You are on page 1of 7

Elements of Set Theory

Chapter 6: Cardinal Numbers and The Axiom of


Choice
Equinumerosity
March 18 & 20, 2014

Lecturer: Fan Yang

1/28

2/28

What is the size of a set?

X ={
Y ={

,
,

}
,

= {0, 1, 2, 3, 4, . . . }
Do A and B have the same size?
Does A have more elements than B?

Are there exactly as many houses as people?


Yes, since there are 5 houses and 5 people.
Yes, since there is a one-to-one correspondence between the two
sets.

Y has more elements than X .


The infinite set has more elements than the finite sets X and Y .
Do and Z have the same size? What about Q and Z?
0
-3

-2 -1
32

-3

-2

-1

1
2

Definition 6.1

A set A is said to be equinumerous or equipotent to a set B (written


A B) iff there is a bijection from A onto B.

11
4

A bijection from A onto B is also called a one-to-one correspondence


between sets A and B.

Given two infinite sets A and B, how to compare their sizes?


3/28

4/28

Example 6.1: Let A = {a, b, c, d} and B = {a, b, c}. Then A 6 B, since


there is no bijection from A onto B. In general, for any two finite sets X
and Y , if Y X , then X 6 Y .
Consider the following infinite sets. Clearly, Even .
= {0, 1, 2, 3, 4, 5, 6, . . . },
Even = {0, 2, 4, 6, 8, 10, 12, . . . } = {2n | n }.

Example 6.2: Even.


Are there exactly as many knives as forks?
Proof. The function f : Even defined by taking

Yes, as there is a one-to-one correspondence between knives and


forks.

f (n) = 2n
is a bijection. Indeed, f is injective, since n 6= m = 2n 6= 2m.
f is surjective, since for all m Even, m = 2n for some n , and
f (n) = 2n = m.
5/28

6/28

Example 6.3: [Galileo] Similarly, Sq, where


Example 6.5: .

= {0, 1, 2, 3, 4, 5, . . . },

Proof. The function f : defined as


1
f (m, n) = [(m + n)2 + 3m + n]
2
is a bijection.

Sq = {0, 1, 4, 9, 16, 25, . . . } = {n2 | n }.


since there is a bijection f : Sq defined as f (n) = n2 .
Example 6.4: \ {0}, since the function f : \ {0} defined as
f (n) = n+ is a bijection.
0

10

6
1

\ {0}

Remark: For infinite sets A, B,

A B 6= A 6 B !

0
7/28

11
12
7
4

13

14

8/28

Example 6.6: Q .
Example 6.7: (0, 1) R, where (0, 1) = {x R | 0 < x < 1}.
Proof. In the following picture, we specify a bijection f : Q.
3
1

3
2

3
3

3
4

3
5

[10] 2

2
2

2
3

[9] 2

2
5

[1] 1

1
2

[2] 1 [8] 1

[0] 0

0
2

1
5

0
3

0
4

0
5

[4] 1

12 [3] 13 [7] 14

15

[5] 2
1

22

23 [6] 24

25

31

32

33

35

34

Proof. By picture. Here (0, 1) has been bent into a semicircle with
center P. Each point in (0, 1) is paired with its projection (from P) on
the real line.
0

... ...

10/28

9/28

Recall : for any sets X and Y ,

An alternative proof: Claim: the mapping f : (0, 1) R defined by


putting
1
1
f (x) =
x
1x
is a bijection.
1
1
For any a, b (0, 1), if f (a) = f (b), namely, if a1 1a
= b1 1b
, then


1
1
(b a) ab
+ (1a)(1b)
= 0 thereby a = b. Hence f is injective.
For any r R, there is xr =

2
r+

4+r 2 +2

Y = {f : X Y | f is a function}.

Example 6.8: For any set A, we have A A 2.


Proof. Define a function H : A A 2 by taking
H(B) = fB ,
where fB : A {0, 1} is the characteristic function of B defined as
(
1 if x B,
fB (x) =
0 if x A \ B.

(0, 1) such that



1
1
1
1
f (xr ) =

=
2
xr
1 xr
xr
1 xr

2
r + 4 + r 2 r + 4 + r2 + 2

2
r + 4 + r2

(r + 4 + r 2 )2 22

=
= r.
2(r + 4 + r 2 )

Claim: H is a bijection.
Indeed, H is injective, since for any B, C A,
H(B) = H(C) = fB = fC = x A(fB (x) = fC (x))
= x A(x B x C) = B = C;
H is surjective, since for any function g A 2, there is
B = {x A | g(x) = 1} A such that H(B) = fB = g.

Hence f is surjective.
11/28

12/28

Theorem 6B (Cantor 1873)


(a) The set is not equinumerous to the set R of real numbers.
(b) No set is equinumerous to its power set.

Theorem 6A

Proof. (a) [Diagonal argument] Given any map f : R. We show


that there exists z R such that z
/ ran f .

For any sets A, B and C:


(a) A A.

f (0) = 236. 0 0 1 2 . . .

(b) If A B, then B A.

f (1) = 7. 7 3 7 4 . . .

(c) If A B and B C, then A C.

f (2) =

3. 1 4 1 5 . . .

f (3) =
..
.

0. 5 2 4 6 . . .

Proof. Exercise.
That is, is an equivalence relation on the class of all sets.

13/28

The integer part of z is 0, and the (n + 1)st decimal place of z is 3


unless the (n + 1)st decimal place of f (n) is 3, in which case the
(n + 1)st decimal place of z is 4. For example, in the case shown,
z = 0.3433 . . .
Clearly, z 6= f (n) for all n, as it differs from f (n) in the (n + 1)st decimal
/ ran f .
place. Hence z

14/28

(b) Suppose g : A A is surjective. Let


B = {x A | x
/ g(x)}.
Since B A and g is surjective, there exists x0 A such that
g(x0 ) = B.

Finite Sets

But then, by the definition of B,


x0 B x0
/ g(x0 ) = B,
which is a contradiction. Hence there is no surjection from A onto A,
thus A 6 A.
Remark: We will soon be able to prove R .
R has smaller size than R, which has smaller size than R, etc.

15/28

16/28

Definition 6.2

A set is finite iff it is equinumerous to some natural number. Otherwise


it is infinite.

That is, for any set A:


A is finite iff n (A n)
A is infinite iff n (A 6 n)

Example 6.9: The set A = {a0 , a1 , a2 , a3 } is finite, since A 4 via the


bijection f : 4 A defined as: f (n) = an .
4 = {0, 1, 2, 3}
A = {a0 , a1 , a2 , a3 }

Example 6.10: Any finite set A is not equinumerous to an infinite set B.


Proof. Suppose A B. As A is finite, A n for some n . From the
transitivity of , it follows that B n, contradicting B being infinite.

There are finitely many forks in the picture. Why?


Because there are exactly 6 forks. This, in turn, is because there is a
one-to-one correspondence between the natural number
6 = {0, 1, 2, 3, 4, 5} and the set F of forks, or in other words, 6 F .

Next, we want to check that each finite set A is equinumerous to a


unique natural number n. This requires the Pigeonhole Principle.
17/28

Proof. (of Pigeonhole Principle) We shall show that for any n , if f : n n


is a one-to-one function, then ran f = n (not a proper subset of n), namely,
that the set T is inductive:

Pigeonhole Principle:
If n items are put into m
pigeonholes with n > m,
then at least one pigeonhole must contain more
than one item.

T = {n | ran f = n for any one-to-one function f : n n}.


Base case: 0 T , since the only function f : 0 0 is the empty function
f = , and ran = .
Inductive step: Assume k T and f : k + k + is a one-to-one function. We
show that ran f = k + . Note that f  k is a one-to-one function from k into k + .

Here n = 10 and m = 9.

Case 1: ran (f  k ) k . Then f  k : k k is a one-to-one function, as


f : k + k + is 1-1. Thus, the assumption k T implies ran f  k = k .
Moreover, since f is 1-1, we must have f (k ) = k . Hence ran f = k {k } = k + .
Case 2: ran (f  k ) * k . Then f (p) = k for some p k . Define f : k + k + as

Pigeonhole Princinple
No natural number is equinumerous to a proper subset of itself.
4 = {0, 1, 2, 3}

18/28

f (p) = f (k ),
f (k ) = f (p) = k ,

Pigeons

f (x) = f (x) for other x k + .


x = {3, 1, 0}

Pigeonholes

Then f : k + k + is one-to-one, and f (k ) = k . By Case 1,


k + = ranf = ranf .

In particular, n 6 m for any m n.


19/28

20/28

Corollary 6C

Corollary 6E

No finite set is equinumerous to a proper subset of itself.

Any finite set A is equinumerous to a unique natural number n, called


the cardinal number or the cardinality of A, denoted by card A or |A|.

Proof. Let A be an arbitrary finite set. Suppose f : A B is a bijection


for some B A. Let g : A n be a bijection for some n .
g 1

The function g f
is a bijection from n onto a proper subset g[B]
of n (see picture on blackboard), contradicting the Pigeonhole
Principle.

Corollary 6D

Proof. Assume that A m and A n for some natural numbers m, n.


It follows that m n. Thus m 6 n and n 6 m, which by trichotomy and
Corollary 4M implies that m = n is the case.
It follows from the corollary that

A n |A| = n.

For example, since n n, we have |n| = n. If a, b, c, d are distinct


objects, then |{a, b, c, d}| = 4, as {a, b, c, d} {0, 1, 2, 3}.

(a) Any set equinumerous to a proper subset of itself is infinite.


(b) The set is infinite.

Clearly, for finite sets A, B:


|A| = |B| (|A| = n |B| = n) (A n B n) A B.

Proof. (a) follows immediately from Corollary 6C.


(b) follows from (a), since e.g. \ {0}, via the bijection
f : \ {0} defined as f (n) = n+ .

22/28

21/28

Definition 6.3
We also want to have cardinal numbers for infinite sets. In fact, what
sets these numbers are is not too crucial, but the essential demand is
that we will define the cardinality |A| for arbitrary set A in such a way
that
|A| = |B| A B
is the case. We postpone until Chapter 7 the actual definition of the set
|A|. The information we need for the present chapter is embodied in
the following promise:
Promise: For any set A, we will define a set |A| in such a way that:
1

|A| = |B| A B,

for a finite set A, the cardinal number |A| is the natural number n
for which A n.

A cardinal number is the cardinality of some set, i.e., is a cardinal


number iff = |A| for some set A.
For example,
Every natural number n is a cardinal number, as n = |n|.
|| is a cardinal number, and name (due to Cantor) || = 0 .
There is a unique set whose cardinality is 0, namely the empty set .
Given a cardinal number > 0, there are many sets A of cardinality .
If one set A of cardinality is a finite set, then all sets of cardinality
are finite sets. In this case, is called a finite cardinal. Otherwise, is
called an infinite cardinal.
Natural numbers are finite cardinals, and they are the only finite
cardinals.
0 , |R|, ||, || are infinite cardinals. Note 0 6= |R| and
0 6= || =
6 || as we have shown 6 R, and 6 6 .

23/28

24/28

Elementary school math:

Cardinal Arithmetic

Definition 6.4 (Addition)


Let and be two cardinal numbers. Define
+ := |K L|,
where K and L are any disjoint sets with |K | = and |L| = .
26/28

25/28

Note: It is always possible to take disjoint sets K , L in the above


definition, as K {0} and L {1} are always disjoint.

Theorem 6H
Assume that K1 K2 and L1 L2 . If K1 L1 = K2 L2 = , then
K1 L1 K2 L2 .

Example 6.11: Prove: 2 + 2 = 4.

Proof. Since K1 K2 and L1 L2 , there are bijections f : K1 K2 and


g : L1 L2 . Define a function h : K1 L1 K2 L2 by taking
(
f (x), if x K1 ,
h(x) =
g(x), if x L1 .

Proof. Let K = 2 {0} and L = 2 {1}. Clearly, K L = and


|K | = 2 = |L|. We need to show that K L 4.
We have that

Since K1 L1 = , h is indeed a function. Claim: h is a bijection.

K L = (2 {0}) (2 {1}) = {(0, 0), (1, 0), (0, 1), (1, 1)}.

For any x1 , x2 K1 L1 such that h(x1 ) = h(x2 ) = y , since K2 L2 = , y is in


exactly one set of K2 and L2 . W.l.o.g., assume that y K2 . Then x1 , x2 K1 .
Since f is injective, x1 = x2 . Hence h is injective.

Clearly, the function f : 4 K L defined by taking


f (0) = (0, 0),

f (1) = (1, 0),

f (2) = (0, 1),

We have that h[K1 L1 ] = h[K1 ] h[L1 ] = K2 L2 , since h[K1 ] = f [K1 ] = K2


and h[L1 ] = g[L1 ] = L2 . Hence h is surjective.

f (3) = (1, 1)

is a bijection.
27/28

28/28

You might also like