Updated Notes
Updated Notes
BY
M. HAIMENE
Definition 1.1. A set is a collection of objects satisfying a certain set of axioms. (These
axioms are stated below.) Each object in the set is called an element of the set.
Remark 1.2. The membership property is the most basic set-theoretic property. We
denote it by ∈. Thus we read X ∈ Y as “X is an element of Y ” or “X is a member of
Y ” or “X belongs to Y ”.
Since the axioms form our definition of a set, we need an axiom to postulate that
sets indeed do exist. More specifically, that at least one set exists. Predicate logic was
designed for developing formal axiomatic theories.
As you have seen in S3512MS Sets and logic, formal logic’s includes:
• logical symbols which include equality symbol =.
• propositional connectives such as:
(a) ∧ “and”
(b) ∨ “or”
(c) ¬ “ not”( negation)
(d) =⇒ “ if, then”
(e) ⇐⇒ , “if and only if”
• quantifiers such as: ∀ “for all”, ∃ “ there exists” (for some) and ∃! “ there exists
a unique”.
Notice that in this course we will use lower case letters for sets. We do this because
the elements of a set will also be sets. Thus a distinction of a set A and its elements a
no longer applies. We will use upper case letters, like R and P for properties that de-
fine sets. We will now present the axioms and derive the most basic elements of set theory.
∃x∀y¬(y ∈ x)
¬.
Axiom of Extensionality If every element of a is an element of b and every element of
b is an element of a, then a = b. In other words, two sets are equal iff they contain the
same elements.
∀a∀b[∀z((z ∈ a) ⇐⇒ (z ∈ b)) =⇒ (a = b)]
1 {1}
0
If we introduce edge 0 ← {1}, this will violate the axiom of extensionality, since 2 6= {1}
although they have the same elements (edges).
Now we know the existence of a set and the extensionality of two sets. So we have the
following result on the uniqueness of a set.
Lemma 1.4. (Uniqueness of sets) There exists one set with no elements.
3
Proof. Suppose there exist two sets with no elements, say a and b. Then x ∈ a =⇒ x ∈ b.
Also x ∈ b =⇒ x ∈ a. Thus by axiom of extensionality, a = b.
Definition 1.5. (Empty set) The unique set with no elements is called empty set, denoted
by ∅.
Axiom of pairing: For any sets a and b there is a set c whose only elements are a
and b.
∀a∀b∃c∀x[(x ∈ c) ⇐⇒ ((x = a) ∨ (x = b))].
By extensionality again, there is for given a and b only one such set c, i.e. by the Axiom
of Extensionality we have {x, y} = {y, x}.
Definition 1.6. For any set a of the universe U there is a set whose only element is a.
This set is called the singleton {a}.
The unordered pair of a and b is the set whose only elements are a and b, denoted by
{a, b}.
An ordered pair (a, b) of a and b is defined by {{a}, {a, b}}.
Definition 1.7. Let a and b be sets. The cartension product of a and b is a set defined
by
a × b = {(x, y) : x ∈ a and y ∈ b}.
One easily proves the following theorem (the proof is left as an exercise).
Theorem 1.8. One has that (a, b) = (a0 , b0 ) if and only if a = a0 and b = b0 .
Axiom Schema of Comprehension: Let P (x) be a property of x. Then for any set
a there is a set b which consists exactly of those elements x of a for which the property
P (x) holds.
∀a∃b∀x[x ∈ b ⇐⇒ (x ∈ a ∧ P (x))].
Lemma 1.9. For every a, there is a unique set b such that x ∈ b if and only if x ∈ a and
P (x).
Proof. Suppose b0 is another set such that x ∈ b0 if and only if x ∈ a and P (x).
If x ∈ b implies x ∈ a and P (x), then x ∈ b0 . Similarly, if x ∈ b0 implies x ∈ a and P (x),
then x ∈ b. Thus we have x ∈ b if and only if x ∈ b0 . Therefore b = b0 .
4 BY M. HAIMENE
Axiom of Union: For any set a there is a set b whose members are precisely the
members of members of a, i.e. For any a, there exists b such that x ∈ b if and only if
x ∈ z for some z ∈ a.
a ∪ b = {x|x ∈ a or x ∈ b}.
Let a and b be sets. We say that a is a subset of b if every element of a is also an element
of b:
(a ⊆ b) ≡ ∀x[x ∈ a =⇒ x ∈ b].
Proposition 1.12. If c is a set such that every element of c is an ordered pair, then ∃a, b
sets such that c ⊆ a × b.
Proof. Let z ∈ c. Then c = (a, b) = {{a}, {a, b}} for some a, b. Thus, the elements of c are
S
sets. Consider d = z∈c z. For any z = (a, b) ∈ c as above, {a} ∈ d and {a, b} ∈ d. We
S
notice that the elements of d are again sets, so we take f = x∈d x. For any z = (a, b) ∈ c,
a ∈ f and b ∈ f . Set a = f = b. Then c ⊆ a × b.
Axiom of Power Set: Let a be a set of the universe U. Then there is a set b whose
elements are precisely the subsets of a.
∀a∃b∀x[(x ∈ b) ⇐⇒ (x ⊆ a)].
5
Remark 1.13. The set b is called the power set of a and we use the notation b = P(a) .
For example, P (∅) = {∅}, P ({∅}) = {∅, {∅}}, P ({∅, {∅}}) = {∅, {∅}, {{∅}}, {∅, {∅}}}.
The Axiom Schema of Replacement: Let P (x, y) be a property such that for
every x there is a unique y for which P (x, y) holds. Let a be any set. Then there is a set
b such that y ∈ b holds if and only if there is some x ∈ a such that the property P (x, y)
holds.
∀a∃b∀y[(y ∈ b) ⇐⇒ ∃x(x ∈ a ∧ P (x, y))].
Remark 1.14. The Axiom Schema of Replacement aims to correct some of the paradoxes
that arise out of the use of the Axiom Schema of Comprehension. The key difference
between the two is that the property P (x, y) [in Replacement] depends both on x as well
as the unique y for which P (x, y) holds, whereas P (x) [in Comprehension] only depends
on x.
The existence of the union of a set a was stipulated as an axiom. We don’t need a
further axiom for the intersection.
Definition 1.15. (The Intersection of a set) Let a be non-empty set. Then there is a set
b whose members are precisely the members of all members of a.
Theorem 1.18. Given two sets a and b, ∃!c such that x ∈ c if and only if x ∈ a and
x ∈ b.
6 BY M. HAIMENE
Definition 1.19. (Complement) The complement of a and b is the set of all x ∈ a such
that x ∈
/ b. We denote it by a \ b.
∀x(x ∈ a ∧ x ∈
/ b).
Theorem 1.20. Given two sets a and b, ∃!c such that x ∈ c ⇐⇒ x ∈ a ∧ ¬(x ∈ b).
Proof. Use the Axiom of Comprehension to postulate the existence of such a c, and then
the Axiom of Extension to prove its uniqueness. (Exercise).
We will revisit the next two Axioms in more depth. Inductive sets will be defined later
in the next section. They are crucial in defining the set of natural numbers.
The Axiom of Foundation: Every non-empty set a contains a set b which is disjoint
to a.
∀a[∃x(x ∈ a) =⇒ ∃b ∈ a(a ∩ b = ∅)].
Axiom of infinity: There is an inductive set. We further prove few results relateted to
the properties of sets.
S
For (c) Note that ∀x(x ∈
/ ∅). So ∀x(x ∈
/ ∅).
7
Proof.
[
x∈ (a ∪ b) ⇐⇒ ∃c(x ∈ c ∧ c ∈ (a ∨ b)), by axiom of union
S S
(c) {a, b} = {a, b} and {a} = a (Exercise).
Proof.
[ [
{a, b} = {{a}, {a, b}}, by definition of ordered pair
[
= {a} ∪ {a, b}, since {a, b} = a ∪ b.
= {a, b}.
T T
(d) [(a ⊆ b) ∧ (∃c)(c ∈ a)] =⇒ b⊆ a.
T
Proof. Let x ∈ b ⇐⇒ ∀c(c ∈ b =⇒ x ∈ c), be definition of intersection. Now
by hypothesis, if c ∈ a and a ⊆ b, then c ∈ a =⇒ c ∈ b. Thus ∀c(c ∈ a =⇒ x ∈
T
c). Hence x ∈ a.
T T T
(e) (∃c(c ∈ a) ∧ ∃d(d ∈ b)) =⇒ (a ∪ b) = ( a) ∩ ( b).
8 BY M. HAIMENE
Proof.
\
x∈ b ⇐⇒ ∀c(c ∈ (a ∨ b) =⇒ x ∈ c), by definition of the intersection.
Proof. Let
=⇒ c ∈ P(b)
1.1. Relation and function. Armed with ordered pairs, we can now de
ne a relation
Proof. (a) follows immediate from the definition of relations, since the empty set has no
members.
For (b): Let x ∈ S. Then by hypothesis x ∈ R, hence, also by our hypothesis, ∃y, z
such that x = (y, z). Thus (y, z) = x ∈ S.
9
Definition 1.27. (Domain and range) If R is a relation then the domain of R (in symbols:
DomR) is the set of all x such that, for some y, (x, y) ∈ R.
The range of R (in symbols: RanR) is the set of all y such that, for some x, (x, y) ∈ R.
The field of a relation R (in symbols: F ldR) is the union of its domain and range.
The next theorem shows the existence of the sets, domain, range and field. Without
any confusions, we still continue using small letters for our sets notation.
Proof.
x ∈ Dom(a ∪ b) ⇐⇒ ∃y(x a ∪ b y)
⇐⇒ ∃y(xay ∨ xby)
⇐⇒ ∃y(xay) ∨ ∃y(xby)
Remark 1.30. Similar results of the range are parallel to those of the domain.
Definition 1.31. (Converse) The converse of a relation R (in symbols: R̃) is the relation
such that ∀x, y, xR̃y if and only if yRx.
R̃ = {(x, y) : yRx}.
10 BY M. HAIMENE
Remark 1.32. The converse of a relation is obtained simply by reversing the order of
the members of all the ordered couples which constitute the relation.
The notion which it is natural to introduce next is that of the relative product of two
sets.
Definition 1.33. (Relative product) If R and S are relations then the relative product
of R and S (in symbols: R/S) is the relation which holds between x and y if and only if
there exists a z such that xRz holds, and zSy holds.
Example 1.34. Let R = {(1, 3), (2, 3)} and S = {(3, 1)}. Then
Proof.
⇐⇒ x(a/b)y ∪ x(a/c)y.
Definition 1.36. Let a be a set and let R be a relation from a to a. That is, domR =
a = ranR.
Proof.
⇐⇒ xIda y by transitive
Conversely, from our hypothesis we have that R/R ⊆ R, which follows that
∃z(xRz ∧ zRy) =⇒ xRy ≡ (xRz ∧ zRy) =⇒ xRy (by quantifier logic).
Remark 1.40. If a and b are sets, a function from (or on) a to (or into) b is a relation f
such that Domf = a and such that for each x ∈ a there is a unique element y ∈ b with
(x, y) ∈ f .
The uniqueness condition can be formulated explicitly as follows: if (x, y) ∈ f and (x, z) ∈
f , then y = z.
if f is a function, we shall write f (x) = y instead of (x, y) ∈ f or xf y.
The symbol f : a → b is sometimes used as an abbreviation for ” f is a function from a
to b.”
The set of all functions from a to b is a subset of the power set P(a × b) and it will be
denoted by ba .
For any function f : a → b, Dom(f ) = a and Ran(f ) ⊆ b.
Proof. Since f and g are relations then f ∩ g is also a relation. It remains to show that
f ∩ g is a function.
⇐⇒ y = z by hyothesis.
∀x, y, z(xg/f y ∧ xg/f z) ⇐⇒ ∀x, y, z[∃a(xga ∧ af y) ∧ ∃a(xga ∧ af z)] by definition of relative product
⇐⇒ y = z by hypothesis.
Remark 2.2. Now that we have established the definition of partial orderings and strict
orderings, we can use ≤ and to denote partial orderings and < and ≺ to denote strict
orderings.
Thus (a, ≤) is an ordered pair consisting of a set a and an ordering ≤, and (b, ≺) is a
strictly ordered pair consisting of a set b and a strict ordering ≺.
13
Example 2.3. Consider N. The inclusion ⊆ relation is a partial order on this set.
Proof. Let n, m, k ∈ N.
1. x ∈ n =⇒ x ∈ n. Therefore, n ⊆ n.
There is a close relationship between orderings and strict orderings as we will see in the
next theorem.
Proof. For (a), we need to show that S is asymmetric. Suppose xSy and ySx both hold
for some x, y ∈ a. Then xRy and yRx both also hold. It follows that x = y because R is
antisymmetric. This is a contradiction since x 6= y. Therefore S is asymmetric.
For (b): We need to show that R is antisymmetric. Suppose xRy and yRx both hold for
some x, y ∈ a. Suppose that x 6= y. Then xSy and ySx both hold. This is a contradiction
since S is asymmetric. Therefore x = y, showing that R is antisymmetric.
In fact, for N, the partial order ⊆ is the familiar inequality: ≤. On N this order has an
additional property: given any n, m ∈ N, either n ≤ m or m ≤ n.
Definition 2.5. A partial order (an ordering) ≤ on a set a is called a total order (or
linear) if ∀x, y ∈ a, either x ≤ y or y ≤ x. The pair (a, ≤) is called a linearly ordered set.
Although we have introduced the set of natural numbers N in our examples, yet we do
not know the clear definition natural numbers.
In defining the natural numbers we begin by examining the most fundamental set, the
empty set. We can very easily create a pattern that is a prime candidate for the definition
14 BY M. HAIMENE
Proposition 2.9. Let x be a set, and let x+ be its successor. Then the successor is
non-empty.
Proposition 2.11. Let x be a set, and let x+ be its successor. Then we have x ∈ x+ and
x ⊆ x+ , that is, the set x is at the same time an element and a subset of the set x+ .
Proof. The assertion follows from the definition of the successor, since x ∈ x ∪ {x} = x+
and x ⊆ x ∪ {x}.
Proposition 2.13. Let x be a set, and let x+ be its successor. Then we have x 6= a+ .
Proof. In contrary, assume that there exists a set x such that x+ = x. It follows that
x = x+ = x ∪ {x}. In particular, it follows that the set x is an element of itself, which
/ x , for any set x. Thus x 6= x+ .
contradict the fact that x ∈
Definition 2.14. (Successor Set) Let a be a set fulfilling the following conditions:
(i) The empty set ∅ is an element of the set a.
(ii) If x is an element of the set a, then its successor x+ is also an element of the set a.
Then the set a is called a successor set.
Note that the axiom of infinity guarantees that there exists at least one successor set.
Proposition 2.15. Let S be a non-empty set of successor sets. Then the intersection
\
J= a
a∈S
the successor a+ .
Remark 2.18. A possible concern is whether we can even define such a set from ZF
axioms. Certainly, the property P (n) : n ∈ I for every inductive set I is a valid property
of n. But unless there exists an inductive set, this property will always create the empty
set under the Axiom Schema of Comprehension. The Axiom of Infinity allows us to move
past this obstacle.
Definition 2.19. (Minimal successor) Let ρ be a successor set with the following property:
If S is a successor set, then the set ρ is a subset of the set S.
The successor set ρ is called the minimal successor set.
Proof. Let ρ be the intersection of every successor set. Then ρ is a successor set itself.
For if not, then for some x ∈ ρ, x+ ∈
/ ρ. But since ρ is the intersection of all successor
sets, then for some such successor set, x ∈ ρ but x+ ∈
/ ρ. This is a contradiction of the
definition of successor set. Then ρ is a successor set and is, by construction, a subset of
all successor sets. It is therefore the smallest successor set.
Theorem 2.21. Let ρ be the minimal successor set, and let S be a successor set which is
a subset of the set ρ. Then we have S = ρ.
Proof. Since there exists a minimal successor set then the set ρ is a subset of the set S.
It follows from ρ ⊆ S and S ⊆ ρ that S = ρ.
∀m, n ∈ N(m ≤ n ⇐⇒ m ∈ n ∨ m = n)
(i) Is P (0) true? For P (0) we have ∀k, m ∈ N(k < 0 ∧ m < 0), but since ∃ no
such m ∈ N. Then P (0) holds.
(ii) Suppose p(n) holds. We need to show that P (n + 1) also holds. Consider
P (n + 1) and suppose k < m and m < n + 1 both hold. Then either m < n
or m = n.
If m < n, then k < n since P (n) holds.
If m = n, then k < n + 1. Thus P (n) =⇒ P (n + 1).
(b) Suppose have n < m and m < n. Then by transitivity n < n. Consider the
property Q(n) : n ≮ n. We need to show this holds ∀n ∈ N.
(i) Suppose Q(0) does not hold. Then we have 0 < 0, which by definition is ∅ ∈ ∅
, which is a contradiction to the definition of ∅. Thus Q(0) is true.
(ii) Induction hypothesis: Suppose Q(n) holds. Consider Q(n + 1). Suppose
Q(n + 1) does not hold. Then n + 1 < n + 1, by definition, is n + 1 ∈ n + 1.
We know n + 1 = n ∪ {n}, which implies that n + 1 ∈ n or n + 1 = n.
If n + 1 ∈ n, then n + 1 < n. But since n < n + 1, by transitivity we have
n < n, which contradicts the induction hypothesis.
If n + 1 = n. This is obviously a contradiction. Thus Q(n + 1) holds for all
n ∈ N. Therefore < is asymmetric on N.
(c) Left as an exercise.
(a) An element b is called a lower bound of S if and only if b ∈ U and b ≤ x holds for
all x ∈ S.
(e) A maximum of the set of lower bounds of S (i.e. a greatest lower bound) is called
infimum of S, denoted inf S.
(f) A minimum of the set of upper bounds of S (i.e. a least upper bound) is called
supremum of S, denoted sup S.
Remark 2.25. A set S is bounded above and below if it has an upper bound and lower
bound respectively. Otherwise unbounded.
If a set S has no finite lower bound, we set inf(S) = −∞ and sup(S) = ∞. Also
S = ∅ =⇒ sup S = ∞ and inf S = −∞.
Theorem 2.27. Let E be a set with ordering relation ≤. Then for any S ⊆ E admitting
both sup(S) and inf(S),
S 6= ∅ =⇒ inf S ≤ sup S
Proof. Since S 6= ∅, there exists s ∈ S. But then s ≤ sup S by the fact that the
supremum is an upper bound and inf S ≤ s by the fact that infimum is a lower bound.
The transitivity of ≤ then gives the claim.
Remark 2.28. The above shows that, for non-empty set, the infimum and supremum
are ordered intuitively.
19
Proof. We will prove by using strong induction. Let ∅ 6= X ⊆ N. Suppose X does not
have a < -least element. Then consider the set N \ X.
Remark 2.34. Observe that if (A, <) is a well ordering and B ⊆ A then (B, <) is a well
ordering.
Definition 2.35. (Order-Isomorphisms) The well-orderings (A, <) and (B, <) are order-
isomorphic iff there is a bijection f : A → B such that: x < y iff f (x) < f (y). In this
case, we write (A, <) ∼
= (B, <), and say that f is an order-isomorphism.
Lemma 2.36. If (A, <) is a well ordering and B ⊆ A is an initial segment of A then
there exists a ∈ A such that B = {b ∈ A : b < a}.
Proof. Let a be the least element in A − B. Since B is an initial segment it follows that
B = {b ∈ A : b < a}.
Remark 2.37. The initial segments of A are A itself and all subsets of the form {b ∈ A :
b < a}. for some element a ∈ A.
∀b∀c(c ∈ b ∧ b ∈ x =⇒ c ∈ x).
Definition 3.3. A set is an ordinal number (an ordinal) if it is transitive and well-ordered
by a ∈.
We shall denote ordinals by lowercase Greek letters α, β, . . . The class of all ordinals is
denoted by Ord.
Lemma 3.5. Let A be a transitive set. Then ∈ is a transitive relation on A if and only
if for every a ∈ A, a is a transitive set.
Proof. We know α ∪ {α} is a set of ordinals, so it just remains to check that α ∪ {α}
is transitive. Assume β ∈ α ∪ {α}. If β ∈ α, then β ⊆ α ⊆ α ∪ {α}. Otherwise,
β = α ⊆ α ∪ {α}.
Definition 3.9. The cardinality of a set A, written |A|, is the smallest ordinal α such
that there is a bijection between α and A.
We say that two sets A and B have the same size or have the same cardinality (equipon-
tent), and write |A| = |B|, if there is a bijection f : A → B.
We say that the set A is at most as big as the set B, and write |A| ≤ |B| if there is a
one-to-one function f : A → B.
Lemma 3.10. For any nonempty sets A and B, if |A| ≤ |B| then there is an onto
function g : B → A.
22 BY M. HAIMENE
Proof. This is apparent: all cardinals are ordinals and hence the cardinals form a sub-
collection of a well-ordered collection. Thus any subset of cardinals has a least element,
since it is a subset of ordinals, and the cardinals are well-ordered.
Before we give the proof of the Well-Ordering Principle we have several mathematical
concepts to develop.
Note that a choice function will pick an element out of a set. However since a choice
function can not pick an element out of an empty set (there is nothing in there) it is
useful at times to define for a set S, F(S) = P(S) \ {∅}. Where P(S) is the power set of
S.
Axiom of choice (AC): Every set of sets has a choice function. i.e If S is a set
and I = {Si } is a nonempty family of nonempty subsets of S, then there exists a choice
function, f : I → S such that ∀i ∈ I, f (Si ) ∈ Si .
Axiom of Choice (AC’): For any nonempty family I of nonempty sets Si , the product
Q
i∈I Si is nonempty. i.e if {Si } is a family of non-empty sets indexed by a non empty set
I, then there exists a family {si }, i ∈ I, such that xi ∈ Si for each i ∈ I.
Theorem 4.2. AC ⇐⇒ AC 0 .
23
Lemma 4.4. Zorn’s lemma (ZL). If X is a partially ordered set such that every totally
ordered subset (chain) of X has an upper bound in X, then X contains a maximal element.
Proof.
Theorem 4.5. The axiom of choice (AC), Zorn’s Lemma (ZL), and the Well- Ordering
Principle (WOP) are all equivalent to each other.
Proof.