Metric spaces

This page is under construction.

1. Definition and examples

Definition 1.1 Metric space

Let $X$ be a set (whose elements we shall call ‘points’). A metric on $X$ is a function $\rho: X \times X \to \mathbb{R}$ such that for all $x, y, z \in X$,

  1. $\rho(x, y) \geq 0$ (non-negativity)
  2. $\rho(x, y) = 0$ if and only if $x = y$
  3. $\rho(x, y) = \rho(y, x)$ (symmetry)
  4. $\rho(x, z) \leq \rho(x, y) + \rho(y, z)$ (triangle inequality)

The pair $(X, \rho)$ is called a metric space. With some abuse of notation, we may refer to $X$ as a metric space with the implied metric $\rho$.

Examples.

Metrics in $\mathbb{R}$. Consider the set $\mathbb{R}$ of real numbers. The absolute difference function $|\cdot - \cdot|: \mathbb{R} \times \mathbb{R} \to \mathbb{R}$ is a metric on $\mathbb{R}$. Indeed for any real numbers $x, y, z$, we have $|x - y| \geq 0$, $|x - y| = 0$ if and only if $x = y$, $|x - y| = |y - x|$, and $|x - z| \leq |x - y| + |y - z|$.

Let $n \geq 1$. The set $\mathbb{R}^n$ is the set of all $n$-tuples of real numbers. For the sake of brevity, we shall denote an element of $\mathbb{R}^n$ by $x = (\xi_1, \xi_2, \ldots, \xi_n)$, where $\xi_i \in \mathbb{R}$ for $1 \leq i \leq n$. We define the Euclidean metric on $\mathbb{R}^n$ to be the map ${\ell^2}: \mathbb{R}^n \times \mathbb{R}^n \to \mathbb{R}$ given by $$ {\ell^2}(x, y) = \sqrt{\sum_{i = 1}^n (\xi_i - \eta_i)^2}. $$

In $\mathbb{R}^3$ and $\mathbb{R}^2$ one can visualize ${\ell^2}(x, y)$ as the distance between the points $x$ and $y$ in the Euclidean plane and space, respectively. The Euclidean metric is the most commonly used metric on $\mathbb{R}^n$ as it is the natural generalization of the absolute difference function on $\mathbb{R}$ and provides a natural notion of distance between points in $\mathbb{R}^n$.

The Manhattan metric (also known as the taxicab metric) on $\mathbb{R}^n$ is the map ${\ell^1}: \mathbb{R}^n \times \mathbb{R}^n \to \mathbb{R}$ given by $$ {\ell^1}(x, y) = \sum_{i = 1}^n |\xi_i - \eta_i| $$ for $x = (\xi_1, \xi_2, \ldots, \xi_n)$ and $y = (\eta_1, \eta_2, \ldots, \eta_n)$. The Manhattan metric is so-called because it is the distance a taxicab would have to travel in a rectangular grid of streets (similar to those found in Manhattan) to get from the point $x$ to the point $y$.

Finally, we define the Chebyshev distance on $\mathbb{R}^n$ as $$ {\ell^\infty}(x, y) = \max_{1 \leq i \leq n} |\xi_i - \eta_i|. $$

These three metrics are related by the inequalities $$ {\ell^\infty}(x, y) \leq {\ell^2}(x, y) \leq {\ell^1}(x, y) \leq n {\ell^\infty}(x, y). $$

Discrete metric. Let $X$ be an arbitrary set. The discrete metric on $X$ is the map $\delta: X \times X \to \mathbb{R}$ given by $$ \delta(x, y) = \begin{cases} 0 & \text{if } x = y \\ 1 & \text{if } x \neq y. \end{cases} $$

2. The topology of a metric space

Definition 2.1

Let $(X, \rho)$ be a metric space. A ball centered at a point $x \in X$ of radius $\varepsilon > 0$ is the set $$ B_{\varepsilon}(x) = \{y \in X: \rho(x, y) \lt \varepsilon\}. $$

Let $A$ be a subset of $X$. Then $x$ is an interior point of $A$ if there exists $\varepsilon > 0$ such that $B_{\varepsilon}(x) \subseteq A$. It is an exterior point of $A$ if there exists $\varepsilon > 0$ such that $B_{\varepsilon}(x) \cap A = \varnothing$. It is a boundary point of $A$ if it is neither an interior point nor an exterior point of $A$.

Collectively, the set of interior, exterior, and boundary points of the set $A \subset X$ (with $X$ a metric space) is called the interior, exterior, and boundary of $A$, respectively and is denoted by $\operatorname{int}(A)$, $\operatorname{ext}(A)$, and $\partial A$, in that order.

If $x$ is an interior point of $A$ then it must be an element of $A$ since $B_{\varepsilon}(x)$ contains at least one point of $A$ (namely, $x$ itself). On the other hand, if $x$ is an exterior point of $A$ then it cannot be an element of $A$ since $B_{\varepsilon}(x)$ contains no points of $A$ for any $\varepsilon > 0$. Boundary points can be either elements or non-elements of $A$.

Example. In $\mathbb{R}^2$ with the $\ell^2$ metric, the open ball or radius $1$ centered at the origin is the open disc $$ B_1((0, 0)) = \{(x, y) \in \mathbb{R}^2: x^2 + y^2 \lt 1\}. $$ On the other hand, using the $\ell^1$ metric, the earlier-defined open ball is the set $$ B_1((0, 0)) = \{(x, y) \in \mathbb{R}^2: |x| + |y| \lt 1\}, $$ which geometrically is a square (diamond) with vertices $(\pm 1, 0)$ and $(0, \pm 1)$. Finally, using the standard discrete metric, the ball is reduced to the singleton $\{(0, 0)\}$.

Example. Let $X$ be an arbitrary set and consider the standard discrete metric $\delta$ in $X$. If $A$ is a subset pf $X$, then $\operatorname{int}(A) = A$ and $\operatorname{ext}(A) = X \setminus A$. $A$ has no boundary points.

A subset $U$ of $X$ is open if and only if for every $x \in U$ there exists $\varepsilon > 0$ such that $B_{\varepsilon}(x) \subset U$. A subset $F$ of $X$ is closed if and only if $X \setminus F$ is open.

Theorem 2.2

Let $(X, \rho)$ be a metric space.

  1. The empty set $\varnothing$ and the whole space $X$ are open.
  2. Let $\mathfrak{U}$ be a collection of open sets. Then $\bigcup_{U \in \mathfrak{U}} U$ is open.
  3. Let $\mathfrak{U}$ be a finite collection of open sets. Then $\bigcap_{U \in \mathfrak{U}} U$ is open.

Proof.

(2) Let $x \in \bigcup_{U \in \mathfrak{U}} U$. Then $x \in U$ for some $U' \in \mathfrak{U}$. Since $U'$ is open, there exists $\varepsilon > 0$ such that $B_{\varepsilon}(x) \subset U'$. But $U' \subset \bigcup_{U \in \mathfrak{U}} U$, so $B_{\varepsilon}(x) \subset \bigcup_{U \in \mathfrak{U}} U$.

3. Sequences and series in metric spaces

Definition 3.1 Sequence

A sequence in a metric space $(X, \rho)$ is a function $a : \mathbb{N} \to X$.

Instead of the usual notation $a(n)$ for the value of the function $a$ at $n$, we write $a_n$, and call it the $n$th term of the sequence. Similarly,we write $(a_n)$ instead of $a : \mathbb{N} \to X$.

The following are common ways to describe a sequence:

Definition 3.2 Convergence of a sequence

A sequence $(a_n)$ in a metric space $(X, \rho)$ is said to converge to a point $a \in X$ if for every $\varepsilon \gt 0$, there exists $N \in \mathbb{N}$ such that $\rho(a_n, a) \lt \varepsilon$ for all $n \geq N$.

If $(a_n)$ converges to $a$, we write $(a_n) \to a$. (Another common notation is $\lim_{n \to \infty} a_n = a$.) If $(a_n)$ does not converge, we say that $(a_n)$ diverges.

(Henceforth we shall assume that all sequences are in some arbitrary metric space $(X, \rho)$ unless otherwise stated.)

Alternatively, we have the following topological definition (framing convergence in terms of open balls): a sequence $(a_n)$ in a metric space $(X, \rho)$ converges to a point $a \in X$ if for every open ball $B_{\varepsilon}(a)$, there exists $N \in \mathbb{N}$ such that $a_n \in B_{\varepsilon}(a)$ for all $n \geq N$.

If $(a_n) \to a$, we say that $a$ is the limit of the sequence $(a_n)$. Note the use of the definite article "the" in the previous sentence. This is because the limit of a sequence is unique, as our next theorem shows.

Theorem 3.3

The limit of a sequence, if it exists, is unique.

Proof. Suppose $(a_n) \to a$ and $(a_n) \to b$. We want to show that $a = b$.

Let $\varepsilon := \frac{1}{3}\rho(a, b) > 0$. Now since $(a_n) \to a$, there exists $N_1 \in \mathbb{N}$ such that $\rho(a_n, a) \lt \varepsilon$ for all $n \geq N_1$. Similarly, since $(a_n) \to b$, there exists $N_2 \in \mathbb{N}$ such that $\rho(a_n, b) \lt \varepsilon$ for all $n \geq N_2$.

Now choose $m := \max\{N_1, N_2\}$. Then for all $n \geq m$, we have $\rho(a_n, a) \lt \varepsilon$ and $\rho(a_n, b) \lt \varepsilon$. Using the triangle inequality, we have $$ \begin{align} \rho(a, b) & \leq \rho(a_n, a) + \rho(a_n, b) = \rho(a, a_n) + \rho(a_n, b) \\ & \lt \varepsilon + \varepsilon = \frac{2}{3}\rho(a, b) \end{align} $$ so that $\rho(a,b) \lt \frac{2}{3}\rho(a, b)$. This is a contradiction since $\rho(a, b) \gt 0$. Therefore $a = b$.

Example. We will show that the sequence $\displaystyle\left(\frac{n+1}{n}\right)$ in $\mathbb{R}$ with the usual metric converges to $1$.

Let $(a_n)$ be the sequence described above. Then if $(a_n) \to 1$, we need to show that for any $\varepsilon \gt 0$ and for sufficiently large $N$ we have $$ \left\lvert \frac{n+1}{n} - 1 \right\rvert = \left\lvert \frac{n+1-n}{n} \right\rvert = \left\lvert \frac{1}{n} \right\rvert \lt \varepsilon $$ and thus $1/n \lt \varepsilon$ and $n \gt 1/\varepsilon$. Therefore choosing $N \in \mathbb{N}$ such that $N \gt 1/\varepsilon$ will suffice.

We can proceed with the formal proof.

Let $\varepsilon \gt 0$. Choose $N \in \mathbb{N}$ such that $N \gt 1/\varepsilon$. Then for all $n \geq N$, we have $n \gt 1/\varepsilon$ and thus $1/n \lt \varepsilon$. Therefore $$ \left\lvert \frac{1}{n} \right\rvert \lt \varepsilon $$ and so $(a_n) \to 1$.

Example. We will show that the sequence $(a_n) = n$ does not converge in $\mathbb{R}$.

Tending towards contradiction, suppose that $(a_n)$ converges to some value $\nu$ and consider $\varepsilon = 1 > 0$. Then because $(a_n)$ converges, there must exist some natural number $N$ such that for any $n \in \mathbb{N}$ with $n \geq N$, we have $\rho(a_n, \nu) \lt \varepsilon$. In particular, we have $\rho(n, \nu) \lt 1$ which implies that $n \lt \nu + 1$, so that $\mathbb{N}$ is bounded above by $\nu + 1$. This is a contradiction and so $(a_n)$ does not converge.

Theorem 3.4 Boundedness of convergent sequences

A sequence $(a_n)$ is bounded if and only if there exists a point $r \in X$ and a real number $M \gt 0$ such that $\rho(a_n, r) \leq M$ for all $n \in \mathbb{N}$. Every convergent sequence in a metric space is bounded.

Proof. To prove that $(a_n)$ is bounded, we need to find a point $r \in X$ and a real number $M \gt 0$ such that $\rho(a_n, r) \leq M$ for all $n \in \mathbb{N}$. Now because $(a_n)$ is convergent (to some point $a$, say), there exists $N \in \mathbb{N}$ such that $\rho(a_n, a) \lt \varepsilon$ for all $n \geq N$. Fix $\varepsilon = 1$. Then in particular, we have $\rho(a_n, a) \lt 1$.

Define $$ M := \max\{\rho(a_1, a), \rho(a_2, a), \dots, \rho(a_{N-1}, a), 1\}. $$ Thus for all $n \geq N$, $\rho(a_n, a) \lt 1 \leq B$ (because $(a_n)$ is convergent) and for all $n \lt N$, $1 \leq M$ (by our definition of $M$). Therefore $\rho(a_n, a) \leq M$ for all $n \in \mathbb{N}$ and so $(a_n)$ is bounded.

Theorem 3.5 Subsequences

A subsequence of a sequence $(a_n)$ is a sequence $(a_{n_k})$ where $n_1 \lt n_2 \lt \cdots$ is a strictly increasing sequence of natural numbers. If $(a_n)$ converges to $a$, then every subsequence of $(a_n)$ converges to $a$.

Proof. Because $(a_n)$ converges to $a$, for every $\varepsilon \gt 0$ there exists $N \in \mathbb{N}$ such that $\rho(a_n, a) \lt \varepsilon$ for all $n \geq N$. Now let $(a_{n_k})$ be a subsequence of $(a_n)$. Since the $n_k$’s are strictly increasing $n_k \geq k$[1] and thus for all $k \geq N$, we have $n_k \geq N$, so that $\rho(a_{n_k}, a) \lt \varepsilon$. Therefore $(a_{n_k})$ converges to $a$.

Theorem 3.6 Algebraic properties of convergent sequences

Let $(a_n)$ and $(b_n)$ be sequences in a metric space $(X, \rho)$ and let $a, b \in X$. If $(a_n) \to a$ and $(b_n) \to b$, then:

  1. $(a_n + b_n) \to a + b$
  2. $(ca_n) \to ca$ for any $c \in V$, where $V$ is a normed vector space
  3. $(a_n b_n) \to ab$
  4. $(a_n / b_n) \to a/b$ if $b \neq 0$

Proof.

(1) Let $\varepsilon > 0$. Then $\varepsilon / 2 > 0$, and since $(a_n) \to a$ and $(b_n) \to b$, there exists $N_1, N_2 \in \mathbb{N}$ such that $\rho(a_n, a) \lt \varepsilon / 2$ for all $n \geq N_1$ and $\rho(b_n, b) \lt \varepsilon / 2$ for all $n \geq N_2$. Let $N := \max\{N_1, N_2\}$. Then for all $n \geq N$, we have $$ \begin{align} \rho((a_n + b_n), (a + b)) & \leq \rho(a_n, a) + \rho(b_n, b) \\ & \lt \frac{\varepsilon}{2} + \frac{\varepsilon}{2} = \varepsilon. \end{align} $$ Therefore $(a_n + b_n) \to a + b$.

(2) To prove this, we want to show that for any $\varepsilon \gt 0$, there exists $N \in \mathbb{N}$ such that $\rho(ca_n, ca) \lt \varepsilon$ for all $n \geq N$. From $\rho(ca_n, ca) \lt \varepsilon$, we have $\|c\| \cdot \rho(a_n, a) \lt \varepsilon$ (where $\|c\|$ is the norm of $c$).

We now proceed with the actual proof. Let $\varepsilon \gt 0$. Then $\displaystyle\frac{\varepsilon}{\|c\|} \gt 0$. Since $(a_n) \to a$, there exists $N \in \mathbb{N}$ such that $\rho(a_n, a) \lt \displaystyle\frac{\varepsilon}{\|c\|}$ for all $n \geq N$. Therefore for all $n \geq N$, we have $$ \begin{align} \rho(ca_n, ca) & = \|c\| \cdot \rho(a_n, a) \\ & \lt \|c\| \cdot \frac{\varepsilon}{\|c\|} = \varepsilon. \end{align} $$ Therefore $(ca_n) \to ca$.

(3) First observe that $$ \begin{align} \rho(a_n b_n, ab) & = \rho(a_n b_n, a b_n + a b - a b_n) \\ & \leq \rho(a_n b_n, a b_n) + \rho(a b_n, a b) + \rho(a b, a b_n) \\ & =

Theorem 3.7 Order properties of convergent sequences

Let $(a_n)$ and $(b_n)$ be sequences in a metric space $(X, \rho)$ and let $a, b \in X$. If $(a_n) \to a$ and $(b_n) \to b$, then:

  1. If $a_n \geq 0$ for all $n \in \mathbb{N}$, then $a \geq 0$;
  2. If there exists $N \in \mathbb{N}$ such that $a_n \leq b_n$ for all $n \geq N$, then $a \leq b$;
  3. If $\xi \leq a_n \leq \eta$ for all $n \in \mathbb{N}$, then $\xi \leq a \leq \eta$.

Proof.

(1) We prove the contrapositive. Suppose $a \lt 0$. Let $\varepsilon \gt 0$ be such that $a + \varepsilon \lt 0$. Then since $(a_n) \to a$, there exists $N \in \mathbb{N}$ such that $\rho(a_n, a) \lt \varepsilon$ for all $n \geq N$. Taking $n = N$, we have $\rho(a_N, a) \lt \varepsilon$ and thus $a_N \lt a + \varepsilon \lt 0$. Therefore $a_N \lt 0$ and so $a_n \not\geq 0$ for all $n \geq N$. Therefore $(a_n)$ does not converge to $a$.

(2) Consider $N = 1$, i.e., $a_n \leq b_n$ for all $n \in \mathbb{N}$. Let $(c_n)$ be the sequence defined by $c_n = b_n - a_n$. Then $c_n \geq 0$ for all $n \in \mathbb{N}$ and $c_n \to b - a$. By (1), $b - a \geq 0$ and thus $a \leq b$. If $N \gt 1$, then we apply the same argument to the $N$-tail of $(a_n)$ and $(b_n)$.

(3) We have $0 \leq a_n - \xi \leq \eta - \xi$ for all $n \in \mathbb{N}$ from the hypothesis, whence, defining the constant sequence $(x_n) := \eta - \xi$ and the sequence $(y_n) := a_n - \xi$, we have $(x_n) \to \eta - \xi$ and $(y_n) \to a - \xi$. By (1) and (2), we have $0 \leq a - \xi \leq \eta - \xi$ and thus $\xi \leq a \leq \eta$.

Definition 3.8 Cauchy sequence

A sequence $(a_n)$ in a metric space $(X, \rho)$ is Cauchy if for every $\varepsilon \gt 0$, there exists $N \in \mathbb{N}$ such that $\rho(a_n, a_m) \lt \varepsilon$ for all $n, m \geq N$.

Cauchy sequences provide us with an alternative way to work with convergence, even if the limit of the sequence is not necessarily known.

Theorem 3.9

Let $(a_n)$ be a sequence in a metric space $(X, \rho)$. Then the following statements are true.

  1. If $(a_n)$ converges, then $(a_n)$ is Cauchy.
  2. If $(a_n)$ is Cauchy, then $(a_n)$ is bounded.
  3. If $(a_n)$ is Cauchy, then $(a_n)$ converges if and only if it has a convergent subsequence.

Proof.

(1) Let $(a_n) \to a$. Then for every $\varepsilon \gt 0$, there exists $N \in \mathbb{N}$ such that $\rho(a_n, a) \lt \varepsilon / 2$ for all $n \geq N$. Therefore for all $n, m \geq N$, we have $$ \rho(a_n, a_m) \leq \rho(a_n, a) + \rho(a, a_m) \lt \frac{\varepsilon}{2} + \frac{\varepsilon}{2} = \varepsilon. $$ Therefore $(a_n)$ is Cauchy.

(2) Let $(a_n)$ be Cauchy. Then for every $\varepsilon \gt 0$, there exists $N \in \mathbb{N}$ such that $\rho(a_n, a_m) \lt \varepsilon$ for all $n, m \geq N$. In particular, for $\varepsilon = 1$, there exists $N \in \mathbb{N}$ such that $\rho(a_n, a_m) \lt 1$ for all $n, m \geq N$. Moreover, since $N \geq N$, we have $\rho(a_n, a_N) \lt 1$ for all $n \geq N$.

Define $$ M := \max\{\rho(a_1, a_N), \rho(a_2, a_N), \dots, \rho(a_{N-1}, a_N), 1\}. $$ Then for all $n \geq N$, we have $\rho(a_n, a_N) \lt 1 \leq M$ (by our choice of $N$ above) and for all $n \lt N$, $\rho(a_n, a_N) \leq M$ (by our definition of $M$). Therefore $\rho(a_n, a_N) \leq M$ for all $n \in \mathbb{N}$ and so $(a_n)$ is bounded.

(3) The necessary condition is trivial. If we suppose that $(a_n) \to a$, then by Theorem 3.5, every subsequence of $(a_n)$ converges to $a$. We prove sufficiency.

Suppose $(a_{n_k})$ is a subsequence of $(a_n)$ that converges to $a$. Let $\varepsilon > 0$. Then since $(a_n)$ is Cauchy, there exists $N \in \mathbb{N}$ such that for all $n, m \geq N$, we have $$\rho(a_n, a_m) \lt \frac{\varepsilon}{2}.$$ Moreover, since $(a_{n_k}) \to a$, there exists $M \in \mathbb{N}$ such that for all $k \geq M$, $$\rho(a_{n_k}, a) \lt \frac{\varepsilon}{2}.$$

If $K > \max\{N, M\}$, then for all $n \geq K$, and since $n_K \geq K$, we have $$ \begin{align} \rho(a_n, a) &\leq \rho(a_n, a_{n_K}) + \rho(a_{n_K}, a) \\ &\lt \frac{\varepsilon}{2} + \frac{\varepsilon}{2} = \varepsilon. \end{align} $$

Therefore $(a_n) \to a$.

The converse of Theorem 3.9(1) (i.e., if $(a_n)$ is Cauchy, then $(a_n)$ converges) is not true in general. For example, consider the sequence $$ a_n = \left(1 + \frac{1}{n}\right)^n $$ in the metric space $\mathbb{Q}$ with the usual metric. Then $(a_n)$ is Cauchy but does not converge in $\mathbb{Q}$ ($(a_n) \to e$ but $e \not\in \mathbb{Q}$). A metric space in which every Cauchy sequence converges is called a complete metric space.

4. Compactness

We introduce the notion of compactness, which generalizes the notion of closed and bounded sets in $\mathbb{R}^n$. We proceed with some definitions.

We can rephrase the first definition above in terms of closed sets. If $\mathfrak{U}$ is an open cover of $X$ and if $\mathfrak{F}$ is the collection of closed sets in $X$ such that each $F \in \mathfrak{F}$ is the complement of some $U \in \mathfrak{U}$, then by De Morgan’s laws, $$ X = \bigcup_{U \in \mathfrak{U}} U = \bigcup_{F \in \mathfrak{F}} F \implies \bigcap_{F \in \mathfrak{F}} F = \varnothing. $$

We now proceed with the main definition.

Theorem 4.1 Equivalent definitions of compactness

Let $X$ be a metric space. Then the following statements are equivalent:

  1. Every open cover of $X$ has a finite subcover.
  2. Every collection of closed sets in $X$ with the finite intersection property has a nonempty intersection.
  3. If $F_1 \supseteq F_2 \supseteq \cdots$ is a nested sequence of nonempty closed sets in $X$, then $\bigcap_{n=1}^{\infty} F_n$ is nonempty.
  4. Every sequence in $X$ has a convergent subsequence.
  5. $X$ is totally bounded and complete.

A metric space that satisfies the equivalent conditions above is called compact. (Most textbooks, nevertheless, use only the first one as the definition of compactness.)

Proof.

$(1) \implies (2).$ Suppose $\mathfrak{F}$ is a collection of closed sets in $X$ with the finite intersection property. Suppose further that $\displaystyle\bigcap_{F \in \mathfrak{F}} F$ is nonempty. Then if we define $\mathfrak{U}$ as the collection of the complements $U := X \setminus F$ for each $F \in \mathfrak{F}$, we have $$ \bigcup_{U \in \mathfrak{U}} U = \bigcup_{F \in \mathfrak{F}} (X \setminus F) = X \setminus \bigcap_{F \in \mathfrak{F}} F = X \setminus \varnothing = X, $$ so that $\mathfrak{U}$ is an open cover of $X$. By hypothesis, there exists a finite subcover $\{U_1, U_2, \dots, U_n\}$ of $\mathfrak{U}$ and thus $$ X = \bigcup_{i=1}^n U_i = \bigcup_{i=1}^n (X \setminus F_i) = X \setminus \bigcap_{i=1}^n F_i, $$ so that $\displaystyle\bigcap_{i=1}^n F_i = \varnothing$. This contradicts the fact that $\mathfrak{F}$ has the finite intersection property. Therefore $\mathfrak{F}$ has a nonempty intersection.

$(2) \implies (3).$

$(3) \implies (4).$

$(4) \implies (5).$

$(5) \implies (1).$

5. Continuous functions on metric spaces

Definition 5.1

Let $(X, \rho)$ and $(Y, \sigma)$ be metric spaces. A function $f: X \to Y$ is continuous at a point $x \in X$ if for every $\varepsilon > 0$ there exists $\delta > 0$ such that $\sigma(f(x), f(y)) \lt \varepsilon$ whenever $\rho(x, y) \lt \delta$. We say that $f$ is continuous if it is continuous at every point of $X$.

Share this post: Email Twitter Reddit