Вы находитесь на странице: 1из 12

1.

Sets and Functions


1.1 Introduction
In English, every word is expressed in terms of the 26 alphabets. They are the
building blocks of the English language. In mathematics, the building blocks are sets
and functions. Everything in mathematics can be expressed in terms of these concepts.

1.2 Basic notions of (naive) set theory


A set is a collection of objects. A set is determined by its members. Let A be a set. If
an object x belongs to A (or x is an element of A), we write x ∈ A . Otherwise we
write x ∉ A . If A and B are sets, we say that A and B are equal (denoted by A = B ) if
A and B contain the same elements. That is,
A = B ⇔ ∀x[ x ∈ A ⇔ x ∈ B ] .
Note that the equality of sets has to be defined before it makes sense. This is
illustrated by the following example.

1.2.1 Example: Let


A = {2,3,5,7}
B = {All prime numbers less than 10}
C = {x : x 4 − 17 x 3 + 91x 2 − 235 x 3 + 210 = 0}
The three sets look quite different, yet they contain exactly the same elements, namely
2, 3, 5 and 7. Thus by the above definition, A = B = C = A . The above example also
illustrates that a set can be defined in various ways.

Let A and B be sets. We say that A is a subset of B if every element of A is an element


of B. We write this as A ⊂ B . In symbols,
A ⊂ B ⇔ ∀x[ x ∈ A ⇒ x ∈ B ] .
From the definition, it is clear that A = B if and only if A ⊂ B and B ⊂ A . This
equivalent condition is often used to prove equalities of sets, especially the more
complicated ones.

For convenience, it is useful to introduce a set that has no elements. We call it the
empty set and denote it by φ . To justify the definition, we need to prove that there is
only one empty set.

1.2.2 Proposition. The empty set is unique. That is, if φ andψ are sets that have no
elements, then φ = ψ .
Proof. Suppose that φ ≠ ψ . Then either φ ⊄ ψ orψ ⊄ φ . Without loss of generality,
assume that ψ ⊄ φ . Thus, there is an object x such that x ∈ψ but x ∉ φ . This
contradicts the fact that ψ has no elements. We conclude that the hypothesis
that φ ≠ ψ is false and hence φ = ψ .

We now introduce some basic set operations. Usually, every set considered is a subset
of some universal set. In the following definitions, we assume that every set is a
subset of a given set Ω .

Let A and B be subsets of Ω . The union of A and B, written A ∪ B , is defined by


A ∪ B = {x ∈ Ω : x ∈ A or x ∈ B} .

The intersection of A and B, written A ∩ B , is defined by


A ∩ B = {x ∈ Ω : x ∈ A and x ∈ B} .
If A ∩ B = φ , we say that A and B are disjoint.

The relative complement of A in B, written B \ A , is defined by


B \ A = {x ∈ X : x ∈ B and x ∉ A} .
For the special case that B = Ω , we write Ω \ A = A c and say that A c is the
complement of A.

The symmetric difference of A and B, written A∆B , is defined by


A∆B = ( A \ B ) ∪ ( B \ A) .

It is useful to visualize these sets using the so-called Venn Diagrams, which usually
consist of intersecting circles with various regions representing certain sets. However,
the Venn Diagram is only a tool to assist our thinking. Derivation solely based on it
cannot substitute formal proofs.

We proceed to investigate some properties of these set operations. These properties


are intuitive and can be visualized easily using Venn Diagrams. Note the special
properties of φ and Ω . To illustrate the method, we only prove a few of the results. The
rest can be proved similarly. Later in this chapter, we will use the method of indicator
function to prove (some of) these results in a simpler way.

1.2.3 Proposition. Let Ω be a given set, and let A, B, C be subsets of Ω .


(1) A ∪ A = A, A ∩ A = A
(2) A ∪ φ = A, A ∩ φ = φ , A ∪ Ω = Ω, A ∩ Ω = A
(3) A ∪ B = B ∪ A, A ∩ B = B ∩ A
(4) A ∪ ( B ∪ C ) = ( A ∪ B ) ∪ C , A ∩ ( B ∩ C ) = ( A ∩ B ) ∩ C
(5) A ∪ ( B ∩ C ) = ( A ∪ B ) ∩ ( A ∪ C ), A ∩ ( B ∪ C ) = ( A ∩ B ) ∪ ( A ∩ C )
(6) A \ B = A ∩ B c
(7) ( A c ) c = A
(8) Ω c = φ
Proof. (3) We prove that A ∪ ( B ∩ C ) = ( A ∪ B ) ∩ ( A ∪ C ) from the definition. For
any object x,
x ∈ A ∪ (B ∩ C)
⇔ ( x ∈ A) ∨ ( x ∈ B ∩ C )
⇔ ( x ∈ A ∨ x ∈ B) ∧ ( x ∈ A ∨ x ∈ C )
⇔ ( x ∈ A ∪ B) ∧ ( x ∈ A ∪ C )
⇔ x ∈ ( A ∪ B) ∩ ( A ∪ C )
Therefore, we conclude that A ∪ ( B ∩ C ) = ( A ∪ B ) ∩ ( A ∪ C ) .

Here we use the logical symbol “ ∧ ” to represent “and” and “ ∨ ” to represent “or”.
For more details of the above proof, see Appendix 1.

1.2.4 Proposition. Using the above notation, we have


(1) A ⊂ A
(2) A ⊂ B, B ⊂ C ⇒ A ⊂ C
(3) φ ⊂ A ⊂ Ω
(4) A ⊂ A ∪ B
(5) A ∩ B ⊂ A
(6) A ⊂ B ⇔ B c ⊂ A c
Proof. (6) By definition, A ⊂ B ⇔ (∀x ∈ Ω)[ x ∈ A ⇒ x ∈ B ] . Since
( x ∈ A ⇒ x ∈ B ) ⇔ ( x ∉ B ⇒ x ∉ A) ,
we have A ⊂ B ⇔ (∀x ∈ Ω)[ x ∉ B ⇒ x ∉ A] . Note that for x ∈ Ω , x ∉ B ⇔ x ∈ B c and
x ∉ A ⇔ x ∈ A c , therefore
A ⊂ B ⇔ (∀x ∈ Ω)[ x ∈ B c ⇒ x ∈ A c ] ⇔ B c ⊂ A c .

The notion of union and intersection can be generalized to more than two sets. Let
Ω be a set. We define the power set of Ω , 2 Ω , to be the set of all subset of Ω . In
symbols,
2 Ω = { A : A ⊂ Ω} .
Now let C be a collection of subsets of Ω . That is, every element of C is a subset
of Ω , or simply C ⊂ 2 Ω . We define the union and intersection of sets in C by

U A = U { A : A ∈ C} =
A∈C
{x ∈ Ω : x ∈ A for some A ∈ C} ,

I A = I{ A : A ∈ C} = {x ∈ Ω : x ∈ A for all A ∈ C} .
A∈C

If C is finite and C = { A1 , A2 ,..., An } , we write, in analogy of the sum of numbers,


n n

U A = U Ai and
A∈C i =1
I A = I Ai .
A∈C i =1

Other notations are introduced as needed.

Next we introduce the important concept of an ordered pair. We have encountered


this before in analytic geometry, where we use an ordered pair of numbers ( x, y ) to
represent a point on the Cartesian plane. In general, an ordered pair of x and y is an
object of the form ( x, y ) , and if ( x1 , y1 ) and ( x 2 , y 2 ) are ordered pairs,
then ( x1 , y1 ) = ( x 2 , y 2 ) if and only if x1 = x 2 and y1 = y 2 . Similarly, for
objects x1 , x2 ,...xn , we can define the (ordered) n-tuple ( x1 , x 2 ,...x n ) .

Let A and B be sets. The Cartesian product of A and B, written A × B , is defined by


A × B = {( x, y ) : x ∈ A, y ∈ B} .
Similarly, the Cartesian product of sets A1 , A2 ,..., An is defined by
A1 × A2 × ... × An = {( x1 , x 2 ,..., x n ) : xi ∈ A,1 ≤ i ≤ n} .
As you may observe, these concepts can be extended to more general cases.

1.2.5 Examples. (1) The classic example of a Cartesian product is the Cartesian
plane R 2 , where R 2 is just R × R .
(2) Let Ω = {1,2,3,4,5,6} be the set of outcomes of throwing a die. Then the set of
possible outcomes of throwing two distinguishable dies can be represented by
Ω × Ω = {(i, j ) : i, j ∈ Ω} .
This set has 6 × 6 = 36 elements. The event that the sum of the numbers is 6 is
represented by the subset
A = {(i, j ) ∈ Ω × Ω : i + j = 6} .
Question: How to represent the outcomes of throwing two indistinguishable dies?

We end this section with some standard notations. Throughout this notes, we
denote N = {0,1,2,...} as the set of natural numbers, Z = {...,−2,−1,0,1,2,...} as the set
m
of integers, Q = { : m, n ∈ Z , n ≠ 0} be the set of rational numbers, R be the set of
n
real numbers, and C = {x + iy : x, y ∈ R} be the set of complex numbers. We have
N ⊂Z ⊂Q⊂R⊂C.

1.3 Functions
In the past, mathematicians thought of “functions” as explicit formulas such
2 ex + e−x
as x + 5 x + 6 , sin x and . The present definition that has evolved is much
2
more abstract and does not require any “formula” at all.

Let X and Y be (non-empty) sets. A function (also called a map, an operation or a


transformation) f from X to Y, denoted by f : X → Y , assigns a unique
element y ∈ Y to each x ∈ X . We call this y the value of f at x, and write y = f (x) . We
say that X is the domain of f and Y is the range of f. The graph1 of f is defined to be
the set
{( x, y ) ∈ X × Y : y = f ( x)} .
If f and g are functions, we say that f and g are equal (denoted by f = g ) if f and g
have the same domain and have the same value at every element of the domain. This
definition is often insufficiently appreciated by students.

1.3.1 Example. Let X = {0,1,2} , Y = {0,1} , Z = {3,4,5} . Define f : X → Z and


g : Y → Z by
f (0) = 3, f (1) = 4, f (2) = 4
g (0) = 3, g (1) = 4
Since f and g have different domains, by the above definition, we have f ≠ g even
if f ( x) = g ( x) for all x ∈ Y . Rather, it is the restriction of f in Y that equals g.

We now introduce some basic terminology concerning functions. Let f : X → Y , and


let A be a subset of X. We define the image of A under f by the set
f [ A] = { y ∈ Y : y = f ( x) for some x ∈ A} .
(You may think of the operation A a f [ A] as a function from 2 X to 2 Y ). Similarly, for
any subset B of Y, we define the pre-image of B under f by the set
f −1 [ B ] = {x ∈ X : f ( x) ∈ B} .
1
In axiomatic set theory, a function is defined by its graph: A function is a relation f ⊂ X × Y such
that for each x ∈ X , there is a unique y ∈ Y such that ( x, y ) ∈ f . We write y = f ( x) if ( x, y ) ∈ f .
See exercise 5.
If for any y ∈ Y , the pre-image of the set { y} under f has at most one element, we say
that f is an injection (or f is one to one). An equivalent condition is that x1 = x 2
whenever f ( x1 ) = f ( x 2 ) .
If for any y ∈ Y , the pre-image of the set { y} under f is non-empty, we say that f is a
surjection (or f is onto). An equivalent condition is that for any y ∈ Y , there
exists x ∈ X such that y = f ( x) . (Prove these!)
A function which is one to one and onto is called a bijection (or f is one-one onto).
Let f : X → Y and g : Y → Z . The composition of f and g, denoted by g o f , is a
function g o f : X → Z defined by
( g o f )( x) = g ( f ( x)) for x ∈ X .

1.3.2 Examples. Define f : [−1,1] → [0,1] by f ( x) = x 2 for x ∈ [0,1] . Then f is onto

since for every y ∈ [0,1] , f ( y ) = ( y ) 2 = y . However, f is not one-to-one since

1 1 1
f ( x) = f (− x) for every x ∈ [0,1] . The image of the set [0, ] is f [0, ] = [0, ] . The
2 2 4
1 1 1 1 1 1
pre-image the set [0, ] is f −1 [0, ] = [− , ] . Thus [0, ] ⊂ f −1 [ f [0, ]] . This is true
4 4 2 2 2 2
in general: For A ⊂ X , A ⊂ f −1 [ f [ A]] . (Prove this! When will A = f −1[ f [ A]] ?)

The following propositions contain some basic properties of general functions. The
proofs are good exercises of set operations.

1.3.3 Proposition. Let f : X → Y be an injection. Then for every A, B ⊂ X ,


(1) f [ A ∩ B ] = f [ A] ∩ f [ B ]
(2) f [ A \ B ] = f [ A] \ f [ B ]
Proof: (1) We prove that the two sets are equal by showing that each is a subset of
another. First we let y ∈ f [ A ∩ B ] . By definition, there exists x ∈ A ∩ B such
that f ( x) = y . But x ∈ A and x ∈ B , thus y ∈ f [ A] and y ∈ f [ B ] . It follows
that y ∈ f [ A] ∩ f [ B ] and f [ A ∩ B ] ⊂ f [ A] ∩ f [ B ] . (Note that this is true for any
function.)
On the other hand, let y ∈ f [ A] ∩ f [ B ] . Then y ∈ f [ A] and y ∈ f [ B ] . Thus, there
exist x1 ∈ A and x 2 ∈ B such that f ( x1 ) = f ( x 2 ) = y . Now since f is injective, we have
x1 = x 2 ( say = x) . Thus x ∈ A ∩ B and y ∈ f [ A ∩ B ] . This shows
that f [ A] ∩ f [ B ] ⊂ f [ A ∩ B ] . Thus f [ A ∩ B ] = f [ A] ∩ f [ B ] .
The proof of (2) is similar.

1.3.4 Proposition. Let f : X → Y . Then for every A, B ⊂ Y ,
(1) f −1 [ A ∪ B ] = f −1 [ A] ∪ f −1 [ B ]
(2) f −1 [ A ∩ B ] = f −1 [ A] ∩ f −1 [ B ]
(3) f −1 [ A \ B ] = f −1 [ A] \ f −1 [ B ]
Proof: Exercises.

1.3.5 Proposition. The composition of injections is an injection, the composition of


surjections is a surjection, and hence the composition of bijections is a bijection.
Proof: Exercises.

Finally, we prove an important proposition concerning the existence of the inverse


function of an injection.

~ ~
1.3.6. Proposition. Let f : X → Y be an injection. Let Z = f [ X ] ⊂ Y . Then the

~
function f : X → Z defined by f ( x) = f ( x) for x ∈ X is a bijection. Moreover, there

exists a unique function g : Z → X such that ( g o f )( x) = x for all x ∈ X . We call g the


inverse of f and write g = f −1 .
Proof. It is clear that f is bijective. Since f is bijective, for every y ∈ Z , there exists a
unique x ∈ X such that f ( x) = y . Define g ( y ) = x . It follows that for all x ∈ X ,
( g o f )( x) = g ( f ( x)) = g ( y ) = x .
The uniqueness of g is clear since the condition ( g o f )( x) = x defines g ( y ) for
every y ∈ Z .

1.4 Indicator Functions


Let Ω be a non-empty set and let A be a subset of Ω . The indicator function (also
called the characteristic function) of A is the function 1 A : Ω → R defined by
1 A ( x) = 1 if x ∈ A and 1 A ( x) = 0 if x ∉ A .
This simple function has a lot of importance uses, for instance in abstract integration
theory and probability. Before we derive its properties, we introduce some notations.

A function is real-valued if its range is R. If f , g : Ω → R , define the sum and


product of f and g, written f + g and fg , by the formulas
( f + g )( x) = f ( x) + g ( x) , ( fg )( x) = f ( x) ⋅ g ( x) for all x ∈ Ω .
It is important to distinguish between f + g and f ( x) + g ( x) : f + g is a function
while f ( x) + g ( x) is a real number. Here we are defining addition of functions by
addition of numbers. The case of product is similar.
We say that f is larger than or equal to g, written f ≥ g , if f ( x) ≥ g ( x) for all x ∈ Ω .

1.4.1 Proposition. Let A, B ⊂ Ω .

(1) 1 A + 1 Ac = 1 (Note: This is a statement of equality of functions. Here, 1 is the

constant function defined by 1( x) = 1 for all x. Then (1) is equivalent to the statement

that for all x ∈ Ω , 1 A ( x) + 1 Ac ( x) = 1 (which is a statement of equality of numbers).)

(2) 1 A∩ B = 1 A ⋅ 1B
(3) 1 A∪ B = 1 A + 1B − 1 A∩ B
In particular, if A ∩ B = φ , then 1A∪ B = 1 A + 1B .
(4) A ⊂ B ⇔ 1 A ≤ 1B
(5) 1 A ⋅ 1 A = 1 A

Proof. (1) For every x ∈ Ω , x ∈ A ⇔ x ∉ A c . Thus 1 A ( x) = 1 ⇔ 1 Ac ( x) = 0 and

1 A ( x) = 0 ⇔ 1 Ac ( x) = 1 . It follows that 1 A ( x) + 1 Ac ( x) = 1 for every x ∈ Ω .

(2) For every x ∈ Ω , x ∈ A ∩ B ⇔ x ∈ A ∧ x ∈ B . Thus


1 A∩ B ( x) = 1 ⇔ 1 A ( x) = 1 ∧ 1B ( x) = 1 ⇔ 1 A ( x)1B ( x) = 1 ⇔ (1 A ⋅ 1B )( x) = 1

(3) Note that by De Morgan’s law, ( A ∪ B ) c = A c ∩ B c . Thus 1( A∪ B )c = 1 Ac ∩ B c . Now by

(1) and (2),

1( A∪ B ) c = 1 − 1 A∪ B and

1 Ac ∩ B c = 1 Ac ⋅ 1B c = (1 − 1 A )(1 − 1B ) = 1 − (1 A + 1B − 1 A∩ B ) .

Equating the two equations, we have


1 A∪ B = 1 A + 1 B − 1 A∩ B .
(4) By definition, A ⊂ B ⇔ (∀x ∈ Ω)[ x ∈ A ⇒ x ∈ B ] . But this is equivalent to
(∀x ∈ Ω)[1 A ( x) = 1 ⇒ 1B ( x) = 1]
Since the range of indicator functions is just {0,1} , this is equivalent to the statement
(∀x ∈ Ω)[1 A ( x) = 1 ≤ 1B ( x) = 1] .
Thus A ⊂ B ⇔ 1 A ≤ 1B .
(5) This is clear.

Proposition 1.4.1 shows how indictor functions convert logical relations to simple
arithmetic. The following proposition, although simple, has remarkable consequences:
It allows us to prove properties of sets using properties of functions, and vice versa.

1.4.2 Proposition. Let A, B ⊂ Ω . Then 1 A = 1B if and only if A = B .


Proof. The direction “ ⇐ ” is clear. For “ ⇒ ”, assume that A ≠ B and
that x ∈ A and x ∉ B . Then 1 A ( x) = 1 and 1B ( x) = 0 ≠ 1 A ( x) . Thus 1 A ≠ 1B .

1.4.3 Examples. (1) We show that A ∪ ( B ∩ C ) = ( A ∪ B ) ∩ ( A ∪ C ) using indicator


functions. Note that

1( A∪ B )∩( A∪C ) = 1 A∪ B ⋅ 1 A∪C


= (1 A + 1B − 1 A ⋅ 1B ) ⋅ (1 A + 1c − 1 A ⋅ 1c )
= 1 A + 1 A ⋅ 1C − 1 A ⋅ 1 A ⋅ 1C + 1B ⋅ 1 A + 1B ⋅ 1C − 1 A ⋅ 1B ⋅ 1C − 1 A ⋅ 1 A ⋅ 1B − 1 A ⋅ 1B ⋅ 1C − 1 A ⋅ 1 A ⋅ 1B ⋅ 1C
= 1 A + 1B ⋅ 1C − 1 A ⋅ 1B ⋅ 1C = 1 A + 1B ∩c − 1 A∩ B ∩C = 1 A∪( B ∩C )

Thus 1 A∪( B ∩C ) = 1( A∪ B )∩( A∪C ) . Therefore, by Proposition 1.4.2, we have

A ∪ ( B ∩ C ) = ( A ∪ B) ∩ ( A ∪ C ) .
(2) The above calculation seems cumbersome, but there are situations where direct
logic falls short. For example, let us show that for any sets A, B, C , we have
A ∩ ( B ∆C ) = ( A ∩ B ) ∆ ( A ∩ C ) .
You (at least I) will quickly get confused if you start considering different cases.
However, the method of indicator function is straightforward:
1( A∩ B ) ∆ ( A∩ C ) = 1A ⋅ 1B ⋅ 1( A∩ C ) c + 1A ⋅ 1C ⋅ 1( A∩ B ) c
= 1A ⋅ 1B ⋅ (1A c + 1C c − 1A c 1C c ) + 1A ⋅ 1C ⋅ (1A c + 1B c − 1A c 1B c )
= 1A ⋅ 1B ⋅ 1C c + 1A ⋅ 1C ⋅ 1B c
= 1A ⋅ (1B ⋅ 1C c + 1C ⋅ 1B C )
= 1A∩ ( B∆C )

(3) As a final example, we derive the inclusion-exclusion formula for a finite union
of sets. Let A1 , A2 ,..., An ⊂ Ω . Then
1 A1 ∪ A2 ∪...∪ An
= 1 − 1( A ∪ A ∪...∪ A )c
1 2 n
(*)
= 1 − 1 Ac ∩ Ac ∩...∩ Ac = 1 − (1 − 1 A1 )(1 − 1 A2 )...(1 − 1 An )
1 2 n

= (1 A1 + ... + 1 An ) − (1 A1 ∩ A2 + ... + 1 An −1 ∩ An ) + −... + (−1) n +11 A1 ∩ A2 ∩...∩ An


This formula leads directly to the following: If A1 , A2 ,..., An are events of a random
experiment, then the probability that at least one of the events occur can be computed
by

P ( A1 ∪ ... ∪ An ) = ( P ( A1 ) + ... + P ( An )) − ( P ( A1 ∩ A2 ) + ... +


P ( An −1 ∩ An )) + −... + (−1) n +1 P ( A1 ∩ ... ∩ An )

To show this, take expectation on both sides of (*) and use the facts that (1) E is a
linear operator and (2) E (1 A ) = P ( A) for all event A.

Appendix 1
The crucial step of the proof of proposition 1.2.3(3) is the third line, which in a more
general form is the following: If p, q and r are statements, then
p ∨ (q ∧ r ) ⇔ ( p ∨ q) ∧ ( p ∨ r ) .
This is a basic principle of logic which we will now derive using truth table:

p q r q∧r p ∨ (q ∧ r ) p∨q p∨r ( p ∨ q) ∧ ( p ∨ r )


T T T T T T T T
T T F F T T T T
T F T F T T T T
T F F F T T T T
F T T T T T T T
F T F F F T F F
F F T F F F T F
F F F F F F F F

Since p ∨ (q ∧ r ) and ( p ∨ q ) ∧ ( p ∨ r ) have the same truth value in all cases, they are
logically equivalent. Thus p ∨ (q ∧ r ) ⇔ ( p ∨ q ) ∧ ( p ∨ r ) . The proofs of other
properties in the proposition depend on other rules of logic, such as
¬( p ∧ q ) ⇔ ¬p ∨ ¬q and p → q ⇔ ¬q → ¬p .

1.5 Exercises
1. Prove Proposition 1.2.3 and 1.2.4 (a) from first principles and (b) using indicator
functions. Note that neither of the two methods is superior to the other.
2. Show that for any sets A, B and C, we have
(a) ( A∆B )∆C = A∆ ( B∆C )
(b) ( A∆B )∆ ( B∆C ) = A∆C
3. Suppose A × B = C × D . Is it true that A = C and B = D ? Assume that the sets are
non-empty.
∞ ∞
4. Let A1 , A2 ,... be a sequence of sets. Let UA
n =1
n and I An be the union and
n =1

intersection of these sets respectively. We define the upper limit of the sequence by
the formula
∞ ∞
lim sup An = I (U Ak ) .
n =1 k = n

Similarly, we define the lower limit of the sequence by the formula


∞ ∞
lim inf An = U (I Ak ) .
n =1 k = n

(The reasons for the names will unfold if you study real analysis.) Show that
(a) x ∈ lim sup An ⇔ x ∈ An for infinitely many n (Note: I did not define what “a set
has infinitely many elements” means. Just use common sense. )
(b) x ∈ lim inf An ⇔ x ∈ An for all but finitely many n (i.e. x ∉ An for at most finitely
many n)

(c) (lim sup An ) c = lim inf Anc

(d) (lim inf An ) c = lim sup Anc

(e) lim inf An ⊂ lim sup An


5. (Recommended) We introduce the concept of a relation. Let A and B be
(non-empty) sets. A relation with domain A and range B is any subset R of A × B . We
write xRy if ( x, y ) ∈ R . If A = B , we say that R is a relation on A.
Let R be a relation on A. We say that R is reflexive if for all x, xRx . R is symmetric
if xRy ⇒ yRx . R is transitive if xRy, yRz ⇒ xRz . A relation that is reflexive,
symmetric and transitive is called an equivalence relation.
(a) Let Ω be a set, define a relation R on 2 Ω by letting ( A, B ) ∈ R if and only if A ⊂ B .
Show that R is reflexive and transitive. Also show that R is anti-symmetric: i.e. if
xRy and yRx , then x = y . (Any relation on a set that is reflexive, transitive and
anti-symmetric is called a partial ordering. Try to give more examples.)
(b) Define a relation ~ on the set of integer Z as follows. Set x ~ y if x and y have the
same reminder when divided by 5. Show that this is an equivalence relation. (This is
called the congruence relation.)
(c) Let ~ be an equivalence relation on a set A. For x ∈ A , define
[ x] = { y ∈ A : x ~ y} .
The set [ x] is called the equivalence class of x. Prove that for any x, y ∈ A ,
either [ x] = [ y ] or [ x] ∩ [ y ] = φ . Thus, show that A can be written as a disjoint union of
equivalence classes.
6. Prove Proposition 1.3.3, 1.3.4 and 1.3.5. Try to generalize Proposition 1.3.3 and
1.3.4 to more than two sets.
7. Show by examples that Proposition 1.3.3 is not true if f is not an injection.
8. Show that for sets A and B, 1 A∪ B = max{1 A ,1B } , 1 A∩ B = min{1 A ,1B } .
(Let f , g : X → Y . The function max{ f , g} : X → Y is defined by
max{ f , g}( x) = max{ f ( x), g ( x)}, x ∈ X . min{ f , g} is defined analogously.)
9. For sets A and B, express 1 A∆B in terms of 1 A and 1B .
10. Define A by
A = {x : x is a set and x ∉ x }
Show that A ∈ A ⇔ A ∉ A . (Hint: Put x = A and use the definition.) This is called
Russel’s Paradox (1901). This paradox shows that “naïve” set theory is not quite
right. A new kind of set theory, called axiomatic set theory, was developed to
exclude such contradictions. In axiomatic set theory, A is a “class” but not a set.

References
Kai Lai Chung, Elementary Probability Theory, 4th ed. Springer, 2006
Paul R. Halmos, Naive Set Theory, Springer, 1974
Yiannis Moshovakis, Notes on Set Theory, 2nd ed. Springer, 2005

Вам также может понравиться