is so that h:XYh : X \to Y is well-defined as a function.)

If AA and BB are closed in XX, consider a closed subset of YY, denoted here by CC. Note that

h1(C)=f1(C)g1(C).h^{-1}(C) = f^{-1}(C) \cup g^{-1}(C).

Since ff is continuous, f1(C)f^{-1}(C) is closed in AA (by Theorem 18.1), and thus closed in XX. Similarly, g1(C)g^{-1}(C) is closed in BB and hence closed in XX. Their (finite) union h1(C)h^{-1}(C) is thus closed in XX, so hh is continuous.

note

If AA and BB are open in XX, it immediately follows from the 'local formulation of continuity' rule in Theorem 18.2.

[Theorem 18.4] (Maps into products). Let f:AX×Yf : A \to X \times Y be given by the equation f(a)=(f1(a),f2(a))f(a)=(f_1(a), f_2(a)), then ff is continuous if and only if the functions f1:AXf_1 : A \to X and f2:AYf_2 : A \to Y are continuous.

(The maps f1f_1 and f2f_2 are called the coordinate functions of ff.)

note

There is no useful criterion for the continuity of maps like f:A×BXf : A \times B \to X with the product space being the domain, one might guess that ff is continuous if it is continuous 'in each variable separately,' but this conjecture is not true. See Exercise 12 for a counterexample proof.

Footnotes

  1. A generalisation of the subspace topology (as well as product topology) is the initial topology, which confusingly is also called the induced topology. Hence, the name 'induced topology' will hereon be avoided so that ambiguity does not arise.