Homepage › Solution manuals › Karel Hrbáček › Introduction to Set Theory › Exercise 7.2.4
Exercise 7.2.4
If and are ordinals and and , then , , (where , , and are ordinal operations).
Answers
Proof. Suppose that and are disjoint sets where and so that, by the definition of cardinal addition, . We then show the result by constructing a bijection from to the ordinal . First, since , there is a bijective . Similarly there is a bijective since . Now consider any . We then set
noting that this is unambiguous since and are disjoint. If then we clearly have by Lemma 6.5.4 so that . On the other hand if then again by Lemma 6.5.4 since , and hence again . This shows that really is a function into .
Next we show that is injective. So consider any and in where .
Case: and are both in . Then since is injective and .
Case: and . Then by Lemma 6.5.4 so that clearly .
Case: and . This is analogous to the previous case.
Case: and are both in . Then by Lemma 6.5.4b since is injective and so that .
Hence in all cases we have , which shows that is injective.
Lastly, we show that is surjective. So consider any in . If then . Since is surjective there is an such that . Then, since , clearly . If then by Lemma 6.5.5 there is an ordinal such that . It then follows that so that by Lemma 6.5.4a. Hence so that there is an such that since is surjective. Then, since , clearly we have . Hence in both cases there is an where . Since was arbitrary, this shows that is surjective.
Thus we have shown that is bijective so that . □
Proof. First, if then
The result also holds when . Hence going forward we can assume that and are nonzero and therefore nonempty.
We then show the result by constructing a bijective . So consider any . It then follows that and so that . We then set . We note that
by Lemma 6.5.4, Exercise 6.5.2, and Exercise 6.5.7. Hence so that is into .
Now we show that is injective. So consider any and in where . Then clearly we have , , , and . We then have
Case: . Then it must be that . We then have
where we have used Lemma 6.5.4b and Exercise 6.5.7b since .
Case: . Without loss of generality we can assume that so that clearly . Then we have
where we have again used Lemma 6.5.4a, Exercise 6.5.2, and Exercise 6.5.7.
Hence in all cases we have , which shows that is injective.
Lastly, we show that is surjective. So consider any ordinal so that . Since , it follows from Theorem 6.6.3 that there is a unique and unique such that . Note that we have since otherwise would imply that
by Lemma 6.5.4 and Exercise 6.5.7, which is impossible since . Hence we have that , and clearly . Since was arbitrary this shows that is surjective.
Thus we have shown that is bijective so that by the definition of cardinal multiplication we have as desired. □
The following two lemmas are straightforward generalizations of Theorems 4.3.9 and 4.3.10, respectively.
Lemma 3. Consider a nonzero ordinals and . Let be a (potentially transfinite) system of at most sets, and let be a system of enumerations for , i.e., for each , is a (potentially transfinite) sequence, and . Then is at most .
Proof. We define a function by simply setting for any and . We show that is onto by considering any so that there is an such that . Then, since is the range of , there is a such that . Hence so that is onto since was arbitrary.
We also have that is well-orderable by Theorem 6.5.8 (for example the lexicographic ordering has order type ). It then follows from Lemma ?? that since is onto. Hence is at most as desired. □
Lemma 4. If is a set with cardinality for some ordinal , then the set of all finite sequences of elements of also has cardinality .
Proof. Let be a bijection from to . We also know from Theorem 7.2.1 that , so let be a bijection from to . For each (i.e. ) we define a transfinite enumeration of , where of course denotes the set of all sequences of elements of of length . We define these enumerations recursively. Clearly for we have so that we can set
Then, having defined , for any , we let and define the sequence of length as follows:
for any .
Clearly each is a sequence of length for any and , but we must show that is in fact an enumeration by showing that it is onto for all . We show this by induction on . Obviously this is true for the trivial case and for the case as well since, for any sequence of length 1 where , we have that there is a an such that since is onto. Hence by definition so that is onto.
Now suppose that is onto and consider any sequence . Now, since is a sequence of length there is an such that by the induction hypothesis. We also have so that there is an such that since again is onto. Now, clearly so that there is an such that since is onto. Now consider any . If then clearly we have since . On the other hand, if , then by definition . Thus since was arbitrary and the cases are exhaustive. This shows that is onto since was arbitrary, hence it is an enumeration.
Clearly we have that and, since we also have an transfinite enumeration (indexed by ) of each it follows from Lemma 3 that
where we have utilized Corollary 7.2.2 since . Also clearly since, for example, the function defined by (for any ) is injective. Thus by the Cantor-Bernstein Theorem we have as desired. □
Corollary 5. If is a nonempty finite set then the set of all finite sequences of elements of is countable.
Proof. First we note that clearly the set is countable (since is finite and is countable) so that is also countable by Lemma 4. Now let be a function from to defined by the identity for any sequence . Note that so that any sequence of elements of is a also a sequence with elements in . Clearly is injective so that .
Since is nonempty there is an . For any consider the finite sequence for any , which is clearly a sequence of length with elements in . Consider then the function from to defined for . Clearly this is a injective function since, for any and in where , we have that and are sequences of different lengths so cannot be equal. Thus we have as well so that by the Cantor-Bernstein Theorem. □
The following is a corollary to Exercise 7.2.3c.
Corollary 6. Suppose that is a set such that for some ordinal . Let denote the set of all finite subsets of . Then .
Proof. First let again denote the set of finite subsets of . Since there is a bijective . We then define by letting for any , noting that clearly and is finite so that . We now show that is bijective.
First consider any and in where . Hence . So consider any so that . Then also so that there is a such that . But since is injective it has to be that so that . Hence since was arbitrary. A similar argument shows that as well so that . This shows that is injective.
Now consider any and let , noting that is a bijective function since is. We claim then that . So consider any so that there is a such that . Then we have by the definition of so that there is a such that . Hence so that . Thus since was arbitrary. Now consider any so that clearly and also . Hence clearly so that since was arbitrary. This shows that so that is surjective since was arbitrary.
We have just shown that is bijective so that by Exercise 7.2.3c. □
Proof. To show this we reference the representation of ordinal exponentiation discussed in Exercise 6.5.16. In that exercise we showed that the set , where for any , can be ordered to be isomorphic to . From this it clearly follows that .
We now construct an injective function from to , where is the set of all finite subsets of and is the set of all finite sequences of elements of . So consider any so that and is a finite subset of . Also clearly is a finite set of ordinals so that there is a unique isomorphism from some natural number to . Clearly then is a finite sequence from to . Thus we have that and , so we set .
To see that this mapping is injective consider and in where . Then there is some where . Let and be the isomorphisms from natural numbers and to and , respectively, as described above. If then clearly . So assume that , from which it follows that and . Since are bijections there is a such that , noting that it has to be that since otherwise we would have . We then have so that . Thus once again , which shows that is injective.
So since is injective it follows that . Suppose first that so it has to be that is infinite so that for some ordinal . Thus by Lemma 6 we have . If then clearly so that . If is finite but nonzero then it is nonempty so that by Corollary 5. Lastly, if is infinite then for some since . Hence by Lemma 4. Thus in all three cases is either a natural number or for some . It then follows from Corollary 7.2.2 that
On the other hand if then it must be that is infinite so that for some ordinal . Thus by Lemma 4 we have as well. If is finite then clearly every subset of is finite so that is finite by Theorem 4.2.8. Hence for some natural number . If is infinite then for some since . We then have that as well by Lemma 6. Hence in either case is a natural number or for some . Hence we have
again by Corollary 7.2.2. Thus in all cases we have shown that as desired. □
Main Problem.
Proof. That follows almost immediately from Lemma 1. We have that , where we have also used property (c) of cardinal numbers after Lemma 5.1.2, and Corollary 7.2.3.
Similarly, follows from Lemma 2. We have that , where we have used property (i) of cardinal numbers following Lemma 5.1.4, and Theorem 7.2.1.
The analogous lemma for ordinal exponentiation (i.e. that for ordinals and ) is evidently not true. As a counterexample consider and . We then have that is countable whereas we know that is uncountable.
However, the somewhat analogous Lemma 7 will help us show the desired result. First, if both and are finite then clearly is also finite so that clearly . On the other hand, if at least one of or is infinite, then have we have that by Lemma 7 as desired, noting that clearly since both and . □