Well-orderings and ordinals

Well-orderings and ordinals#

Star Issue Watch Follow

\(\newcommand{\vd}{\vdash}\) \(\newcommand{\dvd}{\models}\) \(\newcommand{\imp}{\Rightarrow}\)

Well-orderings#

Definition 122 (Strict total order)

Given a set \(X,\) a strict total order or linear order is a pair \((X, <),\) where \(X\) is a set and \(<\) is a relation on \(X\) that satisfies

  1. \(x \not < x\) for all \(x \in X.\)

  2. If \(x < y, y < z,\) then \(x < z\) for all \(x, y, z \in X.\)

  3. \(x < y\) or \(x = y\) or \(y < x\) for all \(x, y \in X.\)

We use the shorthand \(x > y\) meaning \(y < x.\)

Definition 123 (Non-strict total order)

A non-strict total order is a pair \((X, \leq),\) where \(X\) is a set and \(\leq\) is a relation on \(X\) that satisfies

  1. \(x \leq x.\)

  2. If \(x \leq y, y \leq z\) then \(x \leq z\) for all \(x, y, z \in X.\)

  3. If \(x \leq y, y \leq x\) then \(x = y\) for all \(x, y \in X.\)

  4. \(x \leq y\) or \(y \leq x\) for all \(x, y \in X.\)

We use the shorthand \(x \geq y\) meaning \(y \leq x.\)

Definition 124 (Well-ordering)

A strict total order \((X, <)\) is a well-ordering if every non-empty subset has a least element, that is

\[\begin{equation} (\forall S \subseteq X)(S \neq \emptyset \imp (\exists x \in S)(\forall y \in S) y \geq x) \end{equation}\]

Lemma 49 (Well-ordering equivalent condition)

A total order is a well-ordering if and only if it has no infinite strictly decreasing sequence.

Solution: Well-ordering equivalent condition

Suppose \((X, <)\) is a total order.

Implies: Suppose \((X, <)\) is a total order. Let \(x_1 > x_2 > \dots,\) be an infinite strictly decreasing sequence in \(X.\) Then \(\{x_n \in X: n \in \mathbb{N}\}\) has no least element, contradicting the definition of a total order.

Is implied by: Suppose \((X, <)\) does not have an infinite strictly decreasing sequence. Let \(S \subseteq X.\) If \(S\) is non-empty and has no least element, then for any \(x \in S\) we can find \(x' \in S\) such that \(x > x'.\) We can continue in this way to obtain an infinite strictly decreasing sequence

\[\begin{equation} x > x' > x'' > \dots, \end{equation}\]

which is a contradiction. Therefore if \(S\) is non-empty, it must have a least element.

Definition 125 (Order isomorphism)

We say the strict total orders \(X, Y\) are isomorphic if there exists a bijection \(f: X \to Y\) that is order preserving, that is \(x < y \imp f(x) < f(y).\)

Lemma 50 (Principle by induction)

Let \(X\) be a well-ordered set. Suppose \(S \subseteq X\) has the property

\[\begin{equation} (\forall x)\left(((\forall y) y < x \imp y \in S) \imp x \in S\right). \end{equation}\]

Then \(S = X.\)

Proof: Principle by induction

Let \(X\) be a well-ordering. Suppose \(S \neq X\) is a subset for which the property in the theorem statement holds. Since \(X\) is a well-ordering, if \(X \setminus S\) is non-empty, it has a least element, say \(x.\) Since \(y\) is the least element of \(X \setminus S,\) this means that for all \(y \in S\) such that \(y < x,\) we have \(y \in S.\) But since \(S\) has the stated property, it follows that \(x \in S,\) which is a contradiction. Therefore \(X \setminus S = \emptyset\) and \(S = X.\)

Lemma 51 (Uniqueness of well-ordering isomorphisms)

Let \(X\) and \(Y\) be isomorphic well-orderings. Then, there is a unique isomorphism between \(X\) and \(Y.\)

Proof: Uniqueness of well-ordering isomorphisms

Let \(X, Y\) be well-orderings and \(f\) and \(g\) be two isomorphisms from \(X\) to \(Y.\) Note that since \(f, g\) preserve order, they must map the least element of \(X\) to the least element of \(Y,\) so \(f(\min X) = g(\min Y).\) By the principle by induction it suffices to show that if \(f(y) = g(y)\) for all \(y < x,\) then \(f(x) = g(x).\)

Let \(S = \{f(x) \in Y: y < x\}.\) We know that \(Y \setminus S\) is non-empty because \(f(x) \not \in S.\) Therefore \(Y \setminus S\) must have a least element, say \(a = \min(Y \setminus S).\) Now it must be that \(a = f(x)\) because otherwise, by the minimality of \(a\) we would have \(a < f(x),\) which means that \(f^{-1}(a) < x,\) since \(f\) is order preserving. Then \(f^{-1}(a) \in S\) and \(a = f(f^{-1}(a)) \in Y,\) which is a contradiction.

The induction hypothesis is that \(f(y) = g(y)\) for all \(y < x.\) Therefore \(\{g(y): y < x\} = S\) as well. By the same argument as above \(g(x) = \min(Y \setminus S) = f(x).\) Which concludes the proof.

Definition 126 (Initial segment)

A subset \(Y\) of a totally ordered \(X\) is an initial segment if

\[\begin{equation} x \in Y, y < x \imp y \in Y. \end{equation}\]

Lemma 52 (Initial segment equivalent condition)

Every initial segment \(Y\) of a well-ordering \(X\) is of the form

\[\begin{equation} I_x = \{y \in X: y < x\}. \end{equation}\]
Proof: Initial segment equivalent condition

Let \(X\) is a well-ordering, and let \(Y\) to be an initial segment in \(X.\)

Take \(x = \min(X \setminus Y).\) Then for any \(y \in I_x,\) we have \(y < x.\) So \(y \in Y\) by definition of \(x.\) So \(I_x \subseteq Y.\)

Going the other direction, if \(y \in Y,\) then \(y \neq x.\) Also by the definition of \(x,\) we cannot have \(y > x,\) so \(y < x.\) Therefore \(y \in I_x\) and \(Y \subseteq I_x.\)

We conclude that \(Y = I_x.\)

In the following results, we think of each function \(f: X \to Y\) as a subset of \(X \times Y,\) i.e. an element of \(\mathbb{P}(X \times Y).\)

Definition 127 (Restriction of a function)

For \(f: A \to B\) and \(C \subseteq A,\) the restriction of \(f\) to \(C\) is

\[\begin{equation} f|_C =\{(x, f(x)): x \in C\}. \end{equation}\]

Theorem 109 (Definition by recursion)

Let \(X\) be a well-ordered set and \(Y\) be any set. Then for any function \(G: \mathbb{P}(X \times Y) \to Y,\) there exists a function \(f: X \to Y\) such that

\[\begin{equation} f(x) = G(f|_x) \end{equation}\]

for all \(x \in X.\)

Proof: Definition by recursion

We prove this result via a trick notion of an attempt.

Definition 128 (Attempt)

Let \(I\) is an initial segment of \(X.\) We say \(h: I \to Y\) is an attempt if

\[h(x) = G(h|_{I_x})\]

for all \(x \in I.\)

Any two attempts that are defined at some \(x\) must be equal at \(x.\)

Lemma 53 (Attempts agree)

If attempts \(h\) and \(h'\) are defined at \(x,\) then \(h(x) = h'(x).\)

Proof: Attempts agree

Note that \(h\) and \(h'\) must agree on the least element of \(X.\) Therefore by induction it suffices to show that if \(h(y) = h'(y)\) for all \(y < x,\) then \(h(x) = h'(x).\) Suppose \(h(y) = h'(y)\) for all \(y < x.\) Then \(h|_{I_x} = h'|_{I_x}.\) By the definition of attempt, we then have

\[h(x) = G(h|_{I_x}) = G(h'_{I_x}) = h'(x).\]

Now we show that for any \(x \in X\) there must exist an attempt \(h\) that is defined at \(x.\)

Lemma 54 (An attempt exists for each \(x\))

For any \(x \in X,\) there must exist an attempt \(h\) that is defined at \(x.\)

Proof: An attempt exists for each \(~x\)

We proceed by induction, and assume that for each \(y < x,\) there exists an attempt \(h_y\) defined at \(y.\) Then we combine all these functions together and take \(h' = \cup_{y < x} h_y.\) This is a well-defined function since attempts agree, and it is defined for all \(y < x.\) Now, add \((x, G(h'|_{I_x}))\) to \(h',\) to obtain an attempt \(h\) that is defined at \(x.\) We therefore have an attempt that is defined at \(x.\)

Given this result, define \(f: X \to Y\) by \(f(x) = y\) if there exists an attempt \(h,\) defined at \(x\) with \(h(x) = y.\) Finally, we note that by its definition using attempts, this \(f\) is unique, concluding the proof.

Lemma 55 (Subset collapse)

Let \(X\) be a well-ordering and let \(Y \subseteq X.\) Then \(Y\) is isomorphic to an initial segment of \(X.\) Moreover, this initial segment is unique.

Proof: Subset collapse

Let \(X\) be a well-ordering and let \(Y \subseteq X.\) We will construct \(f: Y \to X\) to be an order preserving bijection with an initial segment of \(X.\) Specifically, we will map each \(x \in Y.\) to the smallest element of \(X\) that we have not yet mapped to. Let

\[\begin{equation} f(x) = \min (X \setminus \{f(y): y < x\}). \end{equation}\]

The minimum is well-defined because \(\{f(y): y < x\} \neq X.\) This is because \(f(z) < x\) for all \(z < x\) by induction, so \(x \not \in \{f(y): y < x\}.\) Therefore, by Theorem 109, the result holds.

Definition 129 (Isomorphic notation)

We write \(X \leq Y\) if \(X\) is isomorphic to an initial segment of \(Y.\) We write \(X < Y\) if \(X \leq Y\) but \(X\) is not isomorphic to \(Y,\) and similarly for \(X > Y.\)

Theorem 110 (Well-orderings are ordered)

Let \(X, Y\) be well-orderings. Then either \(X \leq Y\) or \(Y \leq X.\)

Well-orderings are ordered

We attempt to define an isomorphism \(f: X \to Y\) by

\[\begin{equation} f(x) = \min (Y \setminus \{f(y): y < x\}). \end{equation}\]

Now, depending on whether \(Y \setminus \{f(y): y < x\} = \emptyset\) for some \(x \in X,\) this function is either well-defined or not. If it is well-defined, then it is an isomorphism from \(X\) to an initial segment of \(Y.\) If it is not well-defined, there exists an \(x \in X\) such that \(\{f(y): y < x\} = Y.\) In this case, \(f\) is a bijection between \(I_x = \{y: y < x\}\) and \(Y.\) So \(f\) is an isomorphism between \(Y\) and an initial segment of \(X.\)

Theorem 111 (Sufficient condition for isomorphic well-orderings)

Let \(X, Y\) be well-orderings. If \(X \leq Y\) and \(Y \leq X,\) then \(X\) and \(Y\) are isomorphic.

Solution: Sufficient condition for isomorphic well-orderings

Let \(X, Y\) be well-orderings with \(X \leq Y\) and \(Y \leq X.\)

Since \(X \leq Y,\) there is an order preserving function \(f: X \to Y\) that is a bijection between \(X\) and an initial segment of \(Y.\) Similarly, since \(Y \leq X,\) there is an order preserving function \(g: Y \to X\) that is a bijection between \(Y\) and an initial segment of \(X.\)

Now \(g \circ f: X \to X\) is a bijection between \(X\) and an initial segment of \(X.\) It is not possible to have a bijection between a set and a proper subset of the set, it must be that \(g \circ f\) is a bijection between \(X\) and itself. Similarly, \(f \circ g\) must be a bijection between \(Y\) and itself.

Therefore, both \(f\) and \(g\) must be bijections, and \(X\) and \(Y\) are isomorphic.

New well orderings#

Definition 130 (Successor)

Given a well-ordering \(X\) and some \(x \not \in X,\) we define a well-ordering on \(X \cup \{x\}\) such that \(y < x\) for all \(y \in X,\) and call this the successor of \(X,\) written \(X^+.\)

Definition 131 (Extension)

Given well-orderings \((X, <_X)\) and \((Y, <_Y),\) we say that \(Y\) is an extension of \(X\) if \(X\) is a proper initial segment of \(Y\) and \(<_X\) and \(<_Y\) agree when defined.

Definition 132 (Nested family)

We say well-orderings \(\{X_i: i \in I\}\) form a nested set if for all \(i, j \in I\) with \(i \neq j,\) either \(X_i\) extends \(X_j\) or \(X_j\) extends \(X_i.\)

Lemma 56 (Nested family has union well-ordering)

Let \(\{X_i: i \in I\}\) be a nested set of well-orderings. Then, there exists a well-ordering \(X\) with \(X_i \leq X\) for all \(i \in I.\)

Proof: Nested family has union well-ordering

Let \(\{X_i: i \in I\}\) be a nested set of well-orderings. Define \(X = \cup_{i \in I} X_i,\) with the relation \(<\) defined on \(X\) as \(\cup_{i \in I} <_i,\) where \(<_i\) is the ordering of \(X_i.\) This is a total ordering. Since \(\{X_i: i \in I\}\) is a nested family, each \(X_i\) is an initial segment of \(X.\)

To show this is a well-ordering, let \(S \subseteq X\) be a non-empty subset of \(X.\) Then \(S \cap X_i\) must be non-empty for some \(i \in I.\) Since \(X_i\) is a well-ordering, \(S \cap X_i\) has a minimum element, say \(x.\) Then, for any \(y \in S,\) we must have \(x \leq y.\) Otherwise, if \(y \in S\) and \(y < x,\) we would have \(y \in X_i\) because \(X_i\) is an initial segment of \(X.\) Therefore, \(S\) has a least element and \(X\) is a well-ordering.

Ordinals#

Definition 133 (Ordinal)

An ordinal is a well-ordering, with any two regarded as the same if they are isomorphic.

Definition 134 (Order type)

If a well-ordering \(X\) has a corresponding ordinal \(\alpha,\) we say \(X\) has order type \(\alpha\) and write \(\text{otp}(X) = \alpha.\)

Definition 135 (Order type notation)

For each \(k \in \mathbb{N},\) we write \(k\) for the order type of the (unique) well-ordering of size \(k.\) We write \(\omega\) for teh order type of \(\mathbb{N}.\)

Definition 136 (Ordinal inequality signs)

For ordinals \(\alpha, \beta,\) we write \(\alpha \leq \beta\) if \(X \leq Y\) for some \(X\) of type \(\alpha\) and \(Y\) of type \(\beta.\) We define strict inequality sings for ordinals in a similar way.

Definition 137 (Initial segment of set of ordinals)

We write \(I_\alpha = \{\beta: \beta < a\}.\)

Lemma 57 (Ordinals smaller than an ordinal form a well-ordering)

Let \(\alpha\) be an ordinal. Then the ordinals \(< \alpha\) form a well-ordering of order type \(\alpha.\)

Proof: Ordinals smaller than an ordinal form a well-ordering

Let \(\alpha\) be an ordinal. Let \(X\) have order type \(\alpha.\) By Definition 129, the well orderings \(< X\) are exactly those well-orderings which are isomorphic to some proper initial segment of \(X.\) These are the sets \(I_x\) for some \(x \in X.\) We can therefore form a bijection between \(X\) and the well-orderings \(< X\) via the function that maps \(x \in X\) to \(I_x.\) This function is one-to-one and order preserving. Therefore \(\alpha\) is isomorphic to \(X\) which is isomorphic to the collection of ordinals of size \(< \alpha.\)

Theorem 112 (Non-empty set of ordinals has least element)

Let \(S\) be a non-empty set of ordinals. Then \(S\) has a least element.

Theorem 113 (Burali-Forti)

The ordinals do not form a set.

Proof: Burali-Forti

Suppose the ordinals form a set \(X.\) Then they also form a well-ordering. Let the order type of \(X\) be \(\alpha.\) Then \(X\) is to \(I_\alpha,\) which is a proper initial segment of \(X.\) This is a contradiction.