Topology
Reference:
- UIUC math432 Set Theory and Topology by Prof. Roy Araiza.
- Set Theory and Metric Spaces by Irving Kaplansky.
- Topology by James Munkres
1. Metric Spaces
1.1. DONE Definitions
A metric space is a set \(\mathcal{X}\) equipped with a metric \(d: X \times X \to [0, + \infty)\) satisfying:
- \(d(x, x) = 0\) for all \(x\),
- \(d(x, y) > 0\) for \(x \neq y\),
- \(d(x, y) = d(y, x)\),
- \(d(x, z) \leq d(x, y) + d(y, z)\).
1.2. Open Sets, Interior
1.2.1. Neighbourhoods, Open Sets
A (\(r\)-)neighbourhood with center \(x\) and radius \(r\), denoted by \(O_{r}(x)\), is the set \(\{ y \in X: d(x, y) < r \}\).
Also called open ball. In some textbooks, a neighbourhood of \(x\) is a subset containing an open set containing \(x\), and an open set is a set where for any element \(x\) there is an open ball of it completely inside the set. The definitions here are a simplification of those concepts, directly defining open sets in terms of neighbourhoods.
\(x\) is an interior point of \(E \subset X\) if there exists a neighbourhood of \(x\) completely contained in \(E\).
In a metric space a subset \(E\) is open (relative to the metric space) if for every \(x \in E\) there exists a neighbourhood of \(x\) completely inside \(E\), i.e., every point of \(E\) is an interior point of \(E\).
Any neighbourhood in metric space is an open set.
Proof: easy.
In a metric space any union of open sets is open.
Let \(U\) be the union of the open sets \(U_{i}\)’s. Let \(x \in U\). Then \(x \in U_{i}\) for some \(U_{i}\). In that \(U_{i}\) we have a neighbourhood of \(x\) entirely contained in \(U_{i}\) and hence in \(U\). Thus, \(U\) is an open set.
A subset of a metric space is open iff it is a union of neighbourhoods.
By Prop. 1 and Thm. 1, unions of neighbourhoods are open sets.
Conversely, let \(U\) be open. For any set \(x \in U\), there exists a \(O_{r_{x}}(x) \subset U\). Consider the union \(\cup _{x \in U} O_{r_{x}}(x)\). It is inside \(U\) for sure. For any \(x \in U\), \(x \in O_{r_{x}}(x)\) and is hence in the union, so the \(U\) is inside the union. Thus, \(U\) is exactly the union.
In a metric space a finite intersection of open sets is open.
Similar to the proof of Prop. 1. Let \(U\) be the finite intersection of open sets \(U_1, \dots , U_{n}\). Let \(x \in U\). Then \(x \in U_{i}\) for each \(i\). Since \(U_{i}\) is open, we have \(O_{r_{i}}(x) \subset U_{i}\) for some \(r_{i} > 0\). Taking \(r = \min_{i} (r_{i}) > 0\), we have \(O_{r}(x) \subset U\).
1.2.2. Interior
In a metric space, the interior of a set \(E\), denoted by \(E ^{\circ }\), is the set of all interior points of \(E\).
\(E ^{\circ }\) is always open.
For any \(x \in E ^{\circ }\), \(\exists \) a neighbourhood \(O_{r, E}(x)\) of \(x\) relative to \(E\) that is completely in \(E\). Consider \(O_{\frac{r}{2}, E ^{\circ }}(x)\) (relative to \(E ^{\circ }\)). We want to show that this set is completely inside \(E ^{\circ }\).
For any point \(y\) in this set, \(\forall z \in O_{\frac{r}{2}, E}(y)\), \(d(z, x) \leq d(z, y) + d(y, z) < r\), so \(z \in O_{r, E}(x) \subset E\). Thus, \(y \in E ^{\circ }\), as desired. Therefore, \(E ^{\circ }\) is open.
\(E ^{\circ }\) is the union of all open sets contained in \(E\). In particular, \(E ^{\circ }\) is the largest open set contained in \(E\).
By Thm. 1: Let a point \(x\). \(x \in \) the union of all open sets contained in \(E\) \(\iff \) \(x\) is in some open set in \(E\) \(\iff \) \(x\) is in some union of neighbourhoods (not nece. of \(x\)) \(\iff \) some neighbourhood of \(x\) is in \(E\) (easy) \(\iff \) \(x\) is an interior point of \(E\) \(\iff \) \(x \in E ^{\circ }\) by def.
In a metric space a subset \(E\) is open iff it consists of all its interior points, \(E = E ^{\circ }\).
If \(E\) is open, by definition, all points in \(E\) are interior points of \(E\).
Conversely, suppose \(x\) is an interior point of an open set \(E\), then a neighbourhood of \(x\) is inside \(E\), and hence \(x\) itself must also be in \(E\).
If \(G \subset E\) and \(G\) is open, then \(G \subset E ^{\circ }\)
1.3. Limit Points, Closed Sets, Closures
1.3.1. Limit Point
\(x\) is a limit point of \(E \subset X\) if every neighbourhood of \(x\) contains a point in \(E\) different than \(x\), i.e., \(\forall r > 0, (O_{r}(x) \cap E) \setminus \{ x \} \neq \emptyset\). The set of all limit points of \(E\) is denoted by \(E '\).
- Limit points of \(E\) are those can be approximated by points in \(E\) arbitrarily close: Analogous to limit in analysis, \(x\) is a limit point if no matter how small \(r\) is there is some point \(y \neq x\) in \(E\) with distance to \(x\) less than \(r\).
- A limit point is not necessarily in \(E\).
The opposite concept of limit point is isolated point: \(x\) is an isolated point of \(E\) if there exists a neighbourhood of \(x\) that does not contain any other points of \(E\), i.e., \(\exists r > 0, (O_{r}(x) \setminus \{ x \}) \cap E = \emptyset\).
If \(p \in X\) is a limit point of \(E \subset X\), then every neighbourhood of \(p\) contains infinitely many points of \(E\).
Suppose a neighbourhood \(O_{r}(p)\) contains finitely many points of \(E\) except possibly \(p\) itself, \(\{ q_1, \dots , q_{n} \}\). Take \(\hat{r} = \min_{i} d(q_{i}, p) > 0\). Then, \((O_{\hat{r}}(p) \setminus \{ p \}) \cap E = \emptyset\), which contradicts with \(p\) being a limit point.
A finite set has no limit point.
1.3.2. Closed Set
A subset \(E \subset X\) is closed if it contains all its limit points.
A subset of a metric space is closed iff its complement is open.
Let \(F\) be a set and \(E\) be its complement.
Assume \(F\) is closed. Let \(x \in E\), \(x \notin F\). Then \(x\) cannot be a limit point of \(F\). By definition, there exists a neighbourhood \(O\) of \(x\) s.t. \((O \setminus \{ x \}) \cap F = O \cap F = \emptyset\). Thus, \(O \subset E\). \(E\) is open.
Conversely, assume \(E\) is open. Let \(x\) be a limit point of \(F\). Then every neighbourhood of \(x\) intersects with \(F\) at some point other than \(x\). Thus, neighbourhoods of \(x\) cannot be completely inside \(E\) and hence \(x\) cannot be an interior point of \(E\). Since \(E\) is open, this means that \(x \notin E\), i.e., \(x \in F\).
By Thm. 1, Thm. 1, and Thm. 1, we have:
In a metric space any intersection of closed sets is closed.
In a metric space a finite union of closed sets is closed.
In a metric space the complement of the interior of a set is the closure of the complement of the set.
Hint: Prove the two directions.
A subset \(E \subset X\) is perfect if it is closed and all its points are its limit points (none of its points is isolated).
1.3.3. Closure
In metric space, the closure of a set \(E\) consists of all points in \(E\) as well as all limit points of \(E\), denoted by \(\cl{E} = E \cup E'\).
A point \(x\) lies in \(\cl{E}\) iff every neighbourhood of \(x\) intersects \(A\).
Proof: Note that \(x \in E'\) iff every neighbourhood of \(x\) intersects \(A\) in some points over than \(x\). Thus, this condition include also points that are in \(A\).
See another equivalent definition using convergence: Thm. 1.
Let \(X\) be a metric space and Let \(E \subset X\), then
- \(\cl{E}\) is closed,
- \(E\) is closed iff \(E = \cl{E}\),
- \(\cl{E}\) is the intersection of all closed sets containing \(E\). In particular, \(\cl{E}\) is the smallest closed set containing \(A\).
- We prove that \((\cl{E}) ^{C}\) is open. Let \(x \in (\cl{E}) ^{C}\), i.e., \(x \notin E' \cup E\). We want to show that there exists some neighbourhood of \(x\) \(\subset (\cl{E}) ^{C}\).
Not that \(x \notin E' \cup E \implies \exists O_{r'}(x), (O_{r'}(x) \cap E) \setminus \{ x \} = O_{r'}(x) \cap E = \emptyset\). We want to choose some \(r < r'\) such that \(O_{r}(x)\) does not intersect with \(E'\) either. Suppose \(O_{r'}(x) \cap E' \neq \emptyset\) and \(a\) is a point in it. Then, \(a \in E' \implies \) any neighbourhood of \(a\), especially \(O_{r' - d(x, a)}(a)\), intersects \(E\) in some point other than \(a\). Let \(b\) be that point. Then, \(d(x, b) \leq d(x, a) + d(a, b) < r'\), implying that \(O_{r'}(x)\) intersects \(E\) in \(b\), which is a contradiction.
- TODO
- TOOD
\((E ^{\circ }) ^{C} = \cl{E ^{C}}\).
Let \(X\) denote the metric space. By Thm. 1,
\begin{equation} \label{eq:20} (E ^{\circ }) ^{C} = \left(\bigcup _{O \subset E, O \text{ open}} O\right) ^{C} = \bigcap _{O \subset E, O \text{ open}} O ^{C} = \bigcap_{F \subset E ^{C}, F \text{ closed}} F. \end{equation}
1.3.4. Dense Sets
A subset \(E \subset X\) is dense in \(X\) if every point of \(X\) is a limit point of \(E\) or a point in \(E\), i.e., \(\cl{E} = X\).
1.4. TODO Compactness
Intuition: Stack Exchange
In a metric space we say a collection \(\mathcal{A}\) of subsets is a covering of a set \(E\) if \(E \subset \cup _{A \in \mathcal{A}} A\). It is called an open covering of \(E\) if its element are open subsets.
In a metric space we say a set \(K\) is compact (relative to the metric space) if every open covering of \(K\) has a finite subcollection that also covers \(K\).
Suppose \(X\) is a metric space and \(Y \subset X\) is a restricted metric space. Then, \(K\) is compact relative to \(X\) iff \(K\) is compact relative to \(Y\).
TODO
Any compact set is closed.
Let \(K\) be a compact set. We show that \(K ^{C}\) is open.
Let \(x\) be any point in \(K ^{C}\). Consider the open covering \(U = \cup _{p \in K} O_{\frac{1}{2} d(p, x)}(p)\) of \(K\) (can be easily verified). Since \(K\) is compact, there is a finite subcollection that is still of covering of \(K\): \(W = \cup _{i=1}^{n} O_{\frac{1}{2} d(p_{i}, x)} (p_{i})\). Let \(V = \cap _{i=1} ^{n} O_{\frac{1}{2} d(p_{i}, x)}(x)\). \(V\) is an open set. Since \(y \in V\) iff \(y \in O_{\frac{1}{2} d(p_{i}, x)}(x)\) for all \(i\), which implies that \(y \notin O_{\frac{1}{2}d(p_{i}, x)}(p_{i})\) for all \(i\). This means that \(V \cap W = \emptyset\), \(V \cap K = \emptyset\). Thus, \(x\) is contained in an open set in \(K ^{C}\), i.e., \(K\) is open.
Closed subsets of a compact set are compact.
TODO
If \(F\) is closed (not nece. in \(K\)) and \(K\) is compact, then \(F \cap K\) is compact.
TODO
Let \(\{ K_{i} \}\) be a collection of compact sets is a metric space \(X\) such that every finite intersection \(K_{i_1} \cap \dots K_{i_{n}}\) is non-empty, then \(\cap _{i} K_{i}\) is non-empty.
If \(\{ K_{n} \}\) is a sequence of descending compact sets (\(K_1 \supset K_2 \supset \dots \)), then \(\cap _{i=1} ^{\infty} K_{n} \neq \emptyset\).
Suppose \(E\) is an infinite subset of a compact set \(K\), then \(E\) has a limit point in \(K\).
A subset \(K\) of \(\mathbb{R} ^{n}\) is compact iff \(K\) is closed and bounded.
Every bounded infinite subset of \(\mathbb{R} ^{n}\) has a limit point.
Let \(P\) be a perfect set inside \(\mathbb{R} ^{n}\). Then \(P\) is uncountable.
1.5. Sequence, Convergence
A sequence \(\{ x_{n} \} \subset X\) converges to \(x \in X\), denoted by \(x_{n} \to x\), if \(\forall \epsilon > 0, \exists N \in \mathbb{N}\) s.t. \(\forall n > N, d(x, x_{n}) < \epsilon\).
The range of \(\{ x_{n} \}\) is the set of all values of \(\{ x_{n} \}\), i.e., \(\{ x \in X | \exists n \in \mathbb{N} , x = x_{n} \}\).
\(\{ x_{n} \}\) is bounded if the range of \(\{ x_{n} \}\) is bounded.
If \(x, x' \in X\), \(x_{n} \to x\), and \(x_{n} \to x'\), then \(x = x'\).
Hint: \(x\) and \(x'\) can be arbitrarily close.
In a compact metric space, every sequence has some convergent subsequence.
Every bounded sequence in \(\mathbb{R} ^{k}\) contains a convergent subsequence.
1.5.1. Convergence and Set Properties
Let \(A\) be a subset in a metric space \(X\). Then \(x\) is a limit point of \(A\) iff there exists a sequence \(\{ x_{n} \} \subset A\) such that \(x_{n} \neq x\) and \(x_{n} \to x\).
Let’s rewrite the condition for \(x\) being a limit point of \(A \subset X\): \(\forall \epsilon > 0, (O_{\epsilon}(x) \cap A) \setminus \{ x \} \neq \emptyset\). This is the same as \(\forall \epsilon > 0, \exists y \in A, y \neq x, d(x, y) < \epsilon\). It is easy to verify that this is equivalent to the statement in the proposition.
A point \(x\) lies in \(\cl{A}\) iff there exists a sequence of elements of \(A\) coverging to \(x\).
Fix a sequence \(\{ x_{n} \}\) in a metric space \(X\). Then the subsequential limits of \(\{ x_{n} \}\) form a closed subset of \(X\).
Let \(X\) be a metric space, \(K \subset X\) compact, \(F \subset X\) closed, \(K \cap F = \emptyset\). Then, there exists \(\delta > 0\) such that \(d(p, q) > \delta\) if \(p \in K, q \in F\).
This means that if \(K, F\) are disjoint then they are actually separated by some distance.
Suppose the opposite. We can take two sequence \(\{ p_{n} \} \subset K, \{ q_{n} \} \subset F\) that can be arbitrarily close. By Thm. 1, \(\{ p_{n} \}\) has some subsequence \(\{ p_{n_{k}} \}\) that converges to some \(p \in K\). Then, \(\{ q_{n_{k}} \}\) also converges to \(p\). Since \(F\) is closed, \(p\) is in \(F\), which contradicts with the condition that \(F \cap K = \emptyset\).
1.6. Completeness, Cauchy
1.6.1. Cauchy Sequence
A sequence \(\{ x_{n} \}\) in a metric space is called Cauchy if \(\forall \epsilon > 0, \exists N \in \mathbb{N} \) s.t. \(\forall n, m > N, d(x_{n}, x_{m}) < \epsilon\).
Let \(E\) be a nonempty subset in a metric space. Then the diameter of \(E\), denoted by \(\diam E\), is defined as \(\sup \{ d(x, y) : x, y \in E \}\).
A chain \(L\) is said to have the least-upper-bound property if every nonempty subset in \(L\) that is bounded above has a least upper bound in \(L\).
Suppose \(C\) is a chain with the least-upper-bound property. Suppose \(B \subset C\), \(B \neq \emptyset\), and \(B\) is bounded below. Let \(L\) be the set of all lower bounds of \(B\). Then, \(\alpha = \sup L\) exists in \(L\), and \(\alpha = \inf B\).
Since \(B\) is bounded below, \(L\) is non-empty. By the lub property of \(C\), \(L\) has a least upper bound. Let us denote it by \(\alpha\). Since every element in \(B\) is an upper bound of \(L\), \(\alpha \preceq\) all of them. Then, by the definition of \(L\), \(\alpha \in L\). Then, \(\alpha\) is the greatest among \(L\), i.e., the greatest lower bound of \(B\).
In a metric space, \(\diam E = \diam \cl{E}\).
TODO
In a metric space, every convergent sequence is Cauchy.
Triangle inequality.
In a metric space, if \(K_{n}\)’s are compact, \(K_{n} \supset K_{n+1} \supset \dots \), and \(\lim_{n \to \infty} \diam K_{n} = 0\), then \(\cap _{i=1} ^{\emptyset}\) contains totally one point.
TODO
- If \(X\) is a compact metric space, and \(\{ x_{n} \} \subset X\) is a Cauchy sequence, then \(\{ x_{n} \}\) converges to some \(x \in X\).
- In \(\mathbb{R} ^{k}\), every Cauchy sequence converges.
1.6.2. Completeness
We call a metric space complete if every Cauchy sequence in it converges in it.
Suppose \(\{ x_{n} \}\) is monotonic. Then \(\{ x_{n} \}\) is convergent iff \(\{ x_{n} \}\) is bounded.
TODO
1.6.3. Upper and Lower Limits
For a real sequence \(\{ s_{n} \}\), we define \(s_{n} \to + \infty\) if \(\forall M, \exists N_{M}\) s.t. \(n > N_{M} \implies s_{n} \geq M\). \(s_{n} \to - \infty\) is defined analogously.
Let \(\{ s_{n} \}\) be a sequence of real numbers. Let \(E\) be the set of all subsequential limits of \(\{ s_{n} \}\) including possibly \(\pm \infty\). The upper limit and lower limit of \(\{ s_{n} \}\) are denoted and defined as \(\limsup s_{n} = s ^{*} = \sup E, \liminf s_{n} = s_{*} = \inf E\).
Let \(\{ s_{n} \}\) and \(E\) be as before. Then \(s ^{*}\) is characterized by (the only number satisfying) the following two properties:
- \(s ^{*} \in E\),
- If \(x > s ^{*}\), then \(\exists N\) s.t. \(n > N \implies s_{n} < x\).
TODO
If \(s_{n} \preceq t_{n}\) for all \(n \geq N\), then \(\limsup s_{n} \preceq \limsup t_{n}, \liminf s_{n} \preceq \limsup t_{n}\).
1.7. Function, Continuity
1.7.1. Function Limit
Let \(X, Y\) be metric spaces, \(E \subset X\), \(p \in E'\), \(f: E \to Y\). Then \(f(x)\) has a limit \(q \in Y\) at \(p\), denoted by \(\lim_{x \to p} f(x) = q\), if \(\forall \epsilon > 0, \exists \delta > 0\) s.t. \(d_{X}(x, p) < \epsilon \implies d_{Y}(f(x), q) < \epsilon\).
Let \(X, Y, E, p, f\) be as before. Then \(\lim_{x \to p} f(x) = q\) iff \(\lim_{n \to \infty} f(x_{n}) = q\) for every sequence \(\{ x_{n} \} \subset E\) such that \(x_{n} \neq p\) and \(\lim_{n \to \infty} x_{n} = p\).
Note that \(p \in E'\) implies that the existence of such a sequence \(\{ x_{n} \}\).
Suppose \(\lim_{x \to p} f(x) = q\) holds, it is easy to prove the consequence.
Suppose \(\lim_{x \to p} f(x) = q\) does not hold, we want to show that \(f(x_{n}) \not \to q\) for some sequence satsifying the conditions. \(\exists \epsilon > 0\), \(\forall \delta > 0\), \(\exists x \in E\), \(d_{X}(x, p) < \delta\) and \(d_{Y}(f(x), q) \geq \epsilon\). For each \(n\), taking \(\delta_{n} = \frac{1}{n}\), there exists a corresponding \(x_{n}\) with the above properties. Thus, \(\{ x_{n} \} \to p\) but \(f(x_{n}) \not \to q\).
Let everything as before (especially \(p \in E'\)). If \(f(x)\) has a limit at \(p\), then the limit is unique.
The sequence formulation of function limit implies that the existence of some convergent sequence \(f(x_{n})\). By the uniqueness of sequence limit, the function limit is also unique.
Let \(X\) be a metric space, \(E \subset X\), \(p \in E'\), \(f, g: E \to \mathbb{C} \), \(\lim_{x \to p} f(x) = a\) and \(\lim_{x \to p} g(x) = b \), then \(\lim_{x \to p} f(x) + - \cdot /g(x) = a + - \cdot / b\) (for quotient, provided that \(b \neq 0, g(x) \neq 0\)).
From the arithmetics rules for sequential limits.
1.7.2. TODO Continuity
Let \(X, Y\) be two metric spaces, \(E \subset X\), \(p \in X\) (not nece. in \(E\)), \(f: X \to Y\). Then \(f\) is continuous at \(p\) if \(\forall \epsilon > 0, \exists \delta > 0\) s.t. \(d_{X}(x, p) < \delta \implies d_{Y}(f(x), f(p)) < \epsilon\).
\(f\) is said to be continuous on \(E\) if it is continuous at every point in \(E\).
- If \(p \in E'\), the definition is equivalent to \(\lim_{x \to p} f(x) = f(p)\). (We did not define the function limit for \(p \notin E'\)).
The function \(f: E \to Y\) is always continuous at an isolated point of \(E\).
Let \(p\) be an isolated point of \(E\). Pick \(\delta > 0\) s.t. the only \(x \in E\) for which \(d_{X}(x, p) < \delta\) is \(p\). Then \(f\) is continuous at \(p\).
Let \(X, Y, Z\) be metric spaces, \(f: E \to Y, g: f(E) \to Z\). If \(f\) is continuous at \(p\) and \(g\) is continuous at \(f(p)\), then \(h = f \circ g\) is continuous at \(p\).
Do the epsilon-delta twice. Let \(\epsilon > 0\). \(g\) is continuous at \(f(p)\), so \(\exists \eta > 0\) s.t. \(d_{Y}(y, f(p)) < \eta \implies d_{E}(g(y), g(f(p))) < \epsilon\). \(f\) is continuous at \(p\), so \(\exists \delta > 0\) s.t. \(d_{X}(x, p) < \delta \implies d_{Y}(f(x), f(p)) < \eta\). This means that \(d_{X}(x, p) < \delta \implies d_{E}(h(x), h(p)) < \epsilon\).
Let \(X, Y\) be metric spaces. Then a map \(f: X \to Y\) is continuous on \(X\) iff \(f ^{-1}(O) \subset X\) is open for every open set \(O \subset Y\).
TODO
Let \(X, Y\) be metric spaces. Then \(f: X \to Y\) is continuous iff \(f ^{-1}(F)\) is closed for any closed set \(F \subset Y\).
TODO
Let \(f, g\) be complex-valued function on a metric space. If \(f, g\) are continuous, then \(f + g, fg, f / g\) are continuous.
NOPROF
1.7.3. TODO Continuity and Set Properties
Let \(X, Y\) be metric spaces, \(f: X \to Y\) a continuous function. Then for every subset \(A \subset X\), \(f(\cl{A}) = \cl{f(A)}\).
\(y \in f(\cl{A}) \iff \exists x \in \cl{A}, y = f(x) \iff \exists \{ x_{n} \} \subset A, x_{n} \to x, y = f(x)\).
By the continuity of \(f\), this implies \(\exists \{ x_{n} \} \subset A, f(x_{n}) \to f(x) = y\). Thus, \(\exists \{ y_{n} \} \subset f(A), y_{n} \to y\). This is exactly the condition for \(y\) being in \(\cl{f(A)}\).
A mapping \(f: E \to \mathbb{R} ^{k}\) is said to be bounded if \(\exists M\) s.t. \(\forall x \in E, \| f(x) \| \leq M\).
Let \(X, Y\) be metric spaces, \(X\) compact, \(f: X \to Y\). If \(f\) is continuous, then \(Y\) is compact.
TODO
Let \(X\) be a metric space, \(f: X \to \mathbb{R} ^{k}\) continuous. Then \(f(X)\) is closed and bounded.
TODO
Let \(X\) be a compact metric space, \(f: X \to \mathbb{R} \) coninuous. Let \(M = \sup_{p \in X} f(p)\), \(m = \inf_{p \in X} f(p)\). Then \(\exists p, q \in X\) s.t. \(f(p) = M, f(q) = m\).
TODO
Let \(X, Y\) be metric spaces, \(X\) compact, \(f: X \to Y\) a continuous bijection. Then, \(f ^{-1}: Y \to X\) is continous.
TODO
1.7.4. TODO Uniform Continuity
Let \(X, Y\) be metric spaces. A mapping \(f: X \to Y\) is said to be uniformly continuous if \(\forall \epsilon > 0, \exists \delta > 0\) s.t. \(d_{X}(x, y) < \delta \implies d_{Y}(d(x), d(y))\).
Let \(X, Y, f\) be as before. If \(X\) is compact, then \(f\) is uniformly continuous.
Let \(\epsilon > 0\). We want to find a uniform \(\delta > 0\) s.t. \(d_{X}(p, q) < \delta \implies d_{Y}(f(p), f(q)) < \epsilon\). By the continuity, for every \(p \in X\), \(\exists \delta_{p} > 0\) s.t. \(d_{X}(p, q) < \delta_{p} \implies d_{Y}(f(p), f(q)) < \epsilon\). Let \(J(p) := \{ q \in X: d_{X}(p, q) < \frac{1}{2} \delta_{p} \}\). Then \(\cup _{p \in X} J(p)\) is an open cover of \(X\). By the compactness of \(X\), there exists a finite sub collection \(\cup _{i=1} ^{n} J(p_{i})\) which is also an open cover of \(X\). Take \(\delta = \frac{1}{2} \min _{i=1} ^{n} \{ \phi(p_{i}) \} > 0\), where the min makes sense and is \(> 0\) because the collection is finite. Then, for any \(p, q\) with \(d_{X}(p, q) < \delta\), \(p\) must be in some \(J(p_{i})\), and hence both \(p, q\) are within a distance of \(\delta_{p_{i}})\) to \(p_{i}\), implying that \(d_{Y}(p, q) < d_{Y}(p, p_{i}) + d_{Y}(p_{i}, q) = \epsilon\).
A uniform continuous function carries Cauchy sequences into Cauchy sequences.
Proof: Easy.
2. Topology
A topology on a set \(X\) is a collection \(\mathcal{T}\) of subsets of \(X\) satisfying:
- \(\emptyset, X \in \mathcal{T}\),
- Any union of elements of \(\mathcal{T}\) is in \(\mathcal{T}\),
- Any finite intersection of elements of \(\mathcal{T}\) is in \(\mathcal{T}\).
A topological space is an ordered pair \((X, \mathcal{T})\) consisting of a set \(X\) and a topology \(\mathcal{T}\) on \(X\).
Given a topological space \((X, \mathcal{T})\), a subset \(U\) of \(X\) is an open set of \(X\) if \(U \in \mathcal{T}\).
Let \(X\) be any set.
- The collection of all subsets of \(X\) is a topology on \(X\), called the discrete topology,
- The collection consisting of \(X\) and \(\emptyset\) only is a topology, called the indiscrete topology.
- \(\mathcal{T} = \{ O \subset X: X - O \text{ is either finite or all of } X \}\) is a topology, called the finite complement topology.
- \(\mathcal{T} = \{ O \subset X: X - O \text{ is either countable or all of } X \}\) is a topology, called the countable complement topology.
Given two topologies \(\mathcal{T}, \hat{\mathcal{T}}\) on a set \(X\). Then we say \(\hat{\mathcal{T}}\) is finer/stronger than \(\mathcal{T}\), denoted by \(\hat{\mathcal{T}} \succeq \mathcal{T}\) if \(O \subset \mathcal{T} \implies O \subset \hat{\mathcal{T}}\). Or equivalently, \(\mathcal{T}\) is coarser/weaker than \(\hat{\mathcal{T}}\)
2.1. Basis for a Topology
Let \(X\) be a set. A basis for a topology on \(X\) is a collection \(\mathcal{B}\) of subsets of \(X\) such that
- For each \(x \in X\), \(\exists B \in \mathcal{B}, x \in B\).
- If \(x\) belongs to the intersection of two basis elements \(B_1\) and \(B_2\), then there is a basis element \(B_3\) s.t. \(x \in B_3 \subset B_1 \cap B_2\).
As we will show that such a basis generates a (unique) topology, say \(\mathcal{T}\), on \(X\), we call \(\mathcal{B}\) a basis for the topology \(\mathcal{T}\) on \(X\) (provided \(\mathcal{T}\)).
Let \(\mathcal{B}\) be a basis for a topology on \(X\). Then the topology \(\mathcal{T}\) generated by \(\mathcal{B}\) is the defined as
\begin{equation} \label{eq:1} \mathcal{T} = \{ U \subset X: \forall x \in U, \exists B \in \mathcal{B}, x \in B \subset U \}. \end{equation}The collection \(\mathcal{T}\) generated by the basis \(\mathcal{B}\) is in fact a topology on \(X\):
TODO
Let \((X, \mathcal{T})\) be a topological space. Suppose \(\mathcal{C}\) is a collection of open sets of \(X\) such that for each open set \(O\) of \(X\):
\begin{equation} \label{eq:9} \forall x \in O, \exists C \in \mathcal{C}, x \in C \subset U, \end{equation}In other words, every open set in \(\mathcal{T}\) is in the topology generated by the “basis” \(\mathcal{C}\) (though not sure for now).
Then \(\mathcal{C}\) is a basis for the topology \(\mathcal{T}\).
TODO
Let \(X\) be a set, \(\mathcal{B}\) a basis for a topology on \(X\), and \(\mathcal{T}\) the topology generated by \(\mathcal{B}\). Then \(\mathcal{T}\) equals to collection of all unions of elements of \(\mathcal{B}\).
TODO
Let \(\mathcal{B}\) and \(\mathcal{B}'\) be bases for the topologies \(\mathcal{T}\) and \(\mathcal{T}'\), resp., on \(X\). Then \(\mathcal{T}'\) is finer than \(\mathcal{T}\) iff for each basis element \(B \in \mathcal{B}\):
\begin{equation} \label{eq:4} \forall x \in B, \exists B' \in \mathcal{B}', x \in B' \subset B. \end{equation}In other words, every basis element in \(\mathcal{B}\) is in the topology generate by \(\mathcal{B}'\).
TODO
Consider \(\mathbb{R} \).
- Let \(\mathcal{B}\) be the collection of all open intervals \((a, b) = \{ x | a < x < b \}\). The topology generated by \(\mathcal{B}\) is called the standard topology on \(\mathbb{R} \).
- Let \(\mathcal{B}_{l}\) be the collection of all half-open intervals of the form \([a, b) = \{ x | a \leq x < b \}\). The topology generated by \(\mathcal{B}_{l}\) is called the lower limit topology on \(\mathbb{R} \). When \(\mathbb{R} \) is equipped with the lower limit topology, we denote it by \(\mathbb{R} _{l}\).
- Let \(K = \{ 1 / n, n \in \mathbb{Z} _{+} \}\). \(\mathcal{B}_{K}\) be the collection of all open intervals, along with all sets of the form \((a, b) - K\). The topology generated by \(\mathcal{B}_{K}\) is called the \(K\)-topology on \(\mathbb{R} \). When \(\mathbb{R} \) is equipped with the \(K\)-topology, we denote it by \(\mathbb{R} _{K}\).
Properties:
- \(\mathbb{R} _{l}\) and \(\mathbb{R} _{K}\) are not comparable (check by the bases).
2.1.1. Subbasis
A subbasis for a topology on \(X\) is a collection of subsets of \(X\) whose union equals \(X\).
The collection \(\mathcal{B}\) of all finite intersections of elements of a subbasis \(\mathcal{S}\) is a basis. The topology generated by a subbasis \(\mathcal{S}\) is defined to be the topology generated by the basis \(\mathcal{B}\), which is the collection of all unions of finite intersections of elements of \(\mathcal{S}\).
TODO
2.1.2. A Family of Topologies
Let \(\{ \mathcal{T}_{\alpha} \}_{\alpha}\) be a family of topologies on \(X\). Then
- \(\cap_{\alpha} \mathcal{T}_{\alpha}\) is a topology on \(X\),
- \(\cup_{\alpha} \mathcal{T}_{\alpha}\) is not generally a topology on \(X\).
Proof: easy.
Let \(\{ \mathcal{T}_{\alpha} \}_{\alpha}\) be a family of topologies on \(X\). Then
- \(\cap_{\alpha} \mathcal{T}_{\alpha}\) is the unique largest topology contained in all \(\mathcal{T}_{\alpha}\).
- The topology generated by the subbasis \(\mathcal{S} = \cup _{\alpha} \mathcal{T}_{\alpha}\) is the unique smallest topology containing all \(\mathcal{T}_{\alpha}\).
For the second one, any topology \(\mathcal{T} '\) containing all \(\mathcal{T}_{\alpha}\) would have to contain all finite intersections of elements of \(\cup _{\alpha} \mathcal{T}_{\alpha}\). The collection of all finite intersections is the basis \(\mathcal{B}\) corresponding to the subbasis \(\mathcal{S}\). Consequently, \(\mathcal{T}'\) is finer than the topology generated by \(\mathcal{B}\).
If \(\mathcal{A}\) is a basis or a subbasis for a topology on \(X\), then the topology generated by \(\mathcal{A}\) is equal to the intersection of all topologies on \(X\) that contain \(\mathcal{A}\).
2.2. Order Topology
Let \(X\) be a totally ordered set. Given elements \(a\) and \(b\) of \(X\) such that \(a \prec b\), the following four subses are called the intervals determined by \(a\) and \(b\): \((a, b) = \{ x | a < x < b \}\), \((a, b]\), \([a, b)\), \([a, b]\) (analogously defined).
Let \(X\) be a non-empty totally ordered set. Let \(\mathcal{B}\) be the collection of all sets of the following types:
- All open intervals \((a, b)\) in \(X\).
- If \(X\) has the bottom element, \(a_0\): All intervals of the form \([a_0, b)\).
- If \(X\) has the top element, \(b_0\): All intervals of the form \((a, b_0]\).
Then \(\mathcal{B}\) is a basis for a topology on \(X\). The topology generated by \(\mathcal{B}\) is called the order topology.
Let \(X\) be an ordered set (not necessarily totally ordered), and \(a\) is an element of \(X\), the following four subsets are called the rays determined by \(a\): \((a, + \infty) = \{ x | x > a \}\), \((-\infty, a)\), \([a, + \infty)\), \((-\infty, a]\) (analogously defined).
2.3. Product Topology
Let \(X, Y\) be topological spaces. The product topology on \(X \times Y\) is the topology having as basis the collection \(\mathcal{B} = \{ U \times V | U \text{ open in } X, V \text{ open in } Y\}\).
\(\mathcal{B}\) is in fact a basis.
TODO
If \(\mathcal{B}\) is a basis for the topology of \(X\) and \(\mathcal{C}\) a basis for the topology of \(Y\), then the collection
\begin{equation} \label{eq:5} \mathcal{D} = \{ B \times C | B \in \mathcal{B}, C \in \mathcal{C} \} \end{equation}is a basis for the topology of \(X \times Y\).
Let \(\pi_{1}: X \times Y \to X\) be defined by
\begin{equation} \label{eq:6} \pi_{1}(x, y) = x. \end{equation}Let \(\pi_{2}: X \times Y \to Y\) be defined by
\begin{equation} \label{eq:7} \pi_{2}(x, y) = y. \end{equation}The maps \(\pi_{1}, \pi_{2}\) are called the projections of \(X \times Y\) onto its first and second factors, resp.
The collection
\begin{equation} \label{eq:8} \mathcal{S} = \{ \pi_{1} ^{-1}(U) | U \text{ open in } X \} \cup \{ \pi_{2} ^{-1}(V) | V \text{ open in } Y \} \end{equation}is a subbasis for the product topology on \(X \times Y\).
TODO
2.4. Subspace Topology
Let \(X\) be a topological space with topology \(\mathcal{T}\). If \(Y\) is a subset of \(X\), the collection
\begin{equation} \label{eq:2} \mathcal{T}_{Y} = \{ Y \cap O | O \in \mathcal{T} \} \end{equation}is a topology on \(Y\), called the subspace topology (of \(Y\) relative to \(X\)). \((Y, \mathcal{T}_{Y})\) is called a subspace of \((X, \mathcal{T})\).
TODO
If \(\mathcal{B}\) is a basis for the topology of \(X\) then the collection
\begin{equation} \label{eq:3} \mathcal{B}_{Y} = \{ B \cap Y | B \in \mathcal{B} \} \end{equation}is a basis for the subspace topology on \(Y\).
TODO
Let \(Y\) be a subspace of \(X\). If \(U\) is open in \(Y\) and \(Y\) is open in \(X\), then \(U\) is open in \(X\).
\(U\) is open in \(Y\) iff \(U = Y \cap O\) for some \(O\) open in \(X\). Since \(Y\) and \(O\) are both open in \(X\), \(Y \cap O\) is open in \(X\).
Let \(X\) be a topological space, \(Y\) a subspace of \(X\), \(A \subset Y\). Then the subspace topology of \(A\) relative to \(Y\) is equal to the subspace topology of \(A\) relative to \(X\).
Hint: Express \(A = A \cap Y\).
Let \((X, \mathcal{T}(X)), (Y, \mathcal{T}(Y))\) be two topological spaces, and \((A, \mathcal{T}_{A}(X)), (B, \mathcal{T}_{B}(Y))\) are their subspaces, resp. Then the following two topologies are the same:
- The product topology \(\mathcal{T}(A \times B)\) on \(A \times B\). (first subspace, then product)
- The subspace topology \(\mathcal{T}_{A \times B}\) relative to \(X \times Y\). (first product, then subspace)
TODO
2.5. Closed Sets, Limit Points, Closure
In a topological space, a neighbourhood of an element \(x\) is an open set containing \(x\).
Let \(A\) be a subset of the topological space \(X\). Then \(A\) is open iff for every point \(x \in A\) there is a neighbourhood of \(A\) completely in \(A\).
If \(A\) is open, then just take the neighbourhood as \(A\).
Conversely, for every \(x \in A\), let \(U_{x}\) be the neighbourhood. Then \(A = \cup _{x \in A} U_{x}\). By definition, \(A\) is open.
Let \(X\) be a topological space and \(A\) be a subset. Then a point \(x \in X\) is called a limit point of \(A\) if for every neighbourhood of \(x\) intersects \(A\) in some point other than \(x\).
The set of all limit points of \(A\) is denoted by \(A'\).
2.5.1. Closed Set
A subset \(A\) of a topological space \(X\) is said to be closed if \(X - A\) is open.
Let \(X\) be a topological space. Then the following condiions hold:
- \(\emptyset, X\) are closed.
- Arbitrary intersections of closed sets are closed.
- Finite unions of closed sets are closed.
The following properties analogous to open sets holds for closed sets:
Let \(Y\) be a subspace of \(X\). Then a set \(A\) is closed in \(Y\) iff it equals a closed set \(X\) with \(Y\).
Let \(Y\) be a subspace of \(X\). If \(A\) is closed in \(Y\) and \(Y\) is closed in \(X\), then \(A\) is closed in \(X\).
2.5.2. Closure
Let \(X\) be a topological space and \(A\) be a subset. The closure of \(A\) is defined as \(\cl{A} = A \cup A'\).
Compared to the definition of neighbourhood in metric space, the neighbourhood here only has to be an open set, while the neighbourhood there has to be a basic open set. However, we will see the equivalence (up to open and closed sets) in the following:
Let \(A\) be a subset of the topological space \(X\). Then
- \(x \in \cl{A}\) iff every neighbourhood of \(x\) intersects with \(A\).
- Furthermore, suppose the topology of \(X\) is given by a basis \(\mathcal{B}\), then \(x \in \cl{A}\) iff every basis element \(B \in \mathcal{B}\) containing \(x\) intersects \(A\).
TODO
Let \(Y\) be a subspace of \(X\) and \(A\) be a subset of \(Y\). Let \(\cl{A}_{X}\) denote the closure of \(A\) in \(X\) and \(\cl{A}_{Y}\) the closure of \(A\) in \(Y\). Then \(\cl{A}_{Y} = \cl{A}_{X} \cap Y\).
That is equivalent to \(A'_{Y} = A'_{X} \cap Y\) (the set of limit points in \(Y\), the set of limit points in \(X\)).
By Thm. 1, \(x \in A'_{Y}\) iff \(\forall B \in \mathcal{B}_{Y}, x \in B \implies B \cap A \setminus \{ x \} \neq \emptyset\). By Prop. 1, this means that \(\forall B \in \mathcal{B}_{X}, x \in B \cap Y \implies B \cap Y \cap A \setminus \{ x \} \neq \emptyset\), which is exactly the condition for \(x \in A'_{X} \cap Y\).
Let \(A\) be a subset of the topological space \(X\). Then
- \(x\)
2.6. Hausdorff Space
A topological space \((X, \mathcal{T})\) is called a Hausdorff space (in the topology \(\mathcal{T}\)) if for each pair \(u, v\) of \(X\), there exist neighbourhoods \(U, V\) of \(u, v\), resp., that are disjoint.
A set \(X\) is said to satisfy the T2 axiom if it is Hausdorff.
A set \(X\) is said to satisfy the T1 axiom if all finite point sets of \(X\) are closed.
T2 \(\implies\) T1, i.e., every finite point set in a Hausdorff space \(X\) is closed.
Let \(X\) be a space satisfying the T1 axiom. Let \(A\) be a subset of \(X\). Then the point \(x\) is a limit point of \(A\) iff every neighborhood of \(x\) contains infinitely many points of \(A\).
If \(X\) is a Hausdorff space, then a sequence of points of \(X\) converges to at most one point of \(X\).
- Every totally ordered set is a Hausdorff space in the order topology (i.e., open sets are defined by the order topology).
- The product of two Hausdorff spaces is a Hausdorff space.
- A subspace of a Hausdorff space is a Hausdorff space.
2.7. DONE Continuous Functions
Let \((X, \mathcal{J}), (Y, \mathcal{K})\) be topological spaces. A function \(f: X \to Y\) is said to be \((\mathcal{J}, \mathcal{K})\)-continuous if for each \(V \in \mathcal{K}\), the pre-image \(f ^{-1}(V)\) is in \(\mathcal{J}\).
We use \((\mathcal{J}, \mathcal{K})\) to emphasize that the continuity depends not only upon the function but also on the topologies.
Let \(\mathcal{B}\) be a basis for \(\mathcal{K}\). Then \(f\) is continuous if the pre-image of every basis element in \(\mathcal{B}\) is in \(\mathcal{J}\).
Any open set \(V\) in \(Y\) can be writen as a union of basis lements: \(V = \cup _{\alpha} B_{\alpha}\). Then \(f ^{-1}(V) = \cup _{\alpha} f ^{-1}(B_{\alpha})\).
Let \(\mathcal{S}\) be a subbasis for \(\mathcal{K}\). Then \(f\) is continuous if pre-image of every subbasis element in \(\mathcal{S}\) is in \(\mathcal{J}\).
Any basis element \(B\) can be written as a finite intersections of subbasis elements: \(B = S_1 \cap \dots \cap S_{n}\). Then \(f ^{-1}(B) = f ^{-1}(S_1) \cap \dots \cap f ^{-1}(S_{n})\).
Let \(X, Y\) be topological spaces. Let \(f: X \to Y\). Then the followings are equivalent:
- \(f\) is continuous.
- For every subset \(A\) of \(X\), \(f(\cl{A}) \subset \cl{f(A)}\). (Recall Thm. 1)
- For every closed set \(B\) of \(Y\), the pre-image \(f ^{-1}(B)\) is closed in \(X\).
- For each \(x \in X\) and each neighbourhood \(V\) of \(f(x)\), there is a neighbourhood \(U\) of \(x\) s.t. \(f(U) \subset V\).
We show that (1) \(\implies \) (2) \(\implies \) (3) \(\implies \) (1) and that (1) \(\iff \) (2).
(1) \(\implies \) (2): Let \(x \in \cl{A}}\). By Thm. 1, to show that \(f(x) \in \cl{f(A)}\), we need to show that every neighbourhood of \(f(x)\) intersects with \(f(A)\). Let \(V\) be a neighbourhood of \(f(x)\). Then \(f ^{-1}(V)\) is open in \(X\) and contains \(x\), i.e., a neighbourhood of \(x\). By \(x \in \cl{A}\) and again Thm. 1, \(f ^{-1}(V)\) intersects \(A\) in some point \(y\) (does not matter if \(y = x\)). Thus, \(f(y) \in V \cap f (A)\), as desired.
(2) \(\implies \) (3): Let \(B\) be a closed set in \(Y\). We want to show that \(f ^{-1}(B)\) is closed, i.e., \(f ^{-1}(B) = \cl{f ^{-1}(B)}\). By (2), we have \(f(\cl{f ^{-1}(B)}) \subset \cl{f(f ^{-1}(B))} = \cl{B} = B\). Thus, \(f ^{-1}(f(\cl{f ^{-1}(B)})) = \cl{f ^{-1}(B)} \subset f ^{-1}(B)\), so that \(\cl{f ^{-1}(B)} = f ^{-1}(B)\), as desired.
(3) \(\implies \) (1): Let \(V\) be an open set in \(Y\). Then \(B = Y - V\) is closed, and hence \(f ^{-1}(B)\) is closed. \(f ^{-1}(V) = X - f ^{-1}(B)\) is open.
(1) \(\implies\) (4): Let \(x \in X\) and \(V\) be a neighbourhood of \(f(x)\). Then \(f ^{-1}(V) \ni x\) is a neighbourhood of \(x\). Take \(U = f ^{-1}(V)\).
(4) \(\implies\) (1): Let \(V\) be an open set in \(Y\). For each \(x \in f ^{-1}(V)\), \(f(x) \in V\). By (4), there exists a neighbourhood \(U_{x}\) of \(x\) s.t. \(f(U_{x}) \subset V\). \(f ^{-1}(V) = \cup _{x \in f ^{-1}(V)} U_{x}\).
If the condition in (4) holds for the point \(x \in X\), we say \(f\) is continuous at \(x\).
2.7.1. DONE Homeomorphisms
Let \(X, Y\) be topological spaces. Let \(f: X \to Y\) be a bijection. If both the function \(f\) and the inverse function \(f ^{-1}\) are continuous, then \(f\) is called a homeomorphism.
In there exists a homeomorphism \(: X \to Y\), we say \(X\) and \(Y\) are homeomorphic.
Let \(X, Y\) be topological spaces. Let \(f: X \to Y\) be an injective continuous map. If the bijective function \(f': X \to f(X)\) obtained by restricting the range of \(f\) is a homeomorphism, then the map \(f: X \to Y\) is called a topological imbedding of \(X\) in \(Y\).
2.7.2. DONE Constructing Continuous Functions
Let \(X, Y, Z\) be topological spaces.
- Constant function: If \(f: X \to Y\) is defined by \(f(x) = y_0 \in Y\) for all \(x \in X\), then \(f\) is continuous.
- Inclusion: If \(A\) is a subspace of \(X\), the inclusion function \(j: A \to X\) defined by \(f(a) = a\) for all \(a \in A\) is continuous.
- Composites: If \(f: X \to Y\) and \(g: Y \to Z\) are continuous, then \(g \circ f: X \to Z\) is continuous.
- Restricting the domain: If \(f: X \to Y\) is continuous and \(A\) is a subspace of \(X\), then the restricted function \(f|_{A}: A \to Y\) is continuous.
- Restricting or expanding the range: Let \(f: X \to Y\) be continuous.
- If \(Z \supset f(X)\) is a subspace of \(Y\), then the function \(g: X \to Z\) obtained by restricting the range of \(f\) is continuous.
- If \(Z\) is a space having \(Y\) as a subspace, then the function \(h: X \to Z\) obtained by expanding the range of \(f\) is continuous.
- Local formulation of continuity: The map \(f: X \to Y\) is continuous if \(X\) can be written as the union of open sets \(U_{\alpha}\) such that \(f|_{U_{\alpha}}\) os continuous for each \(\alpha\).
- Let \(V\) be open in \(Y\). If \(V \ni x\), \(f ^{-1}(x) = X\); otherwise, \(f ^{-1}(x) = \emptyset\).
- Let \(V\) be open in \(X\). Then \(j ^{-1}(V) = V \cap A\), which is open.
- Let \(V\) be open in \(Z\). Then \(g ^{-1}(U)\) is open in \(Y\) and \(f ^{-1}(g ^{-1}(U))\) is open in \(X\). \((g \circ f) ^{-1}(U) = f ^{-1}(g ^{-1}(U))\).
- \(f_{|A}\) is the composition of the inclusion map \(j: A \to X\) and \(f\).
If \(f(X) \subset Z \subset Y\): Let \(V\) be open in \(Z\). Then \(V = Z \cap U\) for some open set \(U\) of \(Y\). Since \(f ^{-1}\) and \(g ^{-1}\) coincide on \(f(X)\), \(g ^{-1}(V) = g ^{-1}(V \cap f(X)) = f ^{-1}(V \cap f(X)) = f ^{-1}(U \cap f(X)) = f ^{-1}(U)\), which is open.
If \(Y \subset Z\): Then \(g\) is the composition of the inclusion map \(j: Y \to Z\) and \(f\).
- Let \(V\) be open in \(Y\). Then \(f ^{-1}(V) \cap U_{\alpha} = (f|_{U_{\alpha}}) ^{-1}(V)\) for all \(\alpha\). Thus, \(f ^{-1}(V) = \cup _{\alpha}(f ^{-1}(V) \cap U_{\alpha})\) is open.
Let \(X = A \cup B\), where \(A\) and \(B\) are closed in \(X\). Let \(f: A \to Y\) and \(g: B \to Y\) be continuous. If \(f(x) = g(x)\) for every \(x \in A \cap B\), then \(f\) and \(g\) combine to give a continuous function \(h: X \to Y\), defined by setting \(h(x) = f(x)\) if \(\)
P.S., if \(A\) and \(B\) are closed, the same result can be implied by the local continuity.
Use the definition of contiuity by closed sets.
2.8. Product Topology (General)
Generalize indices, coordinates from \(\mathbb{Z} _{+}\) to an arbitrary set.
Let \(J\) be an index set. Given a set \(X\), a \(J\)-tuple of elements of \(X\) is a function: \(\vec{x}: J \to X\), where we use the same font as for vectors. For an element \(\alpha\) of \(J\), we denote the value of \(\vec{x}\) at \(\alpha\) by \(x_{\alpha}\) (again the same as vectors, rather than \(\vec{x}(\alpha)\)) and call it the \(\alpha\)-th coordinate of \(\vec{x}\). And the function \(\vec{x}\) itself can also be denoted by \((x_{\alpha})_{\alpha \in J}\). The set of all possible \(J\)-tuples of elements of \(X\) is denoted by \(X ^{J}\).
Let \(\{ A_{\alpha} \}_{\alpha \in J}\) be an indexed family of sets. The cartesian product of this family is denoted and defined by
\begin{equation} \label{eq:10} \prod_{\alpha \in J} A_{\alpha} := \{ \vec{x}: J \to \cup _{\alpha \in J} A_{\alpha} | \forall \alpha \in J, x_{\alpha} \in A_{\alpha}\}. \end{equation}E.g., \(\prod_{\alpha \in J} X = X ^{J} \).
Let \(\{ X_{\alpha} \}_{\alpha \in J}\) be an indexed family of topological spaces. Then
\begin{equation} \label{eq:11} \mathcal{B}_{\text{box}} := \left\{ \prod_{\alpha \in J} U_{\alpha} | \forall \alpha \in J, U_{\alpha} \text{ is open in } X_{\alpha} \right\} \end{equation}is a basis for a topology on the product space \(\prod_{\alpha \in J} X_{\alpha} \). The topology generated by \(\mathcal{B}_{\text{box}}\) is called the box topology.
TODO
Let \(\{ A_{\alpha} \}_{\alpha \in J}\) be an indexed family of sets. For each \(\beta \in J\), the function \(\pi_{\beta}: \prod_{\alpha \in J} A_{\alpha} \to A_{\beta} \) assigning to each element of the cartesian product its \(\beta\)-th coordinate,
\begin{equation} \label{eq:12} (x_{\alpha})_{\alpha \in J} \overset{\pi_{\beta}}{\mapsto} x_{\beta}, \end{equation}is called the projection mapping associated with the index \(\beta\).
The following definition is comparable with Thm. 1.
Let
\begin{equation} \label{eq:13} \mathcal{S} = \{ \pi_{\beta} ^{-1} (U_{\beta}) | \beta \in J, U_{\beta} \text{ open in } X_{\beta} \}. \end{equation}Then \(\mathcal{S}\) is a subbasis of the product space \(\prod_{\alpha \in J} X_{\alpha}\), and the topology generated by \(\mathcal{S}\) is called the product topology.
Let \((x_{\alpha})_{\alpha \in J}\) be an arbitrary element in \(\prod_{\alpha \in J} X_{\alpha}\). We want to show that there exist a \(\beta \in J\) and an open subset \(U_{\beta}\) in \(X_{\beta}\) s.t. \((x_{\alpha})_{\alpha \in J} \in \pi_{\beta} ^{-1}(U_{\beta})\). We can simply choose any \(\beta\) and \(U_{\beta} = X_{\beta}\).
The product topology on \(\prod_{\alpha \in J} X_{\alpha}\) has the basis:
\begin{equation} \label{eq:16} \mathcal{B}_{\text{product}} := \left\{ \left. \prod_{\alpha \in J} U_{\alpha} \right| U_{\alpha} \text{ open in } X_{\alpha} \text{ for finitely many } \alpha, U_{\alpha} = X_{\alpha} \text{ otherwise }\right\} \end{equation}Consider any finite intersetion of elements belonging to the same one of the sets \(\mathcal{S}_{\beta}\): \(\pi_{\beta} ^{-1}(U_{\beta}, 1) \cap \dots \cap \pi_{\beta} ^{-1}(U_{\beta}, n) = \mu_{\beta} ^{-1}(U_{\beta, 1} \cap \dots \cap U_{\beta, n})\). Since the finite intersection of open sets in \(X_{\beta}\) is still an element in \(X_{\beta}\), we do not get anything new.
We can get something new only when we intersect elements from different sets \(\mathcal{S}_{\beta}\). Let \(\beta_{1}, \dots , \beta_{n}\) be a finite set of distinct indices from \(J\). Let \(U_{\beta_{i}}\) be an open set in \(X_{\beta_{i}}\). Consider the element \(B = \pi_{\beta_{1}} ^{-1} (U_{\beta_{1}}) \cap \dots \cap \pi_{\beta_{n}} ^{-1}(U_{\beta_{n}})\) of the generated basis. A point \((x_{\alpha})_{\alpha \in J}\) is in \(B\) iff its \(\beta_{i}\)-th coordinate is in \(U_{\beta_{i}}\) for all \(i\). There is no restriction on other coordinates (\(\pi_{\alpha}(B) = X_{\alpha}\) for other \(\alpha\)). That is,
\begin{equation} \label{eq:17} B = \prod_{\alpha \in J} U_{\alpha}, U_{\alpha} = X_{\alpha} \forall \alpha \neq \beta_{1}, \dots \beta_{n}. \end{equation}Let \(\{ X_{\alpha} \}_{\alpha \in J}\) be an indexed family of topological spaces, \(\mathcal{T}_{\text{box}}\) be the box topology and \(\mathcal{T}_{\text{product}}\) be the procuct topology. Then \(\mathcal{T}_{\text{box}} \succeq \mathcal{T}_{\text{product}}\), and the equality is reached iff \(J\) is finite.
Suppose the topology on each space \(X_{\alpha}\) is given by a basis \(\mathcal{B}_{\alpha}\). Then
\begin{equation} \label{eq:18} \left\{ \prod_{\alpha \in J} B_{\alpha} | B_{\alpha} \in \mathcal{B}_{\alpha} \forall \alpha \right\} \end{equation}is a basis for the box topology on \(\prod_{\alpha \in J} X_{\alpha}\), and
\begin{equation} \label{eq:21} \left\{ \left. \prod_{\alpha \in J} B_{\alpha} \right| B_{\alpha} \in \mathcal{B}_{\alpha} \text{ for finitely many } \alpha \right\} \end{equation}is a basis for the product topology on \(\prod_{\alpha \in J} X_{\alpha}\).
Let \(A_{\alpha}\) be a subspace of the topological space \(X_{\alpha}\) for each \(\alpha \in J\). Then \(\prod_{\alpha \in J} A_{\alpha}\) is a subspace of \(\prod_{\alpha \in J} X_{\alpha}\) if both are equipped with the box topology, or if both are equipped with the product topology.
If each space \(X_{\alpha}\) is Hausdorff space, then \(\prod_{\alpha \in J} X_{\alpha}\) is a Hausdorff space in both the box and product topology.
Let \(x, y\) be two distinct points in \(\prod_{\alpha \in J} X_{\alpha}\). Then they are different on at least one index, say \(\alpha_{0}\). Since \(X_{\alpha_{0}}\) is Hausdorff, there exist disjoint neighbourhoods \(U_{\alpha_{0}}\) and \(V_{\alpha_{0}}\) of \(x_{\alpha_{0}}\) and \(y_{\alpha_{0}}\), resp.
Define \(U = \prod_{\alpha \in J} U_{\alpha}\), where \(U_{\alpha} = X_{\alpha}\) for \(\alpha \neq \alpha_{0}\). Analogously define \(V\). Then \(U\) and \(V\) are disjoint neighbourhoods of \(x\) and \(y\), resp. \(U\) and \(V\) are open in both the box and product topology.
2.8.1. Product Space and Continuity
Let \(\{ X_{\alpha} \}_{\alpha \in J}\) be an indexed family of topological spaces. Then the projection mappings \(\pi_{\beta}\), \(\beta \in J\), are continuous, if the product has either the box topology or the product topology.
Let \(\beta \in J\). For every open set \(U_{\beta} \in X_{\beta}\), \(\pi_{\beta} ^{-1}(U_{\beta}) = \prod_{\alpha \in J} U_{\alpha}\), where \(U_{\alpha} = X_{\alpha}\) for \(\alpha \neq \beta\). This is an open set in either the box topology or the product topology.
Let \(J\) be an index set, \(f_{\alpha}: A \to \prod_{\alpha \in J} X_{\alpha}\) for each \(\alpha\), \(f: A \to \prod_{\alpha \in J} X_{\alpha}\) given by
\begin{equation} \label{eq:15} f(a) = (f_{\alpha}(a))_{\alpha \in J}. \end{equation}Let \(\prod_{\alpha \in J} X_{\alpha}\) have the product topology. Then the function \(f\) is continuous iff each function \(f_{\alpha}\) is continuous.
Let \(\pi_{\beta}\) be the projection map (Defn. 1) of the product onto its \(\beta\)-th coordinate.
Assume each function \(f\) is continuous. Then \(f_{\beta} = \pi_{\beta} \circ f\) is continuous, as it is the composition of two continuous functions.
Conversely, assume each coordinate function \(f_{\alpha}\) is continuous. By Thm. 1, it suffices to show that the pre-image of each subbasis element is open in \(A\). We take the typical subbasis in the product topology: \(\{ \pi_{\beta} ^{-1}(U_{\beta}) | \beta \in J, U_{\beta} \text{ open in } X_{\beta} \}\). Then \(f ^{-1}(\pi_{\beta} ^{-1}(U_{\beta})) = f _{\beta} ^{-1}(U_{\beta})\) (only has requirements on the \(\beta\)-th coordinate). This set is open in \(A\) because \(f_{\beta} ^{-1}\) is continuous.
Summary: We make use the subbasis elements, whose pre-images are only about one coordinate.
Assume the same. Let \(\prod_{\sigma \in J} X_{\alpha}\) have the box topology. Then every coordinate function \(f_{\alpha}\) is continuous if \(f\) continuous but the reverse does not generally hold.
The first direction can be proved in the same way.
Now, assume each \(f_{\alpha}\) is continuous and try to show that the pre-image of each basis element is open in \(A\) (Thm. 1)). Consider a typical basis element of the box topology, \(B = \prod_{\beta \in J} U_{\beta}\), where \(U_{\beta}\) is open in \(X_{\beta}\). Then,
\begin{equation} \label{eq:19} f ^{-1}(B) = \bigcap _{\beta \in J} f_{\beta} ^{-1}(U_{\beta}). \end{equation}This is where problems may arise: By the continuity of each \(f_{\beta}\), each \(f_{\beta} ^{-1}(U_{\beta})\) is open in \(A\). However, there is no guarantee that (countably) infinite intersection of open sets is still open.
To find a counterexample, we just need to construct a countable infinite intersection of open sets that is not open. E.g., in \(\mathbb{R} \), we have \(\cap _{i=1} ^{\infty} (-1 /n, -1 /n) = \{ 0 \}\), \(\cap _{i=1} ^{\infty} (0, 1 + 1 /n) = (0, 1]\), etc. And we can choose \(f_{\beta}\) to be simply the identity map. All together, here is a complete example: \(f: \mathbb{R} \to \mathbb{R} ^{\omega}\), where \(\mathbb{R} ^{\omega} = \prod_{n=1} ^{\infty} \mathbb{R} \). \(f(x) = (f_{n}(x))_{n \in \mathbb{Z} _{+}}\), where \(f_{n}(x) = x\). \(B = \prod_{n=1} ^{\infty} (-1 /n, 1 /n)\).
2.9. Metric Topology
Let \(X\) be a topological space. \(X\) is said to be metrizable if there exists a metric \(d\) on the set \(X\) that induces the topology of \(X\).
Let \(X\) be a topological space. Let \(A \subset X\). If there is a sequence of points \(\{ x_{n} \} \subset A\) s.t. \(x_{n} \to x\), then \(x \in \cl{A}\). The converse holds if \(X\) is metrizable.
Let \(f: X \to Y\). If the function \(f\) is continuous, then for every convergent sequence \(x_{n} \to x\) in \(X\), the sequence \(f(x_{n})\) converges to \(f(x)\). The converse holds if \(X\) is metrizable.
2.10. Compactness
The definitions are exactly the same:
A collection \(\mathcal{A}\) of subsets of a space \(X\) is said to be a covering of a set \(E\) if \(E \subset \cup _{A \in \mathcal{A}} A\). It is called an open covering of \(E\) if its element are open subsets.
A space \(K\) is compact if every open covering of \(K\) has a finite subcollection that also covers \(K\).
Let \(Y\) be a subspace of \(X\). Then \(Y\) is compact iff every open covering of \(Y\) by sets open in \(X\) contains a finite subcollection open covering of \(Y\).
Definition + intersection with \(Y\).
Every closed subspace of a compact space is compact.
For every open covering of the closed subspace \(Y\), we can add \(X - Y\) (open) to form a open covering of the compact space \(X\).
This is comparable to Thm. 1:
Every compact subspace of a Hausdorff space is closed.
Let \(Y\) be a compact subpace of the Hausdorff space \(X\). We shall prove that \(X - Y\) is open and appeal to Cor. 1.
Let \(x_0\) be any point of \(X - Y\). We want oshow that there is a neighbourhood of \(x_0\) completely in \(X - Y\). Since \(X\) is Hausdorff, for every point \(y \in Y\), there exist neighbourhood \(U_{y}, V_{y}\) of \(x, y\), resp., such that they are disjoint. \(\cup _{y \in Y} V_{y}\) is an open covering of \(Y\) by sets open in \(X\), so, by the compactness of \(Y\) and Thm. 1, there exist a finite subcollection \(\{ V_{y_1}, \dots , V_{y_{n}} \}\) that is an open covering of \(Y\) by sets open in \(X\).
Then, the intersection of the corresponding U’s, \(U_{y_1} \cap \dots \cap U_{y_{n}}\), is disjoint from \(Y\).
We extract a useful statement in the proof:
If \(Y\) is a compact subspace of the Hausdorff space \(X\) and \(x_0 \notin Y\), then there exist disjoint open sets \(U\) and \(V\) of \(X\) containing \(x_0\) and \(Y\), resp.
The image of a compact space under a continuous map is compact.
Use the definition of the continuous map.
Let \(f: X \to Y\) be a bijective continuous map. If \(X\) is compact and \(Y\) is Hausdorff, then \(f\) is a homeomorphism.
That is to prove that the inverse \(f ^{-1}: Y \to X\) is continuous. For every closed subset \(A \subset X\), we want to show that \((f ^{-1}) ^{-1}(A) = f(A)\) is closed in \(Y\). Since \(A\) is closed in a compact set, \(A\) is compact. By Thm. 1, \(f (A)\) is compact. Since \(Y\) is Hausdorff, by Thm. 1, \(f(A)\) is closed.
An arbitrary product space of compact topological spaces is compact.
A collection \(\mathcal{C}\) of subsets of \(X\) is said to have the finite intersection property if for every finite subcollection \(\{ C_1, \dots , C_{n} \}\) of \(\mathcal{C}\), the intersection \(C_1 \cap \dots \cap C_{n}\) is nonempty.
Let \(X\) be a topological space. The \(X\) is compact iff for every collection \(\mathcal{C}\) of closed sets in \(X\) with the finite intersection property, the intersection \(\cap _{C \in \mathcal{C}} C\) is nonempty.
2.11. TODO Quotient Topology
Let \(X\) and \(Y\) be topological spaces. Let \(p : X \to Y\) be a surjective map. Then \(p\) is said to be a quotient map if a subset \(U\) of \(Y\) is open iff \(p ^{-1}(U)\) is open in \(X\).
Note that the condition is stronger than continuity.
Let \(X\) be a topological space and \(A\) be a set. If \(p: X \to A\) is a surjective map, then there exists one topology \(\mathcal{T} = \{ U \subset A : p ^{-1}(U) ~\text{is open in}~X \}\) on \(A\) relative to which \(p\) is a quotient map. \(\mathcal{T}\) is called the quotient topology induced by \(p\).
Let \(X\) be a topologixal space, and let \(\sim\) be an equivalence relation on \(X\). \(\sim\) determines a partition of \(X\) (vice versa), i.e., \(\{ \overline{a} : a \in X \}\), where each subset is an equivalence class. Define a surjective map \(p: X \to X ^{*}\) by \(p(a) = \overline{a}\). Then the space \(\{ \overline{a} : a \in X \}\) paired with the quotient topology induced by \(p\) is called a quotient space of \(X\).