Countable and Uncountable sets

4 minute read

Definition 1.7.2

For each positive integer $n\in \mathbb{N}$, let \(\mathbb{N}_n:=\{1,2,\ldots,n \}\). If $A$ is a set, we say
(a) $A$ is finite if $A\sim \mathbb{N}_n$ for some $n$, or if $A=\emptyset$, where $\sim$ denotes equivalence relation if there exists a one-to-one function of $A$ onto $\mathbb{N}_n$.
(b) If $A$ is infinite if $A$ is not finite.
(c) $A$ is countable if $A\sim \mathbb{N}$.
(d) $A$ is uncountable if $A$ is neither finite nor countable.
(e) $A$ is at most countable if $A$ is finite or countable.

Theorem 1.7.4

$\mathbb{N}\times\mathbb{N}$ is countable

<proof> Construct a one-to-one mapping of $\mathbb{N}\times \mathbb{N}$ onto $\mathbb{N}$. Such a function is given by \(\begin{align} f(m,n):=2^{m-1}(2n-1), \text{where } n,m \in \mathbb{N} \end{align}\) First we want to show that the function $f$ is injective. Suppose that $f(m_1,n_1) = f(m_2, n_2)$ for $(m_1,n_1), (m_2,n_2)\in \mathbb{N}\times\mathbb{N}$. Since the prime factorization is unique, $2^{m_1-1}=2^{m_2-1}$, in other words, $m_1=m_2$. Therefore, $2n_1-1=2n_2-1$, i.e. $n_1=n_2$.

$\therefore f$ is injective.

Now we want to show that $f$ is onto. Let $a\in\mathbb{N}$ be given. Since $(2n-1)$ is odd number, the powers of 2 in the prime decomposition of $a$ are exactly $m-1$ and from there $n$ is determined.

$\therefore f$ is onto

\[\tag*{$\square$}\]

Theorem 1.7.6

Every infinite subset of a countable set is countable.

<proof> Let $A$ be a countable set and let \(\{x_n:n=1,2,\ldots\}\) be an enumeration of $A$. Suppose $E$ is an infinite subset of $A$. Then each $x\in E$ is of the form of $x_k$ for some $k\in \mathbb{N}$.

We inductively construct a function $f:\mathbb{N}\to E$ as follows: Let $n_1$ be the smallest integer such that $x_{n_1}\in E$. Such an integer exists by the well ordering principle. Having chosen $n_1,\ldots, n_{k-1}$, let $n_k$ be the smallest integer greater than $n_{k-1}$ such that \(\begin{align} x_{n_k}\in E\setminus \{x_{n_1},\ldots, x_{n_{k-1}} \} \end{align}\) Set $f(k) = x_{n_k}$. Since $E$ is infinite, $f$ is defined on $\mathbb{N}$.
If $m >k$, then $n_m > n_k$ and thus $x_{n_m} \neq x_{n_k}$. Therefore $f$ is one-to-one. The function $f$ is onto $E$ since if $x\in E$, then $x=x_j$ for some $j$. By construction, $n_k = j$ for some $k$ and thus $f(k) = x_{n_k}=x_j.$

\[\tag*{$\square$}\]

Theorem 1.7.7

If $f$ maps $\mathbb{N}$ onto $A$, then $A$ is at most countable.

<proof> If $A$ is finite, the result is certainly true. Suppose that $A$ is infinite. Since $f$ maps $\mathbb{N}$ onto $A$, each $a\in A$ is of the form $f(n)$ for some $n\in\mathbb{N}$. For each $a\in A$, by the well ordering principle \(\begin{align} f^{-1}(a):= \{ n\in\mathbb{N}:f(n)=a\} \end{align}\) has a smallest integer, which we denote by $n_a$. Consider the mapping $g:a\mapsto n_a$. Since $f$ is a function, $a\neq b \Rightarrow n_a \neq n_b$. It implies that g is one-to-one. By the construction of the function $g$, it is onto. Also, since $A$ is infinite, \(S=\{n_a:a\in A\}\) is an infinite subset of $\mathbb{N}$ by Theorem 1.7.6. Since $g: S \rightarrow A$ is bijective and the set $S$ is countable, $A$ is countable.

$\therefore A$ is at most countable. \(\tag*{$\square$}\)

Theorem 1.7.15

If \(\{E_n\}_{n=1}^\infty\) is a sequence of countable sets and \(\begin{align} S = \bigcup_{n=1}^\infty E_n \end{align}\) then $S$ is countable.

<proof>

Since $E_n$ is countable for each $n\in \mathbb{N}$, we can write \(\begin{align} E_n =\{x_{n,k}: k=1,2,\ldots\} \end{align}\)

Now, we write the elements of $S$ in the form of a table:

\[\begin{align*} \begin{pmatrix} x_{1,1}& x_{1,2}& x_{1,3}&\cdots\\ x_{2,1}& x_{2,2}& x_{2,3}&\cdots\\ x_{3,1}& x_{3,2}& x_{3,3}& \cdots\\ \vdots&\vdots&\vdots&\ddots \end{pmatrix} \end{align*}\]

Then we can enumerate the elements as follows:

\[\begin{align*} x_{1,1}\to x_{1,2}\to x_{2,1}\to x_{3,1}\to x_{2,2}\to x_{1,3}\to \cdots \end{align*}\]

Since there could be duplicates, there is a one-to-one correspondence between $S$ and a subset of $\mathbb{N}$, denoted as $J$. In other words, $\lvert S \rvert =\lvert J\rvert$.

Since $E_n$ is an infinite subset of $S$, the set $S$ is infinite. Thus $J$ is infinite subset of countable set $\mathbb{N}$. By Theorem 1.7.6, $J$ is countable, so $S$ is also countable. \(\tag*{$\square$}\)

Corollary 1.7.16

$\mathbb{Q}$ is countable.

<proof>

For each $m\in\mathbb{N}$, let

\[\begin{align} E_m = \left\{\frac{n}{m}:n\in\mathbb{Z}\right\} \end{align}\]

The $E_m$ is countable, and since $\mathbb{Q} = \bigcup_{m=1}^\infty E_m$, by Theorem 1.7.15 the set $\mathbb{Q}$ is countable.

\[\tag*{$\square$}\]

Theorem

Let $A$ be a countable set. Then $B_n = \underbrace{A\times A\times \cdots \times A}_{n \text{ times}}$ is countable.

<proof> We prove it with mathematical induction on $n$. For $n=1$, $B_1$ is clearly countable. Suppose that $B_{k}$ is countable for $k\geq 1$. For any $a\in A$,

\[\begin{align*} \{a\}\cup B_k \end{align*}\]

is countable. Then, by Theorem 1.7.6,

\[\begin{align*} B_{k+1}=\bigcup_{a\in A} \left( \{a \}\cup B_k \right) \end{align*}\]

is countable set since $A$ is countable. \(\tag*{$\square$}\)

Theorem 1.7.17

The closed interval $[0,1]$ is uncountable.

<proof> We want to show that every countable subset of $[0,1]$ is a proper subset of $[0,1]$. Thus, $[0,1]$ cannot be countable.

Let \(E=\{x_n, n=1,2,\ldots \}\) be a countable subset of $[0,1]$. Then for each $x_n$ has a decimal expansion: \(\begin{align} x_n = 0.x_{n,1}x_{n,2}x_{n,3}\cdots \end{align}\) where for each \(k\in \mathbb{N}, x_{n,k}\in \{0,1,\ldots, 9\}\). We now define a new number \(y:= 0.y_1y_2y_3\cdots\) as follows: If $x_{n,n} \leq 5$, define $y_n :=6$; if $x_{n,n} \geq 6$, define $y_n=3$. Then $y\in [0,1]$, and since $y_n \neq 0$ or 9, $y_n \neq x_{n,n}$, we have $y\neq x_n$ for any $n\in\mathbb{N}$.

$\therefore y\not\in E$
$\therefore E$ is a proper subset of $[0,1]$. \(\tag*{$\square$}\)

Reference

  • Manfred Stoll, Introduction to Real Analysis, Pearson