Skip to main content

Exercises Attempts

Exercise after §13

  1. Let XX be a topological space; let AA be a subset of XX. Suppose that for each xAx \in A there is an open set UU containing xx such that UAU \subseteq A. Show that AA is open in XX.

Since for any xAx \in A, there exists an open set UxU_x such that xUxAx \in U_x \subseteq A, it follows that AxAUxA    A=xAUxA \subseteq \bigcup_{x \in A}U_x \subseteq A \implies A=\bigcup_{x \in A}U_x, so by the definition of a topology, AA is open in XX.

3 (reworded). Show that the countable complement topology Tc\mathcal{T}_c is indeed a topology on the set XX.

Is the collection

T:={U:X\U is infinite or empty or all of X}\mathcal{T}_\infty := \{U : X \backslash U \text{ is infinite or empty or all of } X\}

a topology on XX?

Tc\mathcal{T}_c case.

Indeed, XX and \emptyset are in Tc\mathcal{T}_c because X=XX = X-\emptyset and =XX\emptyset=X-X, and \emptyset is countable.

Consider an indexed family of sets {Uα}\{U_\alpha\} in Tc\mathcal{T}_c, then the (relative) component of its union Uα\bigcup U_\alpha with respect to XX has the following equivalence (due to de Morgan's law):

X\Uα=(X\Uα).X \backslash \bigcup U_\alpha = \bigcap \left(X \backslash U_\alpha \right).

Note that since each UαU_\alpha is in Tc\mathcal{T}_c, each X\UαX \backslash U_\alpha is countable. Due to the fact that an arbitrary intersection of countable sets is countable, X\UαX \backslash \bigcup U_\alpha is thus countable, so Uα\bigcup U_\alpha is in Tc\mathcal{T}_c.

Now, consider a finite family of sets {Ui}i=1,,n\{U_i\}_{i=1,\ldots,n} in Tc\mathcal{T}_c, then by a similar argument, the (relative) component of its intersection i=1nUi\bigcap_{i=1}^{n}U_i with respect to XX has the following equivalence:

X\i=1nUi=i=1n(X\Ui).X \backslash \bigcap_{i=1}^{n}U_i=\bigcup_{i=1}^{n}(X \backslash U_i).

Similarly, each X\UiX \backslash U_i is countable. Since a finite union of countable sets is countable1, X\i=1nUiX \backslash \bigcap_{i=1}^{n}U_i is thus countable, so i=1nUi\bigcap_{i=1}^{n}U_i is in Tc\mathcal{T}_c.

Therefore, by the definition of a topology, Tc\mathcal{T}_c is indeed a topology.

T\mathcal{T}_\infty case.

By the similar reason as above, XX and \emptyset are in T\mathcal{T}_\infty.

Since an arbitrary, in particular finite, union of infinite sets is infinite, we can see from a similar argument as above that a finite intersection of sets in T\mathcal{T}_\infty is in T\mathcal{T}_\infty.

However, an arbitrary intersection of infinite sets may not be infinite nor empty. A relatively straightforward example is when X=RX = \mathbb{R} and a family of sets UαXU_\alpha \subseteq X is defined as sets in the form of {xR:x>r}\{ x \in \mathbb{R} : |x| > r \} where rr is an arbitrary positive real number. Note that each X\UαX \backslash U_\alpha, which is given by {xR:xr}\{ x \in \mathbb{R} : |x| \leqslant r \}, is infinite, so each UαU_\alpha is in T\mathcal{T}_\infty and thus open.

It follows that the intersection of this family of sets is {0}\{0\}, which is finite and non-empty. This fact causes Uα\bigcup U_\alpha, i.e. an arbitrary union of sets in T\mathcal{T}_\infty, not to be in T\mathcal{T}_\infty, due to the similar argument as mentioned in the Tc\mathcal{T}_c case.

This is enough to conclude that T\mathcal{T}_\infty is not a topology.

  1. (a) If {Tα}\{\mathcal{T}_\alpha \} is a family of topologies on XX, show that Tα\bigcap\mathcal{T}_\alpha is a topology on XX. Is Tα\bigcup\mathcal{T}_\alpha a topology on XX?

(b) Let {Tα}\{\mathcal{T}_\alpha \} be a family of topologies on XX. Show that there is a unique smallest topology on XX containing all the collections Tα\mathcal{T}_\alpha, and a unique largest topology contained in all Tα\mathcal{T}_\alpha.

(c) If X={a,b,c}X = \{a,b,c\}, let

T1={,X,{a},{a,b}}andT2={,X,{a},{b,c}}.\mathcal{T}_1 = \{\emptyset, X, \{a\}, \{a,b\}\} \quad \text{and} \quad \mathcal{T}_2=\{\emptyset, X, \{a\},\{b,c\}\}.

Find the smallest topology containing T1\mathcal{T}_1 and T2\mathcal{T}_2, and the largest topology contained in T1\mathcal{T}_1 and T2\mathcal{T}_2.

  1. (a) Indeed, \emptyset and XX are in the intersection as all topologies are required to contain them. For any set that we take from the intersection, they belong to all the topologies that form the intersection, so any arbitrary union and finite intersection of them also belong to all the topologies that form the intersection and thus are in the intersection, so the intersection is a topology.

    The union operation does not preserve topologies in general. Consider X={a,b,c}X=\{a,b,c\} and two of its topologies T1={,X,{a},{b,c}}\mathcal{T}_1=\{\emptyset,X,\{a\},\{b,c\}\} and T2={,X,{b}}\mathcal{T}_2=\{\emptyset,X,\{b\}\}, then T1T2={,X,{a},{b},{b,c}}\mathcal{T}_1 \cup \mathcal{T}_2 = \{\emptyset,X,\{a\},\{b\},\{b,c\}\} and we can see that it is not a topology since {a}{b}={a,b}\{a\} \cup \{b\}=\{a,b\} is not in T1T2\mathcal{T}_1 \cup \mathcal{T}_2.

(b) We know from (a) that Tα\bigcap \mathcal{T}_\alpha is a topology on XX contained in all Tα\mathcal{T}_\alpha. We claim that it is the unique largest such topology. It is the largest because a topology that is contained in all Tα\mathcal{T}_\alpha must contain Tα\bigcap \mathcal{T}_\alpha, and if any xTαx \notin \bigcap \mathcal{T}_\alpha is added in, there must be some Tα\mathcal{T}_\alpha that does not contain xx. The uniqueness comes from the fact that if T1\mathcal{T}_1 and T2\mathcal{T}_2 are the largest topologies, then T1T2\mathcal{T}_1 \subseteq \mathcal{T}_2 and T2T1\mathcal{T}_2 \subseteq \mathcal{T}_1 , so T1=T2\mathcal{T}_1=\mathcal{T}_2.

(c) The smallest topology containing T1\mathcal{T}_1 and T2\mathcal{T}_2 is given by {,X,{a},{b},{a,b},{b,c}}\{\emptyset,X,\{a\},\{b\},\{a,b\},\{b,c\}\}. The largest topology contained in T1\mathcal{T}_1 and T2\mathcal{T}_2 is given by {,X,{a}}\{\emptyset, X , \{a\}\}.

  1. Show that if A\mathcal{A} is a basis for a topology on XX, then the topology generated by A\mathcal{A} equals the intersection of all topologies on XX that contain A\mathcal{A}. Prove the same if A\mathcal{A} is a subbasis.
note

Exercise 13.5 also means that the topology generated by a basis A\mathcal{A} is the 'smallest' topology of a set XX that contains A\mathcal{A}.

Basis case.

Suppose that UTAU \in \mathcal{T}_\mathcal{A}, then by Lemma 13.1, UU is a union of some elements of A\mathcal{A}, i.e. U=CU = \bigcup\mathcal{C} for some CA\mathcal{C} \subseteq \mathcal{A}. Since CAATT\mathcal{C} \subseteq \mathcal{A} \subseteq \bigcap_{\mathcal{A} \subseteq \mathcal{T}}\mathcal{T}, ATT\bigcap_{\mathcal{A} \subseteq \mathcal{T}}\mathcal{T} is a topology as previously shown in 4(a), and any arbitrary union of open sets is open, this implies that UATTU \in \bigcap_{\mathcal{A} \subseteq \mathcal{T}}\mathcal{T}.

Conversely, suppose that UATTU' \in \bigcap_{\mathcal{A} \subseteq \mathcal{T}}\mathcal{T}. It then follows immediately that UTAU' \in \mathcal{T}_\mathcal{A}, because TA\mathcal{T}_\mathcal{A} is one of the topologies that contain A\mathcal{A}, as we know that each basis element is in TA\mathcal{T}_\mathcal{A} and any arbitrary union of open sets is open, which results in the union of the singletons of each basis element, which is exactly A\mathcal{A} itself, to be in TA\mathcal{T}_\mathcal{A} as well.

Therefore, TA=ATT\mathcal{T}_\mathcal{A} = \bigcap_{\mathcal{A} \subseteq \mathcal{T}}\mathcal{T}.

Subbasis case.

If A\mathcal{A} is a subbasis, then there exists a corresponding basis A\mathcal{A}' in XX such that each of its basis elements is a finite intersection of subbasis elements in A\mathcal{A}, and the resulting topology generated respectively by A\mathcal{A} and A\mathcal{A}' are equal.

Since a finite intersection of elements of a topology is still in the topology, for all topologies on XX that contain A\mathcal{A} (which is a subbasis this time), they also contain A\mathcal{A}'. This allows us to revert back to the basis case, and argue similarly to reach the desired conclusion.

Reference.

  1. Show that the topologies of Rl\mathbb{R}_l and RK\mathbb{R}_K are not comparable.

Note that there is an abuse of notation to denote the topologies, instead of the corresponding sets, as Rl\mathbb{R}_l and RK\mathbb{R}_K respectively.

It suffices to find a set (element) in Rl\mathbb{R}_l that is not in RK\mathbb{R}_K and vice versa. We recall that K:={1n:nZ+}K := \left\{\frac{1}{n} : n \in \mathbb{Z}_+ \right\} here.

Take [0,1)Rl[0,1) \in \mathbb{R}_l and (0,1)\KRK(0,1) \backslash K \in \mathbb{R}_K. Since K[0,1)K \subset [0,1), [0,1)RK[0,1) \notin \mathbb{R}_K. We also have that (0,1)\KRl(0,1) \backslash K \notin \mathbb{R}_l because any arbitrary union and finite intersection of elements (sets) in Rl\mathbb{R}_l (which are in the form of [a,b)[a,b) with a,bRa,b \in \mathbb{R}) that contains (0,1)\K(0,1) \backslash K will contain a non-empty subset of KK.

  1. (a) Apply Lemma 13.2 to show that the countable collection
B={(a,b):a<b,  a and b rational}\mathcal{B} = \{ (a,b) : a<b,\; a \text{ and } b \text{ rational}\}

is a basis that generates the standard topology on R\mathbb{R}.

(b) Show that the collection

C={[a,b):a<b,  a and b rational}\mathcal{C} = \{[a,b) : a<b,\; a \text{ and } b \text{ rational}\}

is a basis that generates a topology different from the lower limit topology on R\mathbb{R}.

(a) Let UU be an open set in the standard topology on R\mathbb{R}.

We know that rational numbers are dense in real numbers, i.e. for any x,yRx,y \in \mathbb{R} with x<yx<y, there exists a rational number rr such that x<r<yx<r<y. Let uUu \in U, then uu is contained in some open interval (a,b)U(a,b) \subseteq U where a,bRa,b \in \mathbb{R} and a<ba < b, so that a<u<ba<u<b. By the density of rational numbers, there exists rational numbers r1r_1 and r2r_2 such that a<r1<ua < r_1 < u and u<r2<bu < r_2 < b.

This implies that r1<u<r2r_1 < u < r_2, so uu is contained in the open interval (r1,r2)(r_1,r_2), which is in B\mathcal{B} as well as contained in the interval (a,b)(a,b), and thus in UU. Therefore, by Lemma 13.2, B\mathcal{B} is a basis that generates the standard topology.

(b) We first show that C\mathcal{C} satisfies the two conditions for a basis.

For the first condition, let xRx \in \mathbb{R} be arbitrary. If xx is rational, then we see that xx is contained in [x,y)[x,y), where yy is rational y>xy > x, and the interval is in C\mathcal{C}. If xx is irrational, then the density of irrational numbers in the real numbers, and thus rational numbers in particular, gives rise to the existence of an interval [a,b)[a',b') in C\mathcal{C} that contains xx.

For the second condition, let [c,d)[c,d) and [c,d)[c',d') be in C\mathcal{C}, then their intersection is given by [max{c,c},min{d,d})[\max\{c,c'\},\min\{d,d'\}), which is in C\mathcal{C}, so the second condition is satisfied.

Consider the set U:=[2,3)RU := [\sqrt{2},3) \subseteq \mathbb{R}. It is open in the lower limit topology on R\mathbb{R} but it is not open in the topology generated by C\mathcal{C} as there is no interval CC in C\mathcal{C} such that 2C\sqrt{2} \in C and CUC \subseteq U, as any interval in C\mathcal{C} that contains 2\sqrt{2} must contain some numbers strictly less than 2\sqrt{2}. Thus, the topology generated by C\mathcal{C} is different from the lower limit topology.

Exercise after §16

  1. Show that if YY is a subspace of XX, and AA is a subset of YY, then the topology AA inherits as a subspace of YY is the same as the topology it inherits as a subspace of XX.

Suppose that XX is a set endowed with a topology T\mathcal{T}, YY is a subspace of XX endowed with the topology T:={YU:UT}\mathcal{T}' := \{ Y \cap U : U \in \mathcal{T}\}, and AYA \subseteq Y.

Let TY:={AV:VT}\mathcal{T}_Y := \{A \cap V : V \in \mathcal{T}' \} be the topology AA inherits as a subspace of YY and TX:={AW:WT}\mathcal{T}_X := \{ A \cap W : W \in \mathcal{T} \} as the topology AA inherits as a subspace of XX.

Suppose that a set SS lies in TY\mathcal{T}_Y, then S=AVS = A \cap V where VV is in T\mathcal{T}'. This also means that S=A(YW)S = A \cap (Y \cap W) where WW is in T\mathcal{T}. Now, note that A(YW)=(AY)WA \cap (Y \cap W) = (A \cap Y) \cap W and since AA is a subset of YY, AYA \cap Y is just AA. Therefore, S=AWS = A \cap W where WW is in T\mathcal{T}. Thus, TYTX\mathcal{T}_Y \subseteq \mathcal{T}_X.

Since it is valid to reverse the arguments above, we have that TXTY\mathcal{T}_X \subseteq \mathcal{T}_Y as well. Thus, TY=TX\mathcal{T}_Y = \mathcal{T}_X, so the topology AA inherits as a subspace of YY is the same as the topology it inherits as a subspace of XX.

  1. If T\mathcal{T} and T\mathcal{T}' are topologies on XX and T \mathcal{T}' is strictly finer than T\mathcal{T}, what can you say about the corresponding subspace topologies on the subset YY of XX?

We can say that the subspace topologies on the subset YY of XX inherited from T\mathcal{T}' is strictly finer than that inherited from T\mathcal{T}.

To see this, let's look into the definitions of the respective subspace topologies inherited from T\mathcal{T} and T\mathcal{T}'; we denote them by TY:={YU:UT}\mathcal{T}_Y := \{ Y \cap U : U \in \mathcal{T} \} and TY:={YV:VT}\mathcal{T}'_Y := \{ Y \cap V : V \in \mathcal{T}' \}.

Since T\mathcal{T}' is strictly finer than T\mathcal{T}, VT    VTV \in \mathcal{T}' \implies V \in \mathcal{T} but the converse is in general not true. It follows that YVTY    YVTYY \cap V \in \mathcal{T}'_Y \implies Y \cap V \in \mathcal{T}_Y but the converse is also not true in general. This shows that TY\mathcal{T}'_Y is strictly finer than TY\mathcal{T}_Y.

  1. A map f:XYf : X \to Y is said to be an open map if for every open set UU of XX, the set f(U)f(U) is open in YY. Show that π1:X×YX\pi_1 : X \times Y \to X and π2:X×YY\pi_2 : X \times Y \to Y are open maps.

(Note: π1\pi_1 and π2\pi_2 are projection maps of X×YX \times Y onto its first and second factors respectively.)

For every open set of X×YX \times Y (under the product topology), it is generated by a basis given by sets in the form of U×VU \times V, where UU and VV are open respectively in XX and YY.

Consider the images π1(U×V)\pi_1(U \times V) and π2(U×V)\pi_2(U \times V). Indeed, by the definition of function images as well as π1\pi_1 and π2\pi_2,

π1(U×V)={xX:(x,y)U×V}=U, and π2(U×V)={yY:(x,y)U×V}=V.\pi_1(U \times V) = \{x \in X : (x, y) \in U \times V \} = U, \text{ and } \pi_2(U \times V) = \{y \in Y : (x, y) \in U \times V \} = V.

Since for both π1\pi_1 and π2\pi_2, the images of the basis elements of X×YX \times Y are open (respectively in XX and YY), the same applies for every open set of X×YX \times Y (since images preserve arbitrary unions; as images do not preserve intersections in general, for finite intersections, some work needs to be done: note that if U1,,UnXU_1,\ldots,U_n \subseteq X and V1,,VnYV_1,\ldots,V_n \subseteq Y are open, then i=1n(Ui×Vi)=(i=1nUi)×(i=1nVi)\bigcap_{i=1}^{n}(U_i \times V_i)=\left(\bigcap_{i=1}^{n}U_i\right) \times \left(\bigcap_{i=1}^{n}V_i\right), with first and second factors being open as they both are finite intersections of open sets, and this is enough to apply the same argument as above to show that its images under π1\pi_1 and π2\pi_2 are open), so we can conclude that they are open maps.

  1. Show that the countable collection
{(a,b)×(c,d):a<b and c<d and a,b,c,d are rational}\{(a,b) \times (c,d) : a < b \text{ and } c < d \text{ and } a,b,c,d \text{ are rational}\}

is a basis for R2\mathbb{R}^2.

(Note: (a,b)(a,b) denotes {xR:a<x<b}\{x \in \mathbb{R} : a < x < b\}, not Cartesian coordinates.)

Using Theorem 15.1, it suffices to prove that the open intervals (x,y)(x,y), where x<yx < y and xx and yy are rational numbers, form a basis for R\mathbb{R}, which is already shown in Exercise 13.8(a).

Exercise after §17

  1. Let C\mathcal{C} be a collection of subsets of the set XX. Suppose that \emptyset and XX are in C\mathcal{C}, and that finite unions and arbitrary intersections of elements of C\mathcal{C} are in C\mathcal{C}. Show that
T={X\C:CC}\mathcal{T} = \{ X \backslash C : C \in \mathcal{C} \}

is a topology on XX.

Indeed, since X\X=X \backslash X = \emptyset and X\=XX \backslash \emptyset = X, whence \emptyset and XX are in C\mathcal{C}, so \emptyset and XX are in T\mathcal{T}.

Let {Cα}\{C_\alpha\} be an arbitrarily indexed family of sets in C\mathcal{C}, then since

α(X\Cα)=X\α(Cα),\bigcup_{\alpha}(X \backslash C_\alpha)=X \backslash \bigcap_{\alpha}(C_\alpha),

and arbitary intersections of C\mathcal{C} are in C\mathcal{C}, it follows that arbitrary unions of elements in T\mathcal{T} are in T\mathcal{T}.

Let C1,,CnC_1,\ldots,C_n be a finite number of elements in C\mathcal{C}, then since

i=1n(X\Ci)=X\i=1n(Ci),\bigcap_{i=1}^{n}(X \backslash C_i) = X \backslash \bigcup_{i=1}^{n}(C_i),

and finite unions of elements of C\mathcal{C} are in C\mathcal{C}, it follows that finite intersections of elements in T\mathcal{T} are in T\mathcal{T}.

Thus, T\mathcal{T} is a topology on XX.

  1. Show that if AA is closed in YY and YY is closed in XX, then AA is closed in XX.

Suppose that AA is closed in YY and YY is closed in XX, then Y\AY \backslash A and X\YX \backslash Y are open.

Note that (X\Y)(Y\A)=(XY)\(YA)(X \backslash Y) \cup (Y \backslash A) = (X \cup Y) \backslash (Y \cap A) (why?). As AA is a subset of YY and YY is a subset of XX, it follows that YA=AY \cap A = A and XY=XX \cup Y = X, so we can say that X\A=(X\Y)(Y\A)X \backslash A = (X \backslash Y) \cup (Y \backslash A). Since X\AX \backslash A is a union of open sets, it is also open. Therefore, AA is closed in XX by definition.

Alternatively, using Theorem 17.2, since AA is closed in YY, it equals the intersection of a closed set of XX with YY. Meanwhile, YY itself is closed in XX, so this means that AA is an intersection of closed sets in XX. It follows from Theorem 17.1 that AA is closed in XX.

  1. Show that if AA is closed in XX and BB is closed in YY, then A×BA \times B is closed in X×YX \times Y.

Suppose that AA is closed in XX and BB is closed in YY, then X\AX \backslash A is open in XX and Y\BY \backslash B is open in YY. This implies that (X\A)×Y(X \backslash A) \times Y and X×(Y\B)X \times (Y \backslash B) are open in X×YX \times Y.

Note that (X×Y)\(A×B)=((X\A)×Y)(X×(Y\B))(X \times Y) \backslash (A \times B) = ((X \backslash A) \times Y) \cup (X \times (Y \backslash B)) (why?), so it is open. Therefore, A×BA \times B is closed in X×YX \times Y.

  1. Show that if UU is open in XX and AA is closed in XX, then U\AU \backslash A is open in XX, and A\UA \backslash U is closed in XX.

Suppose that UU is open in XX and AA is closed in XX, then X\AX \backslash A is open in XX.

Note that U\A=(X\A)UU \backslash A = (X \backslash A) \cap U, so it is open in XX.

Note that X\(A\U)=(X\A)UX \backslash (A \backslash U) = (X \backslash A) \cup U, so it is open in XX, and thus A\UA \backslash U is closed in XX.

  1. Let XX be an ordered set in the order topology. Show that (a,b)[a,b]\overline{(a,b)} \subseteq [a,b]. Under what conditions does equality hold?

Suppose that x(a,b)x \in \overline{(a,b)}, we show that x[a,b]    axbx \in [a,b] \iff a \leqslant x \leqslant b. From Theorem 17.5, we know that x(a,b)x \in \overline{(a,b)} if and only if every open set that contains xx intersects (a,b)(a,b).

By contrapositivity, we prove that if x[a,b]    x<a or x>bx \notin [a,b] \iff x < a \text{ or } x > b, then x(a,b)x \notin \overline{(a,b)}. If x<ax < a, then we obtain an open set, specifically {yX:y<a}\{y \in X : y < a \}, that contains xx but does not intersect (a,b)(a,b). This implies that x(a,b)x \notin \overline{(a,b)}. If x>bx > b, then we obtain an open set, specifically {yX:y>b}\{y \in X : y > b \}, that contains xx but does not intersect (a,b)(a,b), which again implies that x(a,b)x \notin \overline{(a,b)}. This thus concludes the proof.

For the equality to hold, by Theorem 17.6, aa and bb need to be the only (two) points in XX that are limit points of (a,b)(a,b) but not in (a,b)(a,b).

  1. Criticise the following 'proof' that AαAα\overline{\bigcup A_\alpha} \subseteq \bigcup \overline{A}_\alpha: if {Aα}\{A_\alpha\} is a collection of sets in XX and if xAαx \in \overline{\bigcup A_\alpha}, then every neighbourhood UU of xx intersects Aα\bigcup A_\alpha. Thus UU must intersect some AαA_\alpha, so that xx must belong to the closure of some AαA_\alpha. Therefore, xAαx \in \bigcup \overline{A}_\alpha.

Let UU be a neighbourhood of xAαx \in \overline{\bigcup A_\alpha}. UU intersecting some AαA_\alpha does not mean that xx must belong to the closure of some AαA_\alpha.

For a counterexample, consider the standard topology on R\mathbb{R}. Define AnA_n by (1n+1,1n)\left(\frac{1}{n+1},\frac{1}{n}\right) where nZ+n \in \mathbb{Z}_+. It follows that An=[0,1]\overline{\bigcup A_n} =[0,1]. Take 0An0 \in \overline{\bigcup A_n}. Indeed, every neighbourhood VV of 00 intersects some AnA_n by Theorem 17.5, but the closure of any AnA_n is given by [1n+1,1n]\left[\frac{1}{n+1},\frac{1}{n}\right], and 00 is not in any of them.

  1. Show that every order topology is Hausdorff.

Let XX be an ordered set in the order topology, and x1,x2Xx_1, x_2 \in X such that x1<x2x_1 < x_2 without loss of generality.

If there exists some yXy \in X such that x1<y<x2x_1 < y < x_2, then let U={xX:x<y}U = \{x \in X : x < y\} and V={xX:x>y}V = \{x \in X : x > y\}. We can see that UU is a neighbourhood of x1x_1, VV is a neighbourhood of x2x_2 and UV=U \cap V = \emptyset.

If there is no such yy, then let U={xX:x<x2}U = \{ x \in X : x < x_2 \} and V={xX:x>x1}V = \{ x \in X : x > x_1 \}. Similarly, UU is a neighbourhood of x1x_1, VV is a neighbourhood of x2x_2 and UV=U \cap V = \emptyset.

  1. Show that the product of two Hausdorff spaces is Hausdorff.

Let X×YX \times Y be the product topology constructed from two Hausdorff spaces XX and YY. Let (x1,y1),(x2,y2)X×Y(x_1,y_1) , (x_2,y_2) \in X \times Y be distinct, then three cases need to be considered.

Case I: x1=x2:=xx_1=x_2 := x and y1y2y_1 \neq y_2. Let UxU_x be a neighbourhood of xx in XX. Let Uy1U_{y_1} be a neighbourhood of y1y_1 and Uy2U_{y_2} be a neighbourhood of y2y_2 in YY such that Uy1Uy2=U_{y_1} \cap U_{y_2} = \emptyset due to YY being Hausdorff. It follows that (Ux×Uy1)(Ux×Uy2)=Ux×(Uy1Uy2)=(U_x \times U_{y_1}) \cap (U_x \times U_{y_2}) = U_x \times (U_{y_1} \cap U_{y_2}) = \emptyset, whence Ux×Uy1U_x \times U_{y_1} and Ux×Uy2U_x \times U_{y_2} are respectively neighbourhoods of (x1,y1)(x_1,y_1) and (x2,y2)(x_2,y_2).

Case II: x1x2x_1 \neq x_2 and y1=y2:=yy_1=y_2 := y. Let Ux1U_{x_1} and Ux2U_{x_2} be respectively neighbourhoods of x1x_1 and x2x_2 in XX such that Ux1Ux2=U_{x_1} \cap U_{x_2}=\emptyset due to XX being Hausdorff. Let UyU_y be a neighbourhood of yy in YY. It follows that (Ux1×Uy)(Ux2×Uy)=(Ux1Ux2)×Uy=(U_{x_1} \times U_y) \cap (U_{x_2} \times U_y) = (U_{x_1} \cap U_{x_2}) \times U_y = \emptyset, whence Ux1×UyU_{x_1} \times U_y and Ux2×UyU_{x_2} \times U_y are respectively neighbourhoods of (x1,y1)(x_1,y_1) and (x2,y2)(x_2,y_2).

Case III: x1x2x_1 \neq x_2 and y1y2y_1 \neq y_2. Let Ux1U_{x_1} and Ux2U_{x_2} be respectively neighbourhoods of x1x_1 and x2x_2 in XX such that Ux1Ux2=U_{x_1} \cap U_{x_2} = \emptyset due to XX being Hausdorff. Let Uy1U_{y_1} and Uy2U_{y_2} be respectively neighbourhoods of y1y_1 and y2y_2 in YY such that Uy1Uy2=U_{y_1} \cap U_{y_2} = \emptyset due to YY being Hausdorff. It follows that (Ux1×Uy1)(Ux2×Uy2)=(Ux1Ux2)×(Uy1Uy2)=(U_{x_1} \times U_{y_1}) \cap (U_{x_2} \times U_{y_2}) = (U_{x_1} \cap U_{x_2}) \times (U_{y_1} \cap U_{y_2}) = \emptyset, whence Ux1×Uy1U_{x_1} \times U_{y_1} and Ux2×Uy2U_{x_2} \times U_{y_2} are respectively neighbourhoods of (x1,y1)(x_1,y_1) and (x2,y2)(x_2,y_2).

Combining all the cases above tells us that the product of two Hausdorff spaces is Hausdorff.

  1. Show that a subspace of a Hausdorff space is Hausdorff.

Let XX be a Hausdorff space and YY be a subset of XX equipped with the subspace topology inherited from XX.

Let y1,y2Yy_1, y_2 \in Y such that y1y2y_1 \neq y_2. Note that y1,y2Xy_1, y_2 \in X as well and XX is Hausdorff, so there exists a neighbourhood UU of y1y_1 and VV of y2y_2 in XX such that UV=U \cap V = \emptyset. It follows that YUY \cap U and YVY \cap V are neighbourhoods of y1y_1 and y2y_2 respectively in YY as a subspace such that (YU)(YV)=Y(UV)=(Y \cap U) \cap (Y \cap V)= Y \cap (U \cap V) = \emptyset.

Thus, a subspace of a Hausdorff space is Hausdorff.

  1. Show that XX is Hausdorff if and only if the diagonal Δ={(x,x):xX}\Delta = \{ (x,x) : x \in X\} is closed in X×XX \times X.

Suppose that XX is Hausdorff. We claim that (X×X)\Δ(X \times X) \backslash \Delta is open in X×XX \times X, so that Δ\Delta is closed in X×XX \times X. Note that for any element in (X×X)\Δ(X \times X) \backslash \Delta, the first and second coordinates are not equal.

Since for x1,x2Xx_1, x_2 \in X such that x1x2x_1 \neq x_2, there exists a neighbourhood UU of x1x_1 and a neighbourhood VV of x2x_2 such that UV=U \cap V = \emptyset. It follows that U×VU \times V is open and does not contain (x1,x1)(x_1,x_1) nor (x2,x2)(x_2,x_2). Taking unions of such U×VU \times V over all distinct pairs of elements in XX, it follows that (X×X)\Δ(X \times X) \backslash \Delta can be expressed as a union of open sets, so it is open.

Conversely, if Δ\Delta is closed in X×XX \times X, then (X×X)\Δ(X \times X) \backslash \Delta is open in X×XX \times X. Take a point (x1,x2)(X×X)\Δ(x_1,x_2) \in (X \times X) \backslash \Delta, then x1x2x_1 \neq x_2. Since (X×X)\Δ(X \times X) \backslash \Delta is open, there exists a neighbourhood UXU \subseteq X of x1x_1 and a neighbourhood VXV \subseteq X of x2x_2 such that U×V(X×X)\ΔU \times V \subseteq (X \times X) \backslash \Delta. This implies that UU and VV do not have common elements, i.e. UV=U \cap V = \emptyset; if it did have one, say, xx', then (x,x)(X×X)\Δ(x',x') \in (X \times X) \backslash \Delta, a contradiction. Therefore, XX is Hausdorff.

  1. In the finite complement topology on R\mathbb{R}, to what point or points does the sequence xn=1nx_n=\frac{1}{n} converge?

Recall that a sequence xnx_n converges to xx if there is a positive integer NN that depends on each neighbourhood UU of xx such that xnUx_n \in U whenever nNn \geqslant N.

Denote the finite complement topology on R\mathbb{R} by Tf\mathcal{T}_f. Each open set VV of Tf\mathcal{T}_f must satisfy that R\V\mathbb{R} \backslash V is finite or all of R\mathbb{R}.

Since every open set under Tf\mathcal{T}_f contains either all, or, all but finitely many values of xnx_n (because each non-empty open set excludes only finitely many elements of R\mathbb{R}), we can say that the sequence xnx_n converges to every real number in R\mathbb{R}.

To see this, we fix some xRx \in \mathbb{R}. For each neighbourhood UU of xx (so UU is non-empty), as the number of elements in R\mathbb{R} that UU excludes is finite, it suffices to choose NN to be N+1N'+1 where NN' is the maximum positive integer such that 1N\frac{1}{N'} is not in UU if UU does not contain some (finitely many) numbers in the form of 1n\frac{1}{n}, or choose NN to be 11 if UU contains every number in the form of 1n\frac{1}{n}.

  1. Show that T1T_1 axiom is equivalent to the condition that for each pair of points of XX, each has a neighbourhood not containing the other (point).

Let XX be a topological space that satisfies the T1T_1 axiom, then every one-point subsets of XX is closed. Take x1,x2Xx_1, x_2 \in X such that x1x2x_1 \neq x_2, then X\{x1}X \backslash \{x_1\} is a neighbourhood of x2x_2 that does not contain x1x_1 whereas X\{x2}X \backslash \{x_2\} is a neighbourhood of x1x_1 that does not contain x2x_2, thus proving one direction of the equivalence.

Conversely, suppose that x,xXx, x' \in X with xxx \neq x' such that xx' has a neighbourhood UU that does not contain xx (and vice versa), then U{x}=U \cap \{x\} = \emptyset. By Theorem 17.5, this implies that xx' is not in the closure of {x}\{x\}. Since the choice of xx' is arbitrary, it follows that the closure of {x}\{x\} is {x}\{x\} itself, so {x}\{x\} is closed. Since the choice of xXx \in X is also arbitrary, every one-point subset of XX is closed.

  1. Consider the lower limit topology on R\mathbb{R} and the topology given by the basis C\mathcal{C} of Exercise 8 of §13. Determine the closures of the intervals A=(0,2)A = (0,\sqrt{2}) and B=(2,3)B = (\sqrt{2},3) in these two topologies.

Using Theorem 17.6, it suffices to find all the limit points of AA and BB under the two respective topologies.

Under the lower limit topology on R\mathbb{R}, if xx is a limit point of AA, then every neighbourhood UU of xx intersects AA in a point other than xx. This implies that UU must contain some half-open interval in the form of [a,b)[a,b) that contains some point of AA. If xAx \notin A, it follows that x=0x = 0 necessarily because if x=εx=-\varepsilon where ε>0\varepsilon > 0 is arbitrary, then we obtain a neighbourhood [ε,ε2)\left[-\varepsilon,-\frac{\varepsilon}{2}\right) that does not intersect AA. Similarly, if x=2+δx = \sqrt{2}+\delta where δ0\delta \geqslant 0, then [2+δ,)[\sqrt{2}+\delta,\infty) is a neighbourhood that does not intersect AA. On the other hand, if x=0x = 0, then every neighbourhood must contain some points in between 00 and 2\sqrt{2}, whether or not 00 is the lower endpoint of a half-open interval. Therefore, the closure of AA under this topology is [0,2)[0,\sqrt{2}). By a similar argument, the closure of the BB under this topology is [2,3)[\sqrt{2},3).

Under the topology given by the basis C\mathcal{C}, if xx is a limit point of AA, then every neighbourhood UU of xx must contain some half-open interval in the form of [a,b)[a,b) where aa and bb are rational that contains some point of AA. If xAx \notin A, then x=0x=0 or x=2x=\sqrt{2} necessarily because any neighbourhood that contains 00 (resp. 2\sqrt{2}) must contain some points in between 00 and 2\sqrt{2}, since 2\sqrt{2} cannot be the lower endpoint of a half-open interval under the C\mathcal{C}-generated topology. This implies that the closures of AA and BB under this topology are respectively [0,2][0,\sqrt{2}] and [2,3][\sqrt{2},3].

  1. If AXA \subseteq X, we define the boundary of AA by the equation
Bd  A=A(X\A). \mathrm{Bd} \; A = \overline{A} \cap \overline{(X \backslash A)}.

(a) Show that IntA\mathrm{Int} A and Bd  A\mathrm{Bd} \; A are disjoint, and A=IntABd  A\overline{A}=\mathrm{Int} A \cup \mathrm{Bd} \; A.

(b) Show that Bd  A=    A is both open and closed \mathrm{Bd} \; A = \emptyset \iff A \text{ is both open and closed}.

(c) Show that U is open     Bd  U=U\UU \text{ is open } \iff \mathrm{Bd}\; U=\overline{U} \backslash U.

(d) If UU is open, is it true that U=Int(U)U = \mathrm{Int}(\overline{U})? Justify your answer.

  1. (Kuratowski) Consider the collection of all subsets AA of the topological space XX. The operations of closure AAA \to \overline{A} and complementation AX\AA \to X \backslash A are functions from this collection to itself.

(a) Show that starting with a given set AA, one can form no more than 14 distinct sets by applying these two operations successively.

(b) Find a subset AA of R\mathbb{R} (in its usual topology) for which the maximum of 14 is obtained.

Exercise after §18

  1. Prove that for functions f:RRf : \mathbb{R} \to \mathbb{R}, the ε\varepsilon-δ\delta definition of continuity implies the open set definition.
note

The arguments to show Exercise 18.1 can be modified to prove the same for real multivariate functions, though this may require the notion of box topology to define the standard topology on Rn\mathbb{R}^n.

Let x0Rx_0 \in \mathbb{R}. Suppose that for a every ε>0\varepsilon > 0, there exists δ>0\delta > 0 such that for xRx \in \mathbb{R} such that xx0<δ|x-x_0|<\delta, we have f(x)f(x0)<ε|f(x)-f(x_0)|<\varepsilon.

Let UU be an open subset of R\mathbb{R} that contains f(x0)f(x_0), then it contains some open interval about f(x0)f(x_0), which we can express in the form (f(x0)ε,f(x0)+ε)(f(x_0)-\varepsilon, f(x_0)+\varepsilon), for some ε>0\varepsilon>0. For that open interval, there then exists a corresponding δ\delta such that xx0<δ| x - x_0 | < \delta where xRx \in \mathbb{R}, and {xR:xx0<δ}=(x0δ,x0+δ)\{x \in \mathbb{R} : |x-x_0|<\delta \} = (x_0-\delta,x_0+\delta) is open in R\mathbb{R}. This implies that f1((f(x0)ε,f(x0)+ε))=(x0δ,x0+δ)f^{-1}((f(x_0)-\varepsilon,f(x_0)+\varepsilon)) = (x_0-\delta,x_0+\delta) is open.

Note that U=(Uf(R))(U\f(R))U = (U \cap f(\mathbb{R})) \cup (U \backslash f(\mathbb{R})) (why?). For points in Uf(R)U \cap f(\mathbb{R}), we can apply the argument above to apply an open subset of the domain. For subsets of U\f(R)U \backslash f(\mathbb{R}), their inverse images are empty, which is also open.

Since arbitrary unions of open sets are open, and inverse images preserve unions, if follows that f1(U)f^{-1}(U) is an open subset of R\mathbb{R}, thus proving the open set definition.

  1. Suppose that f:XYf : X \to Y is continuous. If xx is a limit point of the subset AA of XX, is it necessarily true that f(x)f(x) is a limit point of f(A)f(A)?

Yes. We prove this assertion by contraposition.

Suppose that f(x)f(x) is not a limit point of f(A)f(A). There then exists a neighbourhood UYU \subseteq Y of f(x)f(x) that does not intersect f(A)f(A) in a point different from f(x)f(x). Note that f(X)f(X) must intersect UU since f(x)Uf(x) \in U, so since f:XYf : X \to Y is continuous, f1(U)f^{-1}(U) must be a neighbourhood of xx. However, since UU does not contain any point in f(A)f(A) other than f(x)f(x) itself, it follows that f1(U)f^{-1}(U) does not contain any point in AA other than xx itself. This implies that xx is not a limit point of AA.

  1. Given x0Xx_0 \in X and y0Yy_0 \in Y, show that the maps f:XX×Yf : X \to X \times Y and g:YX×Yg : Y \to X \times Y defined by

    f(x)=(x,y0)andg(y)=(x0,y)f(x) = (x, y_0) \quad \text{and} \quad g(y)=(x_0,y)

    are imbeddings.

We prove the desired statement through a number of steps.

  1. We claim that ff and gg are injective.

    Indeed, for x1,x2Xx_1, x_2 \in X and y1,y2Yy_1, y_2 \in Y, f(x1)=f(x2)    (x1,y0)=(x2,y0)    x1=x2f(x_1)=f(x_2) \implies (x_1,y_0)=(x_2,y_0) \implies x_1=x_2 and g(y1)=g(y2)    (x0,y1)=(x0,y2)    y1=y2g(y_1)=g(y_2) \implies (x_0,y_1)=(x_0,y_2) \implies y_1=y_2.

  2. We claim that ff and gg are continuous.

    Using Theorem 18.4, it suffices to prove that the respective coordinate functions of ff and gg are continuous. If we write f(x)=(f1(x),f2(x))f(x)=(f_1(x),f_2(x)) and g(y)=(g1(y),g2(y))g(y)=(g_1(y),g_2(y)), then f1(x)=xf_1(x)=x, f2(x)=y0f_2(x)=y_0, g1(y)=x0g_1(y)=x_0 and g2(y)=yg_2(y)=y. Since f1f_1 and g2g_2 are identity functions, they are continuous as for every open subset UU of XX (resp. YY), we have f1(U)=Uf^{-1}(U)=U. Since f2f_2 and g1g_1 are constant functions, they are also continuous (by Theorem 18.2).

  3. We claim that the functions f:Xf(X)f' : X \to f(X) and g:Yg(Y)g' : Y \to g(Y) obtained by restricting their codomains to their respective images are homeomorphisms.

    Indeed, the inverses of ff' and gg' are respectively given by the projections f1(x,y)=xf'^{-1}(x,y)=x and g1(x,y)=yg'^{-1}(x,y)=y. We know that projections are continuous, so ff' and gg' are homeomorphisms.

This concludes the proof that ff and gg are embeddings.

  1. Show that the subspace (a,b)(a,b) of R\mathbb{R} is homeomorphic with (0,1)(0,1) and the subspace [a,b][a,b] of R\mathbb{R} is homeomorphic with [0,1][0,1].

Define the function f:(a,b)(0,1)f : (a,b) \to (0,1) by f(x)=xabaf(x)=\dfrac{x-a}{b-a}. The inverse is then given by f1(x)=x(ba)+af^{-1}(x)=x(b-a)+a. Since ff is a rational function whereas f1f^{-1} is a polynomial function, they are both continuous. Thus, (a,b)(a,b) and (0,1)(0,1) are homeomorphic. A very similar argument can be used to show that [a,b][a,b] and [0,1][0,1] are homeomorphic.

  1. Find a function f:RRf : \mathbb{R} \to \mathbb{R} that is continuous at precisely one point.
  1. Let YY be an ordered set in the order topology. Let f,g:XYf,g : X \to Y be continuous.

(a) Show that the set {x:f(x)g(x)}\{x : f(x) \leqslant g(x)\} is closed in XX.

(b) Let h:XYh : X \to Y be the function

h(x)=min{f(x),g(x)}.h(x)=\min\{f(x),g(x)\}.

Show that hh is continuous. [Hint: Use the pasting lemma.]

(a) We show that its complement, given by the set W:={x:f(x)>g(x)}W := \{x : f(x)>g(x)\}, is open in XX. Take an element x0Wx_0 \in W, then f(x0)>g(x0)f(x_0)>g(x_0). Since ff is continuous, we can take a neighbourhood UU of f(x0)f(x_0) such that Ug(A)U \subseteq g(A) and f(x0)>uf(x_0)>u for all uUu \in U; there then exists a neighbourhood UU' of x0x_0 with f(U)Uf(U') \subseteq U. Note that this implies that UWU' \subseteq W. Similarly, since gg is continuous, we can take a neighbourhood VV of g(x0)g(x_0) such that Vf(A)V \subseteq f(A) and v>g(x0)v>g(x_0) for all vVv \in V; there then exists a neighbourhood VV' of x0x_0 with f(V)Vf(V') \subseteq V, so that VWV' \subseteq W.

Taking unions of such UU' and VV' over all x0Wx_0 \in W, we obtain the entire WW. This implies that WW is a union of open sets, so it is open, thus {x:f(x)g(x)}\{x : f(x) \leqslant g(x)\} is closed in XX.

(Alternative proofs.)

(b) Note that f:XYf : X \to Y and g:XYg : X \to Y are continuous. Note as well that from (a), U:={x:f(x)g(x)}U := \{x : f(x) \leqslant g(x)\} and V:={x:f(x)g(x)}V := \{x : f(x) \geqslant g(x) \} are closed subsets of XX (since ff and gg as continuous functions are arbitrary).

Consider the restriction of ff to UU and the restriction of gg to VV. It follows that UV=XU \cup V = X, as for any xXx \in X, exactly one of f(x)g(x)f(x) \leqslant g(x) or g(x)f(x)g(x) \leqslant f(x) holds. We also have that f(x)=g(x)f(x)=g(x) for any xUVx \in U \cap V, since UVU \cap V is precisely {x:f(x)=g(x)}\{x : f(x)=g(x)\}. By Pasting Lemma, these two restrictions combine to give a continuous function h:XYh : X \to Y defined by h(x)={f(x)if xU,g(x)if xVh(x)=\begin{cases} f(x) & \text{if } x \in U, \\ g(x) & \text{if } x \in V \end{cases}, which in this case is equivalent to h(x)=min{f(x),g(x)}h(x)=\min\{f(x),g(x)\}.

  1. Let f:ABf : A \to B and g:CDg : C \to D be continuous functions. Let us define a map f×g:A×CB×Df \times g : A \times C \to B \times D by the equation

    (f×g)(a,c)=(f(a),g(c)).(f \times g)(a,c)=(f(a),g(c)).

    Show that f×gf \times g is continuous.

  1. Let F:X×YZF : X \times Y \to Z. We say that FF is continuous in each variable separately if for each y0y_0 in YY, the map h:XZh : X \to Z defined by h(x)=F(x,y0)h(x)=F(x, y_0) is continuous, and for each x0x_0 in XX, the map k:YZk : Y \to Z defined by k(y)=F(x0,y)k(y)=F(x_0, y) is continuous. Show that if FF is continuous, then FF is continuous in each variable separately.
  1. Let F:RRF : \mathbb{R} \to \mathbb{R} be defined by the equation
F(x,y)={xyx2+y2if (x,y)(0,0),0if (x,y)=(0,0).F(x,y)=\begin{cases} \frac{xy}{x^2+y^2} & \text{if } (x,y) \neq (0,0), \\ 0 & \text{if } (x,y)=(0,0)\end{cases}.

(a) Show that FF is continuous in each variable separately.

(b) Compute the function g:RRg : \mathbb{R} \to \mathbb{R} defined by g(x)=F(x,x)g(x)=F(x,x).

(c) Show that FF is not continuous.

note

The function given by Exercise 18.12 is a classic example in the study of multivariable calculus to demonstrate a multivariate function that is not differentiable at the origin.

  1. Let AXA \subseteq X; let f:AYf : A \to Y be continuous; let YY be Hausdorff. Show that if ff may be extended to a continuous function g:AYg : \overline{A} \to Y, then gg is uniquely determined by ff.

Exercise after §19

  1. Prove Theorem 19.2.
  1. Prove Theorem 19.3.
  1. Prove Theorem 19.4.
  1. Show that (X1×Xn1)×Xn(X_1 \times \cdots X_{n-1}) \times X_n is homeomorphic with X1××XnX_1 \times \cdots \times X_n.
  1. One of the implications stated in Theorem 19.6 holds for the box topology. Which one?
  1. Let x1\mathbf{x}_1, x2\mathbf{x}_2 be a sequence of the points of the product space Xα\prod X_\alpha. Show that this sequence converges to the points x\mathbf{x} if and only if the sequence πα(x1)\pi_\alpha(\mathbf{x}_1), πα(x2)\pi_\alpha(\mathbf{x}_2), ... converges to πα(x)\pi_\alpha(\mathbf{x}) for each α\alpha. Is this fact true if one uses the box topology instead of the product topology?

Given sequences (a1,a2,)(a_1,a_2,\ldots) and (b1,b2,)(b_1,b_2,\ldots) of real numbers with ai>0a_i > 0 for all ii, define h:RωRωh : \mathbb{R}^\omega \to \mathbb{R}^\omega by the equation

h((x1,x2,))=(a1x1+b1,a2x2+b2,).h((x_1,x_2,\ldots)) = (a_1 x_1+b_1, a_2 x_2+b_2, \ldots).

Show that if Rω\mathbb{R}^\omega is given the product topology, hh is a homeomorphism of Rω\mathbb{R}^\omega with itself. What happens if Rω\mathbb{R}^\omega is given the box topology?

Show that the choice axiom is equivalent to the statement that for any indexed family {Aα}αJ\{A_\alpha\}_{\alpha \in J} of nonempty sets, with J0J \neq 0, the cartesian product

αJAα\prod_{\alpha \in J}A_\alpha

is not empty.

Exercise after §20

  1. (a) In Rn\mathbb{R}^n, define
d(x,y)=x1y1++xnyn.d'(\mathbf{x},\mathbf{y})=|x_1-y_1|+\cdots+|x_n-y_n|.

Show that dd' is a metric that induces the usual topology of Rn\mathbb{R}^n. Sketch the basis elements under dd' when n=2n=2.

(b) More generally, given p1p \geqslant 1, define

d(x,y)=[i=1nxiyip]1pd'(\mathbf{x},\mathbf{y})=\left[\sum_{i=1}^{n}|x_i-y_i|^p\right]^{\frac{1}{p}}

for x,yRn\mathbf{x} , \mathbf{y} \in \mathbb{R}^n. Assume that dd' is a metric. Show that it induces the usual topology on Rn\mathbb{R}^n.

  1. Show that R×R\mathbb{R} \times \mathbb{R} in the dictionary order topology is metrizable.
  1. Let XX be a metric space with metric dd.

(a) Show that d:X×XRd : X \times X \to \mathbb{R} is continuous.

(b) Let XX' denote a space having the same underlying set as XX. Show that if d:X×XRd' : X \times X \to \mathbb{R} is continuous, then the topology of XX' is finer than the topology of XX.

note

One can summarise the result of this exercise as follows: if XX has a metric dd, then the topology induced by dd is the coarsest topology relative to which the function dd is continuous.

  1. Show that the Euclidean metric dd on Rn\mathbb{R}^n is a metric, as follows: if x,yRn\mathbf{x}, \mathbf{y} \in \mathbb{R}^n and cRc \in \mathbb{R}, define
x+y=(x1+y1,,xn+yn),\mathbf{x}+\mathbf{y}=(x_1+y_1,\ldots,x_n+y_n),cx=(cx1,,cxn),c\mathbf{x}=(cx_1,\ldots,cx_n),xy=x1y1++xnyn.\mathbf{x} \cdot \mathbf{y} = x_1 y_1 +\cdots + x_n y_n.

(a) Show that x(y+z)=(xy)+(xz)\mathbf{x} \cdot (\mathbf{y}+\mathbf{z})=(\mathbf{x}\cdot\mathbf{y})+(\mathbf{x}\cdot\mathbf{z}).

(b) (Cauchy-Schwarz inequality.) Show that xyxy| \mathbf{x} \cdot \mathbf{y} | \leqslant \Vert\mathbf{x}\Vert\Vert\mathbf{y}\Vert. [Hint: If x,y0\mathbf{x}, \mathbf{y} \neq 0, let a=1xa = \frac{1}{\Vert\mathbf{x}\Vert} and b=1yb = \frac{1}{\Vert\mathbf{y}\Vert}, and use the fact that ax+by0\Vert a\mathbf{x}+b\mathbf{y}\Vert \geqslant 0.]

(c) Show that x+yx+y\Vert \mathbf{x}+\mathbf{y} \Vert \leqslant \Vert\mathbf{x}\Vert + \Vert\mathbf{y}\Vert. [Hint: Compute (x+y)(x+y)(\mathbf{x}+\mathbf{y}) \cdot (\mathbf{x}+\mathbf{y}) and apply (b).]

(d) Verify that dd is a metric.

  1. Show that if dd is a metric for XX, then
d(x,y)=d(x,y)1+d(x,y)d'(x,y)=\frac{d(x,y)}{1+d(x,y)}

is a bounded metric that gives the topology of XX.

[Hint: If f(x)=x1+xf(x) = \dfrac{x}{1+x} for x>0x > 0, use the mean-value theorem to show that f(a+b)f(b)f(a)f(a+b)-f(b) \leqslant f(a).]

Following the hint, define f(x)=x1+xf(x)=\dfrac{x}{1+x} with x>0x>0. It follows that f(x)=1(1+x)2f'(x)=\dfrac{1}{(1+x)^2}. Let a,b>0a,b>0 such that a>ba>b without loss of generality. By mean-value theorem, since b<a<a+bb < a < a+b, we have that f(a)=f(a+b)f(b)a    af(a)=f(a+b)f(b)f'(a)=\dfrac{f(a+b)-f(b)}{a} \implies af'(a)=f(a+b)-f(b). Meanwhile, a(1+a)2a1+a=a2(1+a)20    af(a)f(a)\dfrac{a}{(1+a)^2}-\dfrac{a}{1+a}=-\dfrac{a^2}{(1+a)^2} \leqslant 0 \implies af'(a) \leqslant f(a), so f(a+b)f(b)f(a)    f(a+b)f(a)+f(b)f(a+b) - f(b) \leqslant f(a) \implies f(a+b) \leqslant f(a)+f(b).

Note as well that f(x)f(x) is increasing within x>0x>0, as f(x)=111+xf(x)=1-\dfrac{1}{1+x}, whence 11+x\dfrac{1}{1+x} decreases as xx increases and is positive.

Let x,y,zRx,y,z \in \mathbb{R} be arbitrary. Since d(x,y)0d(x,y) \geqslant 0 and the equality holds if and only if x=yx=y, it follows that 1+d(x,y)01+d(x,y) \geqslant 0 as well, so d(x,y)0d'(x,y) \geqslant 0 and the equality holds if and only if x=yx=y as the numerator in dd' is determined solely by dd. By definition, we can also see that d(x,y)=d(y,x)d'(x,y)=d'(y,x). To show the triangle inequality property of dd', we can apply the fact proven in the previous paragraph and let a=d(x,y)a=d(x,y) and b=d(y,z)b=d(y,z); note as well that d=fdd' = f \circ d. We then have that d(x,z)=f(d(x,z))f(d(x,y)+d(y,z))f(d(x,y))+f(d(y,z))=d(x,y)+d(y,z)d'(x,z) = f(d(x,z)) \leqslant f(d(x,y)+d(y,z)) \leqslant f(d(x,y))+f(d(y,z))=d'(x,y)+d'(y,z), with the first \leqslant relation being a consequence of the triangle inequality of dd and ff increasing when x>0x>0.

Since for any x,yRx,y \in \mathbb{R}, d(x,y)1+d(x,y)1+d(x,y)1+d(x,y)=1\dfrac{d(x,y)}{1+d(x,y)} \leqslant \dfrac{1+d(x,y)}{1+d(x,y)}=1, we conclude that dd' is bounded by M=1M=1.

Now, we show that dd' gives the topology of XX that is equal to that induced by dd. [TODO]

Exercise after §21

  1. Let AXA \subseteq X. If dd is a metric for the topology of XX, show that dA×Ad|_{A \times A} is a metric for the subspace topology on AA.
  1. Let XX and YY be metric spaces with metrics dXd_X and dYd_Y, respectively. Let f:XYf : X \to Y have the property that for every pair of points x1,x2x_1, x_2 of XX,
dY(f(x1),f(x2))=dX(x1,x2).d_Y(f(x_1),f(x_2))=d_X(x_1,x_2).

Show that ff is an embedding. It is caleld an isometric embedding of XX in YY.

  1. Let XnX_n be a metric space with metric dnd_n, for nZ+n \in \mathbb{Z}_+.

(a) Show that

ρ(x,y)=max{d1(x1,y1),,dn(xn,yn)}\rho(x,y)=\max\{d_1(x_1,y_1), \ldots, d_n(x_n,y_n)\}

is a metric for the product space X1××XnX_1 \times \cdots \times X_n.

(b) Let di=min{di,1}\overline{d_i}=\min\{d_i,1\}. Show that

D(x,y)=sup{di(xi,yi)i}D(x,y)=\sup\left\{\frac{\overline{d_i}(x_i,y_i)}{i}\right\}

is a metric for the product space Xi\prod X_i.

  1. Theorem (i.e. very useful result). Let xnxx_n \to x and ynyy_n \to y in the space R\mathbb{R}. Then
xn+ynx+y,x_n+y_n \to x+y,xnynxy,x_n-y_n \to x-y,xnynxy,x_n y_n \to xy,

and provided that each yn0y_n \neq 0 and y0y \neq 0,

xn/ynx/y.x_n/y_n \to x/y.

[Hint: Apply Lemma 21.4; recall from the exercises of §19 that if xnxx_n \to x and ynyy_n \to y, then (xn,yn)(x,y)(x_n,y_n) \to (x,y).]

  1. Define fn:[0,1]Rf_n : [0,1] \to \mathbb{R} by the equation fn(x)=xnf_n(x)=x^n. Show that the sequence (fn(x))(f_n(x)) converges for each x[0,1]x \in [0,1], but that the sequence (fn)(f_n) does not converge uniformly.
  1. Let XX be a topological space and let YY be a metric space. Let fn:XYf_n : X \to Y be a sequence of continuous functions. Let xnx_n be a sequence of points of XX converging to xx. Show that if the sequence (fn)(f_n) converges uniformly to ff, then (fn(xn))(f_n(x_n)) converges to f(x)f(x).
  1. Prove the following standard facts about infinite series:

(a) (Monotone convergence theorem). Show that if (sn)(s_n) is a bounded sequence of real numbers and snsn+1s_n \leqslant s_{n+1} for each nn, then (sn)(s_n) converges.

(b) Let (an)(a_n) be a sequence of real numbers; define

sn=i=1nai.s_n=\sum_{i=1}^{n}a_i.

If snss_n \to s, we say that the infinite series

i=1ai\sum_{i=1}^{\infty}a_i

converges to ss also. Show that if ai\sum a_i converges to ss and bi\sum b_i converges to tt, then (cai+bi)\sum (ca_i+b_i) converges to cs+tcs+t.

(c) Prove the comparison test for infinite series: If aibi|a_i| \leqslant b_i for each ii, and if the series bi\sum b_i converges, then the series ai\sum a_i converges. [Hint: show that the series ai\sum |a_i| and ci\sum c_i converge, where ci=ai+aic_i=|a_i|+a_i.]

(d) Given a sequence of functions fn:XRf_n : X \to \mathbb{R}, let

sn(x)=i=1nfi(x).s_n(x)=\sum_{i=1}^{n}f_i(x).

Prove the Weierstrass M-test for uniform convergence: If fi(x)Mi|f_i(x)| \leqslant M_i for all xXx \in X and for all ii, and if the series Mi\sum M_i converges, then the sequence (sn)(s_n) converges uniformly to a function ss. [Hint: Let rn=i=n+1Mir_n=\sum_{i=n+1}^{\infty}M_i. Show that if k<nk<n, then sk(x)sn(x)rn|s_k(x)-s_n(x)| \leqslant r_n; conclude that s(x)sn(x)rn|s(x)-s_n(x)| \leqslant r_n.]

  1. Prove continuity of the algebraic operations on R\mathbb{R}, as follows: Use the metric d(a,b)=abd(a,b)=|a-b| on R\mathbb{R} and the metric on R2\mathbb{R}^2 given by the equation
ρ((x,y),(x0,y0))=max{xx0,yy0}.\rho((x,y),(x_0,y_0))=\max\{|x-x_0|,|y-y_0|\}.

(a) Show that addition is continuous. [Hint: Given ε\varepsilon, let δ=ε/2\delta=\varepsilon/2 and note that d(x+y,x0+y0)xx0+yy0d(x+y,x_0+y_0) \leqslant |x-x_0|+|y-y_0|.]

(b) Show that multiplication is continuous. [Hint: Given (x0,y0)(x_0,y_0) and 0<ε<10 < \varepsilon < 1, let 3δ=ε/(x0+y0+1)3\delta=\varepsilon/(|x_0|+|y_0|+1) and note that d(xy,x0y0)x0yy0+y0xx0+xx0yy0d(xy,x_0y_0) \leqslant |x_0||y-y_0|+|y_0||x-x_0|+|x-x_0||y-y_0|.]

(c) Show that the operation of taking reciprocals is a continuous map from R\{0}\mathbb{R} \backslash \{0\} to R\mathbb{R}. [Hint: show the inverse image of the interval (a,b)(a,b) is open. Consider five cases, according as aa and bb are positive, negative, or zero.]

(d) Show that the subtraction and quotient operations are continuous.

Exercise after §22

  1. Check the details of Example 3.

    i.e.

    Let pp be the map of the real line R\mathbb{R} onto the three-point set A={a,b,c}A = \{a,b,c\} defined by

    p(x)={aif x>0,bif x<0,cif x=0.p(x)=\begin{cases} a & \text{if } x>0, \\ b & \text{if } x<0, \\ c & \text{if } x=0. \end{cases}

    Check that the quotient topology on AA induced by pp is given by T:={,A,{a},{b},{a,b}}\mathcal{T} := \{\emptyset, A, \{a\},\{b\}, \{a,b\}\}.

  1. (a) Let p:XYp : X \to Y be a continuous map. Show that if there is a continuous map f:YXf : Y \to X such that pfp \circ f equals the identity map of YY (i.e. ff is the right inverse [section] of pp), then pp is a quotient map.

    (b) If AXA \subseteq X, a retraction of XX onto AA is a continuous map r:XAr : X \to A such that r(a)=ar(a)=a for each aAa \in A. Show that a retraction is a quotient map.

  1. Let π1:R×RR\pi_1 : \mathbb{R} \times \mathbb{R} \to \mathbb{R} be projection on the first coordinate. Let AA be the subspace of R×R\mathbb{R} \times \mathbb{R} consisting of all points (x,y)(x,y) for which either x0x \geqslant 0 or y=0y=0 (or both); let q:ARq : A \to \mathbb{R} be obtained by restricting π1\pi_1.

    Show that qq is a quotient map that is neither open nor closed.

  1. Let p:XYp : X \to Y be an open map. Show that if AA is open in XX, then the map q:Ap(A)q : A \to p(A) obtained by restricting pp is an open map.

Footnotes

  1. Note that finitude is a special case of countability.