0% found this document useful (0 votes)
7 views47 pages

Dms Notes

The document discusses the concepts of propositions, compound propositions, and their truth values, along with logical connectives and truth tables. It also covers conditional statements, argument validity, predicates, and quantifiers, including universal and existential quantifiers, as well as their negations. Additionally, it presents laws of propositions and examples illustrating logical equivalence and tautologies.

Uploaded by

rushalverma8
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
7 views47 pages

Dms Notes

The document discusses the concepts of propositions, compound propositions, and their truth values, along with logical connectives and truth tables. It also covers conditional statements, argument validity, predicates, and quantifiers, including universal and existential quantifiers, as well as their negations. Additionally, it presents laws of propositions and examples illustrating logical equivalence and tautologies.

Uploaded by

rushalverma8
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 47

Lecture-(1-2)

Proposition: A proposition is a declarative sentence which is either true or false but not
both. Propositions are generally expressed by small alphabets p, q, r, . . ..
Examples: 1: Paris is in France (true),
2: London is in Denmark (false),
3: 2 < 4 (true),
4: 4 = 7 (false).
However the following are not propositions:
1: what is your name? (this is a question),
2: do your homework (this is a command),
3: this sentence is false (neither true nor false),
4: x is an even number (it depends on what x represents),
5: Socrates (it is not even a sentence).
The truth or falsehood of a proposition is called its truth value

Compound Proposition: A proposition that is constructed by combining one or more


propositions is called a compound proposition. We denote compound propositions by capital
alphabets L, M, X, Y, . . .. The propositions in a compound proposition are called primitives.
1. P: If you work hard, then you will get A grade. Here primitives are: p := You work
hard, and q := You will get A grade.
2. Q: Amit is good in study and he plays football every day. Here p := Amit is good in
study, q := Amit plays football everyday.

Connective: Connectives are used for making compound propositions. The main ones are
the following (p and q represent given propositions):
Name Notation Meaning
Negation ¬p not p
Conjunction p∧q p and q
Disjunction p∨q p or q
Exclusive OR p⊕q either p or q, but not both
Implication p⇒q p implies q
Bi-conditional p⇔q p if and only if q

Truth Table: A table showing ouput (truth or falsity) of a proposition from all possible
inputs (all combinations of Truth and False for the inputs). Let p, q be propositions.

1
p q ¬p p∧q p∨q p⊕q p⇒q p⇔q
T T F T T F T T
T F F F T T F F
F T T F T T T F
F F T F F F T T
Conditional statement: Let p and q be two propositions. The conditional proposition
p ⇒ q is the proposition “if p, then q”. A conditional statement has two parts, one is
hypothesis (p) and other is conclusion (q).
Example: If you do your homework, you will not be punished. Here, the hypothesis p :=
“you do your homework” and the conclusion q := “you will not be punished”.
Inverse, Converse, Contra-positive: We can form new conditional propositions from
an existing conditional proposition. These are: Inverse, Converse and Contra-positive. So
if p ⇒ q is a conditional proposition, then inverse is ¬p ⇒ ¬q, converse is q → p, and
contra-positive is ¬q ⇒ ¬p.
Tautology, Contradiction, Contingency: A compound statement which is always true
is called a tautology. A compound statement which is always false, is called a contradiction.
If a compound statement is neither tautology nor contradiction, then it is called contingency.
Example: Let p and q be propositions. Then (p∧¬p), (p∨¬p) and (p∧q) are contradiction,
tautology and contingency respectively. To see this, construct their truth tables.
Example: Construct the truth table for compound proposition (p ∨ ¬q) ⇒ (p ∧ q).
p q ¬q p ∨ ¬q p∧q (p ∨ ¬q) ⇒ (p ∧ q)
T T F T T T
T F T T F F
F T F F F T
F F T T F F

Propositional Equivalence: Two propositions X and Y are logically equivalent or equiv-


alent, denoted as X ≡ Y , if the bi-conditional proposition X ⇔ Y is a tautology or if the
columns giving their truth values agree.
Example: Show that ¬(p ∨ q) ≡ [(¬p) ∧ (¬q)].
p q ¬p ¬q p∨q ¬(p ∨ q) ¬p ∧ ¬q ¬(p ∨ q) ⇔ (¬p ∧ ¬q)
T T F F T F F T
T F F T T F F T
F T T F T F F T
F F T T F T T T
In the above truth table, we see that truth value of ¬(p ∨ q) and [(¬p) ∧ (¬q)] are same (see
columns six and seven) or ¬(p ∨ q) ⇔ (¬p ∧ ¬q) is a tautology. Therefore the propositions
are equivalent.
Exercise: Show that p ⇔ q ≡ (p ⇒ q) ∧ (q ⇒ p).

2
Laws of propositions: Let p, q, r be primitive statements.
1. Double negation: ¬¬p ≡ p.
2. De Morgan’s Laws: ¬(p ∧ q) ≡ ¬p ∨ ¬q and ¬(p ∨ q) ≡ ¬p ∧ ¬q
3. Commutative Laws: p ∨ q ≡ q ∨ p and p ∧ q ≡ q ∧ p.
4. Associative Laws: p ∧ (q ∧ r) ≡ (p ∧ q) ∧ r and p ∨ (q ∨ r) ≡ (p ∨ q) ∨ r.
5. Distributive Laws: p ∧ (q ∨ r) ≡ (p ∧ q) ∨ (p ∧ r) and p ∨ (q ∧ r) ≡ (p ∨ q) ∧ (p ∨ r).
6. Idempotent Laws: p ∧ p ≡ p and p ∨ p ≡ p
7. Identity Laws: p ∧ T ≡ p and p ∨ F ≡ p.
8. Inverse Laws: p ∧ ¬p ≡ F and p ∨ ¬p ≡ T
9. Dominations Laws: p ∨ T ≡ T and p ∧ F ≡ F ,
10. Absorption Laws: p ∨ (p ∧ q) ≡ p and p ∧ (p ∨ q) ≡ p
One can also show the equivalence of propositions by using the laws of propositions. Here
are examples.
Example: Show that (p ∨ q) ∧ ¬(¬p ∧ q) ≡ p.
Solution: (p ∨ q) ∧ ¬(¬p ∧ q)
≡ (p ∨ q) ∧ ¬¬p ∨ ¬q (by De Morgan’s Law)
≡ (p ∨ q) ∧ p ∨ ¬q (by Double negation Law)
≡ p ∨ (q ∧ ¬q) (by Distributive Law)
≡p∨F (by Inverse Law)
≡p (by Identity law)
Example: Show that ¬(p ∨ (¬p ∧ q)) ≡ ¬p ∧ ¬q.
Solution: ¬(p ∨ (¬p ∧ q))
≡ ¬p ∧ ¬(¬p ∧ q) (by De Morgan’s Law)
≡ ¬p ∧ (p ∨ ¬q) (by De Morgan’s Law and Double negation Law)
≡ (¬p ∧ p) ∨ (¬p ∧ ¬q) (by Distributive Law)
≡ F ∨ (¬p ∧ ¬q) (by Inverse Law)
(¬p ∧ ¬q). (by Identity Law)
Exercise: Show that p ⇒ q ≡ ¬p ∨ q.
Example: Show that (p ∧ q) ⇒ (p ∨ q) is a tautology.
Solution: By above exercise: (p ∧ q) ⇒ (p ∨ q) ≡ ¬(p ∧ q) ∨ (p ∨ q)
≡ (¬p ∨ ¬q) ∨ (p ∨ q) (by De Morgan’s Law)

3
≡ (¬p ∨ p) ∨ (¬q ∨ q) (by Associative and Commutative Laws)
≡T ∨T (by Inverse Law)
≡ T. (by Dominations Law)
Thus (p ∧ q) ⇒ (p ∨ q) is a tautology.
Argument and its validity: An argument is a sequence of statements in which the con-
junction of the initial statements (called the premises/hypotheses) p1 , p2 , . . . , pn is said to
imply the final statement (called the conclusion) q.
An argument is valid if the truth of all its premises implies that the conclusion is true or
(p1 ∧p2 ∧. . .∧pn ) → q is a tautology. Here pi ’s are premises or hypothesis and q is conclusion.
Example: Let p, q be primitive propositions. Let P : p and Q : p ⇒ q be premises and q be
conclusion. Check the validity of the argument.
Solution: Let us construct the truth table.
p q p⇒q p ∧ (p ⇒ q) [p ∧ (p ⇒ q)] ⇒ q
T F F F T
T T T T T
F T T F T
F F T F T
Method 1: In the above truth table, we see that there is only one case when both premises
are two (see second row) and in this cases the conclusion is also true. Thus the argument is
valid.
Method 2: Note that [p ∧ (p ⇒ q)] ⇒ q is a tautology, therefore the argument is valid.

Exercise: Let P : p ⇒ q and Q : ¬p be premises and ¬q be conclusion. Show that the


argument is not valid.

4
Lecture-3

Note that the sentence “P (x) := x + 2 = 2x” is not a proposition. However, if we assign
a value for x then it becomes a proposition. As for each value of x the sentence is either
true or false. Thus the sentence can be treated as a function for which input is a value of x
and the output is a proposition. Such sentence is an example of predicate or a propositional
function. We define it more precise way as follows:
Predicate or Propositional function: Let A be a given set. A propositional function
defined on A is an expression P (x) which has the property that P (a) is true or false for each
a ∈ A. That is P (x) becomes a statement whenever x is replaced by any value a ∈ A. In
short, predicate is the part of a sentence that attributes a property to the subject.

The set A is called the domain of P (x), and the set Tp of all elements of A for which P (a)
is true is called the truth set of P (x). In other words, Tp = {x : x ∈ A, P (x) is true}
Example: Find the truth set Tp of each propositional function P (x) defined on the set N.
1. Let P (x) be “x + 5 > 1”. Then Tp = {x : x ∈ N, x + 5 > 1} = N.
2. Let P (x) be “x + 2 > 7”. Then Tp = {x : x ∈ N, x + 2 > 7} = {6, 7, 8, . . .} consists of
all integers greater than 5.
3. Let P (x) be “x + 5 < 3”. Then Tp = {x : x ∈ N, x + 5 < 3} = ∅.
Remark: The above example shows that if P (x) is a propositional function defined on a
set A then P (x) could be true for all x ∈ A, for some x ∈ A or for no x ∈ A. In the next
paragraph, we discuss this quantifiers related notion to such proposition function.
A word which is usually used before noun to express the quantity of object is called quantifier.
Here we discuss few quantifiers which are used in propositional functions.
Universal Quantifier:
Let P (x) be a propositional function defined on a set A. Consider the expression

(∀x ∈ A) P (x) or ∀x P (x)

which reads as “for every x in A, P (x) is true statement. The symbol ∀ which reads “for all”
or “for every” is called universal quantifier. In this case Tp = A (the entire domain).
Existential Quantifier: Let P (x) be a propositional function defined on a set A. Consider
the expression

(∃x ∈ A) P (x) or ∃x P (x)

which reads as “there exists x in A such that P (x) is true statement. The symbol ∃ which

1
reads “there exists” or “for some” or “for at least one” is called existential quantifier. In this
case Tp ̸= ∅.
Precedence of Quantifiers: The quantifiers ∀ (universal quantifier) and ∃ (existential
quantifier) have higher precedence than all logical operators from propositional calculus.
For example, ∀xP (x) ∨ Q(x) is the disjunction of ∀xP (x) and Q(x). In other words, it means
(∀xP (x)) ∨ Q(x) rather than ∀x(P (x) ∨ Q(x)).

Negation of Quantified Statements


Consider the statement “All maps are linear”. Its negation is either of the following equivalent
statements:
“It is not the case that that all maps are linear”
“There exists at least one map which is not linear”.
Symbolically, let S denote the set of all maps. Then the above negation can be written as

¬(∀x ∈ S) (x is linear) ≡ (∃x ∈ S) (x is not linear).

Or when P (x) denotes “ x is linear”,

¬(∀x ∈ S) P (x) ≡ (∃x ∈ S)¬P (x) or ¬∀x P (x) ≡ ∃x ¬P (x).

Thus we have Negating Quantified Expressions:


1. ¬(∀x ∈ S) P (x) ≡ (∃x ∈ S)¬P (x)
2. ¬(∃x ∈ S) P (x) ≡ (∀x ∈ S)¬P (x)
The above rules for negations for quantifiers are called De Morgan’s laws for quantifiers.

Example: What are the negations of the statements ∀x (x2 > x) and ∃x (x2 = 2)?
Solution: The negation of ∀x (x2 > x) is the statement ¬∀x (x2 > x), which is equiva-
lent to ∃x ¬(x2 > x), that is, ∃x(x2 ≤ x). The negation of ∃x(x2 = 2) is the statement
¬ ∃x(x2 = 2), which is equivalent to ∀x ¬(x2 = 2), that is, ∀x(x2 ̸= 2).

Example: Show that ¬∀x(P (x) → Q(x)) ≡ ∃x(P (x) ∧ ¬Q(x)).


Solution: By De Morgan’s law for universal quantifiers, we know that ¬∀x(P (x) → Q(x))
and ∃x(¬(P (x) → Q(x))) are logically equivalent. Since P (x) → Q(x) ≡ ¬P (x) ∨ Q(x), it
follows that ¬∀x(P (x) → Q(x)) ≡ ∃x(P (x) ∧ ¬Q(x)).

Nested Quantifiers: Two quantifiers are nested if one is within the scope of the other.

2
Example: Assume that the domain for the variables x and y consists of all real numbers.
The statement
∀x ∀y (x + y = y + x)
says that x + y = y + x for all real numbers x and y. This is the commutative law for addition
of real numbers. Likewise, the statement

∀x ∃y (x + y = 0)

says that for every real number x there is a real number y such that x + y = 0. This states
that every real number has an additive inverse. Similarly, the statement

∀x ∀y ∀z (x + (y + z) = (x + y) + z)

is the associative law for addition of real numbers.


Quantifications of Two Variables:
• Statement: ∀x ∀y P (x, y) OR ∀y ∀x P (x, y)
When True? P (x, y) is true for every pair x, y.
When False? There is a pair x, y for which P (x, y) is false.
• Statement: ∀x ∃yP (x, y)
When True? For every x there is a y for which P (x, y) is true
When False? There is an x such that P (x, y) is false for every y.
• Statement: ∃x ∀y P (x, y)
When True? There is an x for which P (x, y) is true for every y.
When False? For every x there is a y for which P (x, y) is false.
• Statement: ∃x ∃y P (x, y) OR ∃y ∃x P (x, y)
When True? There is a pair x, y for which P (x, y) is true.
When False? P (x, y) is false for every pair x, y.
Example: We can express that a function f : X → Y is one-to-one using quantifiers as

∀a ∀b f (a) = f (b) → a = b .

Example: A function f : X → Y is onto if

∀y ∃x (f (x) = y).

Example: Use quantifiers to express the definition of the limit of a real-valued function
f (x) of a real variable x at a point a in its domain.
Solution: Recall that the definition of the statement

lim f (x) = L
x→a

is: For every real number ϵ > 0 there exists a real number δ > 0 such that |f (x) − L| < ϵ

3
whenever 0 < |x − a| < δ. This definition of a limit can be phrased in terms of quantifiers by
 
∀ϵ > 0 ∃δ > 0 ∀x 0 < |x − a| < δ → |f (x) − L| < ϵ .

Example: Negate each of the following statement:


1. ∃x ∀y, P (x, y),
2. ∃x ∃y ∀z, p(x, y, z)
Solution:
1. ¬(∃x ∀y, P (x, y)) ≡ ∀x ∃y, ¬P (x, y).
2. ¬(∃x ∃y ∀z, P (x, y, z)) ≡ ∀x ∀y ∃z, ¬P (x, y, z)
Example from “A basic course in Real Analysis by S Kumaresan”: Suppose we have a sen-
tence: “In each tree in the orchard, we can find a branch in which all the leaves are green”.
Let us convert the above sentence as a mathematical sentence: Let T denote the set of all
trees in the orchard. Let t ∈ T be a tree. Let Bt denote the set of all branches of the tree
t. Let b ∈ Bt be a branch of tree t. Let lb denote the set of all leaves on the branch b. Then
the above sentence can be written as:
∀t ∈ T ∃ b ∈ Bt ∀ l ∈ Lb , l is green. The negation is:

¬(∀t ∈ T ∃ b ∈ Bt ∀ l ∈ Lb , l is green) ≡ ∃ t ∈ T ∀ b ∈ Bt ∃ l ∈ Lb , l is not green.

4
Lecture-4

Set: A set is defined as a well defined collection of well defined distinct objects. The objects
are called the elements or members of the set. We denote a set usually by capital letters,
such as, A, B, X, Y, . . ., whereas the lower-case letters a, b, p, q, . . . will usually be used to
denote elements of sets. The set having no element is called empty set or nul set denoted by
ϕ. If x is an element of a set X then we denote it by x ∈ X. The cardinality or the number
of elements in a set S is denoted by |S|.
Let A and B be two sets such that elements of A are also the elements of B then we say
that A is a subset of B, denoted as A ⊆ B. Two sets A and B are said to be equal, written
as A = B, if A ⊆ B and B ⊆ A. Let A be a set. A collection of all subsets of A is called
power set of A, denoted as P (A). If |A| = n, then |P (A)| = 2n .
Examples:
1. Natural numbers N, Integers Z, Rationals Q, Real numbers R.
2. The solution of the equation x2 − 4x + 4.
3. The set of nobel laureates in the world.
4. The set of points in R2 .
5. The people living in India.
Operations on sets Let A and B be two sets. Then:
1. A union B, denoted as: A ∪ B = {x : x ∈ A or x ∈ B}.
2. A intersection B, denoted as: A ∩ B = {x : x ∈ A and x ∈ B}.
3. A minus B denoted as: A − B = {x : x ∈ A and x ∈
/ B}.
4. A complement, denoted as: Ac = {x : x ∈ U and x ∈
/ A}, where U is universal set.
5. The symmetric difference of A and B written as: A ⊕ B = (A ∪ B) − (A ∩ B).
Algebra of sets Let A and B be two sets and U be universal set. Then:
1. Associative Law: (A ∪ B) ∪ C = A ∪ (B ∪ C) and (A ∩ B) ∩ C = A ∩ (B ∩ C)
2. Commutative Law: A ∪ B = B ∪ A and A ∩ B = B ∩ A
3. Distributive Law: A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C) and A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C)
4. De Morgan’s Law: (A ∪ B)c = Ac ∩ B c and (A ∩ B)c = Ac ∪ B c .
5. Identity Law: A ∪ ϕ = A, A ∪ U = U and A ∩ ϕ = ϕ, A ∩ U = A.
6. Complement Law: A ∪ Ac = U , A ∩ Ac = ϕ, U c = ϕ and ϕc = U
7. Involution Law: (Ac )c = A
Inclusion and exclusion principle:

1
1. For two sets A1 and A2 : |A1 ∪ A2 | = |A1 | + |A2 | − |A1 ∩ A2 |.
2. For three sets A1 , A2 and A3 : |A1 ∪ A2 ∪ A3 | = |A1 | + |A2 | + |A3 | − |A1 ∩ A2 | − |A1 ∩
A3 | − |A2 ∩ A3 | + |A1 ∩ A2 ∩ A3 |.
n
S Pn P P
3. General form: | Ai | = i=1 |Ai | − 1≤i<j≤n |Ai ∩ Aj | + 1≤i<j<k≤n |Ai ∩ Aj ∩ Ak | +
i=1
. . . + (−1)n−1 |A1 ∩ A2 ∩ . . . ∩ An |.
Multiset: A multiset is a set in which the multiplicity of an element may be one or more.
The multiplicity of an element is the number of times the element repeated in the multiset.
Operations on multiset: Let A and B be two multisets. Then
1. Union of multisets: The union of two multiset A and B is a multiset C such that the
multiplicity of an element in C is equal to the maximum of the multiplicity of the
element in A and B.
2. Intersection of multisets: The intersection of two multiset A and B is a multiset C such
that the multiplicity of an element in C is equal to the minimum of the multiplicity of
the element in A and B.
3. Difference of multisets: The difference of two multisets A and B is a multiset C such
that the multiplicity of an element in C is equal to the multiplicity of the element in
A minus the multiplicity of the element in B if the difference if positive, and if the
difference is negative multiplicity is considered as 0.
4. Sum of multisets: The sum of two multisets A and B is a multiset C such that the
multiplicity of an element in C is the sum of multiplicity of the element in A and B.
5. Cardinality of multiset: The cardinality of a multiset is the number of distinct elements
in the multiset without considering the multiplicity of an element.
Cartesian product: Let A and B be two sets. Then the cartesian product A × B of the
sets is defined as A × B = {(a, b) : a ∈ A, b ∈ B}. The elements of A × B are called ordered
pairs. Note that if |A| = n, |B| = m, then |A × B| = n.m.
Examples 1: Let A = {1, 2} and B = {a, b, c}. Then A×B = {(1, a), (1, b), (1, c), (2, a), (2, b),
(2, c)}, B×A = {(a, 1), (a, 2), (b, 1), (b, 2), (c, 1), (c, 2)}, and A×A = {(1, 1), (1, 2), (2, 1), (2, 2)}.
Example 2: Let A = R. Then R2 = R × R.

2
Lecture-5
Binary Relation: Let A and B be non-empty sets. A binary relation or simply a relation
R from A to B is a subset of A × B, that is, R ⊆ A × B. If (a, b) ∈ R, then we also say that
a is related to b by R or aRb. If A = B, then we say that R is a relation on A.
The domain of R is a subset of A which are related to some elements in B. The range of R
is set of all element b ∈ B for which there is some element a ∈ A such that aRb. Let A, B
be a sets with |A| = m and |B| = n. Then there are 2mn relations from A to B.
Examples:
1. Let A = {1, 2, 3} and B = {x, y, z}, and let R = {(1, y), (1, z), (3, y)}. Since R ⊆ A×B,
R is a relation from A to B. The domain of R is {1, 3} and the range of R is {y, z}.
2. Let S be a collection of sets. Then set inclusion ⊆ is a relation on A.
3. The divisibility of two numbers in N is a relation on N.
4. Let L be the set of lines in the plane. Then perpendicularity of two lines l1 and l2 in
the plane gives a relation on L.
Complement of relation: Let R be a relation from A to B. The complement of R, denoted
by R̄, is a relation from A to B such that R̄ = {(a, b) : (a, b) ̸∈ R}.
Inverse of relation: Let R be a relation from A to B. The inverse of R, denoted by R−1
is a relation from B to A such that R−1 = {(b, a) : (a, b) ∈ R}.
Composition of relation: Let A, B and C be sets, and let R be a relation from A to B
and S be a relation from B to C, that is, R ⊆ A × B and S ⊆ B × C. Then
R ◦ S = {(a, c) : there exists b ∈ B for which (a, b) ∈ R and (b, c) ∈ S}.
Example: Let A = {1, 2, 3, 4}, B = {a, b, c, d}, C = {x, y, z}. Let R = {(1, a), (2, d), (3, a),
(3, b), (3, d)} and S = {(b, x), (b, z), (c, y), (d, z)} be relations from A to B and from B to C
respectively. Then R ◦ S = {(2, z), (3, x), (3, z)}.
Types of relation: Let A be a set and R be a relation on A.
1. Reflexive Relation: R is reflexive if (a, a) ∈ R, that is, aRa for all a ∈ A.
2. Symmetric Relation: R is symmetrix if aRb then bRa.
3. Antisymmetric Relation: R is called antisymmetric if aRb and bRa then a = b.
4. Transitive Relation: R is called transitive: if aRb and bRc then aRc.
Example. Let A = {1, 2, 3, 4}. Consider the following relations on A.
R1 = {(1, 1), (1, 2), (2, 3), (1, 3), (4, 4)},
R2 = {(1, 1), (1, 2), (2, 1), (2, 2), (3, 3), (4, 4)},
R3 = {(1, 3), (2, 1)},
R4 = ∅, the empty relation,

1
R5 = A × A.
Determine, which of the relations are: (a) reflexive, (b) symmetric, (c) antisymmetric, (d)
transitive.
Solution: Since (2, 2) ∈/ R1 , R3 , R4 . Hence, these relations are not reflexive. Since (a, a) ∈
R2 , R5 for every a ∈ A, R2 and R5 are reflexive.
R1 is not symmetric since (1, 2) ∈ R1 but (2, 1) ∈
/ R1 . Similarly R3 is not symmetric. All
other relations are symmetric.
R2 is not antisymmetric since (1, 2), (2, 1) ∈ R2 but 1 ̸= 2. Similarly R5 . All the other
relations are antisymmetric.
R3 is not transitive since (2, 1), (1, 3) ∈ R3 but (2, 3) ∈
/ R3 . All the other relations are
transitive.
Equivalence Relation: A relation R on a set S is called an equivalence relation if it is
reflexive, symmetric, and transitive.
Examples:
1. Let S be a set of lines in the plane. The relation of parallel is an equivalence relation.
2. The relation of inclusion ⊆ is not equivalence relation. It is reflexive and transitive
but not symmetric, since A ⊆ B does not imply B ⊆ A.
3. Let m be a fixed positive integer. Two integers a and b are said to be congruent moulo
m, written as a ≡ b (mod m), if m divides a − b. This relation of congruence modulo
m is an equivalence relation on Z.
Equivalence Class: Let R be an equivalence relation on a set S. For a ∈ S, the set
[a] = {x : (a, x) ∈ R} is called the equivalence class of a.
The collection of all such equivalence classes is denoted by S/R, that is, S/R = {[a] : a ∈ S}.
The set S/R is also called quotient set of S by R.
Example In the above Example 3, the relation of congruent modulo m on the set of integers
Z. Let m = 5. Then we see that
[0] = {. . . , −10, −5, 0, 5, 10, . . .}, that is, [0] = {5k : k ∈ Z},
[1] = {. . . , −9, −4, 1, 6, 11, . . .}, that is, [1] = {5k + 1 : k ∈ Z},
[2] = {. . . , −8, −3, 2, 7, 12, . . .}, that is, [2] = {5k + 2 : k ∈ Z}.
[3] = {. . . , −7, −2, 3, 8, 13, . . .}, that is, [3] = {5k + 3 : k ∈ Z}.
[4] = {. . . , −6, −1, 4, 9, 14, . . .}, that is, [3] = {5k + 4 : k ∈ Z}.
The above are the only distinct equivalence classes. Thus Z/R = {[0], [1], [2], [3], [4]}.
Theorem 1: Let R be an equivalence relation on a set S.
1. For each a ∈ S, a ∈ [a], that is, every element lies in its own equivalence class.

2
2. For each a, b ∈ S, a R b if and only if [a] = [b], that is, if any two elements are related
by R then they have same equivalence class.
3. For each a, b ∈ S, [a] = [b] or [a] ∩ [b] = ∅.
Proof: Since R is reflexive, a R a for each a ∈ S. So a ∈ [a]. This proves first part.
Second Part: Suppose a R b and x ∈ [a]. Then x R a. Since a R b and R is transitive, x R b.
So x ∈ [a], and [a] ⊆ [b]. Similarly we see that [b] ⊆ [a]. Combining both, we get [a] = [b].
Conversely, let [a] = [b]. This means, if x ∈ [a] then x ∈ [b] and therefore xRa (or aRx since
R is symmetrix) and xRb. Since R is transitive, aRb.
Third Part: let [a] ∩ [b] ̸= ∅ and x ∈ [a] ∩ [b]. Then xRa (so aRx) and xRb imples aRb. By
second part, [a] = [b].
Partition of a set: Let S be a non-empty set. A collection P , containing subsets A1 , A2 , . . .
of S, is called a partition of S if: ∪Ai = S and Ai ∩ Aj = ∅ for i ̸= j.
Example: Let S = {1, 2, . . . , 9}. Consider the following collections of subsets of S.
P1 = [{1, 3, 5}, {2, 6}, {4, 8, 9}],
P2 = [{1, 3, 5}, {2, 4, 6, 8}, {5, 7, 9}],
P3 = [{1, 3, 5}, {2, 4, 6, 8}, {7, 9}].
P1 is not a partition, since 7 ∈
/ P1 . P2 is not a partition, since {1, 3, 5} and {5, 7, 9} are not
disjoint. Note that P3 is a partition of S.
Theorem 2: Let R be an equivalence relation on a nonempty set S. The collection S/R of
all equivalence classes gives a partition of S.
Proof: Proof follows from Theorem 1.

3
Lecture-6

Partial order relation: Let R be a relation on a set X satisfying the following three
properties:
Reflexive: For any a ∈ S, we have aRa.
Antisymmetric: If aRb and bRa, then a = b.
Transitive: If aRb and bRc, then aRc.
Then R is called partial order and the set X with such R is called partially ordered set.
Examples:
1. The relation (≤) “less than or equal to" is partial order on R.
2. The inclusion relation of sets (⊆) is partial order on a collection of sets.
3. The relation divisibility (|) is a partial order on N.
4. The ralation divisibility (|) is not a partial order on Z. As, 2| − 2 and −2|2 but 2 ̸= −2.
Totally ordered set: A partial order R on a set X is called total ordering if given every
pair of x, y ∈ X, either xRy or yRx. A set X with total ordering is called totally ordered
set or a chain.
Examples:
1. R is a chain with the relation ≤. What if we replace ≤ by <?
2. A collection of sets with the inclusion relation ⊆ is not a chain.
3. The set S = {2, 5, 6, 8, 9, 10} is not a chain with the divisibility relation.
First and Last element: Let X be a partially order set with the relation R. If a ∈ X such
that aRx for every x ∈ X, then we say that a is the first element of X. Similarly if b ∈ X
such that xRb for every x ∈ X, then b is called the last element of X.
Well ordered set: A partially ordered set X is called well ordered if every non-empty
subset of X contains the first element.
Example. N is well ordered set under the usual relation ≤.
Maximal and Minimal element: Let X be a partially ordered set with the ralation R.
An element a ∈ X is called a minimal element of X if no element of X related to a, that is,
if xRa implies x = a.
An element b ∈ X is called a maximal element of X if b is not related to any element of X,
that is, if bRx implies x = b.
Example: Consider the divisibility relation on the set S = {2, 3, 4, 6, 9, 10, 12, 36}. Then 2
and 3 are minimal elements and 10 and 36 are maximal elements.

1
Upper and lower bounds: Let A be a subset of a partially orederd set X. An element
a ∈ X is called a lower bound of A if aRx for every x ∈ A. Similarly an element b ∈ X is
an upper bound of A if xRb for every x ∈ A. A set A may have no upper bound or lower
bound.
Infimum and Supremum: Let A∗ denote the collection of all upper bounds of A and A∗
denote the collection of all lower bounds of A. Then the first element of A∗ , if it exists, is
called the least upper bound or the supremum of A. Similarly the last element of A∗ , if it
exists, is called the greatest lower bound or infimum of A.
Example: Let R with usual order relation ≤ and let A = {x ∈ R : 1 < x < 2}. Here
A∗ = {x ∈ R : x ≤ 1}. and A∗ = {x ∈ R : x ≥ 2}. So, the supremum is 2 and infimum is 1. yeh
galat
Order Completeness Axiom: A partial order set X is said to be order complete if every hai by
tanishq
non-empty subset of X which has an upper bound (or which has a lower bound) has a and
supremum (or infimum). verified
by
Example: The sets N and R with usual oreder ≤ are order complete. The set Q is not hemant
order complete. Consider A := {x ∈ Q : 2 < x2 < 5}, has no supremum and infimum. singh

Function: A function f from A to B is an assignment which assigns each element of A to


a unique element of B. Equivalentely, a function is a relation from A to B such that for
each element a ∈ A there is a unique element b ∈ B such that aRb. A and B are called the
domain and co-domain of the function respectively. The set of all those elements in B which
are mapped by some elemets in A is called the range or image of f .
Types of Functions: Let f : A → B.
1. Injective (one-to-one): If distinct elements of domain have distinct images. That
is, f is one-to-one if f (x) = f (y), then x = y. Or x ̸= y implies f (x) ̸= f (y). Find the
number of injective functions from one set to another.
2. Surjective (onto): If every element of co-domain are mapped by some elements of
domain. That is f is onto, if for each b ∈ B there is a ∈ A such that f (a) = b. Find
the number of surjective functions from one set to another.
3. Bijective: If it is one one and onto. Find the number of bijective functions from one
set to other.
Similar set: Sets A and B are called similar, if there is a bijective map between them.
Example: N and E (set of all even natural number) are similar. Define f : N → E
by f (n) = 2n.
Countable set: A set S is called countable if it is similar to N.
Example: The set S = { 21 , 23 , 34 , 45 , . . .} is countable. Define f : N → S as f (n) = n
(n+1)
for all n ∈ N .
Uncountable set: A set which is not countable is called uncountable set.
Example: The set of real numbers R is not countable, that is, an uncountable set.

2
Proof: Suppose R is countable. We know that a subset of a countable set is countable.
consider A = (0, 1) ⊆ R. We show that A is not countable. On the contrary, suppose A is
countable. Then we can write elements of A as r1 , r2 , r3 . . ., where ri can be written in the
decimal expansion form as follows:
r1 = d11 d12 d13 . . .
r2 = d21 d22 d23 . . .
r3 = d31 d32 d33 . . .
...........................
rn = dn1 dn2 dn3 . . .
...........................
, where dij ∈ {0, 1, 2, . . . , 9}. Now consider r = d1 d2 d3 . . . as follows:


1 ̸ 1
dii =
di =
2 dii = 1.

Then r is an element of A which is not equal to ri . Thus A is uncountable and therefore R


is uncountable.
Schröder-Bernstein Theorem: If A and B are sets with |A| ≤ |B| and |B| ≤ |A|, then
|A| = |B|. In other words, if there are one-to-one functions f from A to B and g from B to
A, then there is a one-to-one correspondence between A and B.
Example: Show that |(0, 1)| = |(0, 1]|
Solution: Let A = (0, 1) and B = (0, 1]. Then consider f : A → B defined as f (x) = x and
g : B → A defined as g(x) = x2 . Then f and g are one-to-one. Therefore |(0, 1)| = |(0, 1]|.

3
Lecture-(7-9)
(Proof Techniques)

Mathematical system: A system consists of Axioms, Definitions, and Terms is called a


Mathematical system. We prove or disprove any statement within a mathematical system. Let
us define some terms which are related to a mathematical system directly or indirectly.
1. Definition: A precise description of meaning of a mathematical term.
2. Theorem: A proposition that has been proved to be true. A theorem is of two kinds:
Lemma and Corollary.
3. Lemma: A theorem that is usually not too interesting in its own right but is useful in
proving another theorem.
4. Corollary: A theorem that follows immediately from another theorem.
5. Conjecture: A statement that is suspected to be true but yet to prove.
Example: The 4-color conjecture, the 3x + 1 conjecture, Goldbach’s conjecture, Hadwiger
conjecture, the abc conjecture, etc.
6. Axiom: A statement that is assumed to be true without proof.
Example: 2+2=4.
7. Paradox: A statement that can be shown, using a given set of axioms and definitions,
to be both true and false at the same time.
Example: Nobody goes to Murphy’s Bar anymore as it’s too crowded.

1 Methods of Proof:
By a proof, of a proposition p ⇒ q, we mean an argument that establishes the truth value of
the proposition. Since the argument can be given in different forms and hence we can have
different proof techniques.
1. Direct Method: Using p is true and with the help of other axioms, definitions and
previously derived theorems, we here show that q is true.
(a) Example: If m is odd and n is even integer, then show that m + n is odd integer.
Proof: We use the definitions of even and odd integer.
m is odd if there is an integer k1 such that m = 2k1 + 1 and n is even integer if there
is an integer k2 such that n = 2k2 .
Then m + n = 2k1 + 1 + 2k2 = 2(k1 + k2 ) + 1 = 2k + 1, where k = k1 + k2 . So, m + n
is odd.

1
2. Proof by Contradiction In this technique, we assume that q is false, that is, ¬q is true.
Note that ¬(p → q) ≡ (p ∧ ¬q), that is to say, p → q is true if and only if (p ∧ ¬q) is false.
In other words, p ∧ ¬q is a contradiction.
(a) Example: For any integer x if x2 is even, then x is even.
Proof: Suppose x is not even and x2 is even. So x = 2k1 + 1 and x2 = 2k2 for some
integers k1 , k2 . Then we have (2k1 + 1)2 = 2k2 . This implies 4(k12 + k1 ) + 1 = 2k2 .
But 4(k12 + k1 ) + 1 is odd and 2k2 is even, so these cannot be equal. Thus we have a
contradiction.

(b) Example: Prove that 2 is irrational.
√ √
Proof: Suppose 2 is rational. Then we can write pq = 2, where (p, q) = 1.
Then squaring both sides, we get p2 = 2q 2 . This implies p is even, that is, p = 2k
for some integer k. But then q 2 = 2k 2 , that is, q is even. This gives a contradiction
that (p, q) = 1.
(c) Example: Prove that primes are infinite.
Proof: Suppose there are only k primes p1 , p2 , . . . , pk . Now consider n = p1 p2 . . . pk +
1. Since n is not a prime so there is some prime pi such that pi divides n. Also pi
divides p1 p2 . . . pk . This implies pi divides n − p1 p2 . . . pk = 1. This is a contradiction
as the smallest prime is 2.
(d) Example: Prove that there are no integers x and y such that x2 = 4y + 2.
Proof: Suppose there are integers x and y such that x2 = 4y + 2 = 2(2y + 1). So
x2 is even and therefore x is even. Let x = 2k for some integer k. Then substituting
this, we get 2k 2 = 2y + 1. But 2k 2 is even while 2y + 1 is odd, so these cannot be
equal. Thus we have a contradiction.
3. Proof by Contrapositive: Note that p ⇒ q ≡ ¬(p ∧ ¬q) ≡ ¬(¬q ∧ p) ≡ ¬((¬q) ∧
¬(¬p)) ≡ (¬q ⇒ ¬p).
Thus p ⇒ q is logically equivalent to ¬q ⇒ ¬p. In other words, saying that if p is true
then q is true is equivalent to if q is false then p is false.
(a) Example: For any integer x if x2 is even, then x is even.
Proof: Suppose x is not even. So x = 2k1 + 1 for some integer k1 . Then we have
x2 = (2k1 + 1)2 = 4(k12 + k1 ) + 1. This shows that x2 is not even.
(b) Example: Let a and b be integers. If a + b is even, then a and b are either both
odd or both even.
Proof: Suppose that a and b are not both odd and both even. So one of a and b
is odd and other is even. Without loss of generality, assume that a is even and b is
odd. So a = 2k and b = 2l + 1 for some integers k, l. Therefore a + b = 2(k + l) + 1.
So a + b is odd.

2
4. Proof by Cases: If p ⇒ q and p is partitioned into cases r, s, that is, p ≡ r ∨ s. Then
from the below truth table, we see that p ⇒ q ≡ (r ∨ s) ⇒ q ≡ (r ⇒ q) ∧ (s ⇒ q).
r s q r∨s (r ∨ s) ⇒ q r⇒q s⇒q (r ⇒ q) ∧ (s ⇒ q)
T T T T T T T T
T T F T F F F F
T F T F T T T T
T F F T F F T F
F T T T T T T T
F T F T F T F F
F F T F T T T T
F F F F T T T T
So if p as a proposition involves “or”, it is sufficient to consider each of the possibilities
for p separately.
(a) Example: Prove that there is no possible integer n such that n2 + n3 = 100.
Proof (Method 1): If n2 + n3 = 100 then we have
n2 ≤ 100 and n3 ≤ 100. This implies n ≤ 10 and n ≤ 4. So we have to check for the
cases n = 1, 2, 3, 4. This gives the following cases:
For n = 1, n2 + n3 = 1 + 1 = 2 ̸= 100,
For n = 2, n2 + n3 = 4 + 8 = 12 ̸= 100,
For n = 3, n2 + n3 = 9 + 27 = 36 ̸= 100,
For n = 4, n2 + n3 = 16 + 64 = 80 ̸= 100.
Proof (Method 2): n2 + n3 = 100 is equivalent to n2 (1 + n) = 100. This is an
expression of factors of 100 into two numbers n2 and 1 + n.
Note that possible divisors of 100 are : 2,4,5,10,25,50 and out of then for the possi-
bility of n2 = 4 and n2 = 25.
Thus for n2 = 4, n = 2 and (1 + n) = 3, then we get n2 .(1 + n) = 4.3 = 12 ̸= 100,
Similarly, for n2 = 25, n = 5 and (1 + n) = 6, then we get n2 .(1 + n) = 25.6 = 150 ̸=
100.
5. Proof by Counterexample: Suppose we have problem: Prove or disprove A ⇒ B.
Thus if the proposition A ⇒ B is not true then to show that ¬(A ⇒ B) is true for some
instances.
If the problem is of the form ∀ x, A(x) ⇒ B(x), then its negation is ∃ x A(x) ̸⇒ B(x).
Recall that A ⇒ B ≡ B ∨ ¬A. So
∃x A(x) ̸⇒ B(x)
≡ ∃ x¬(A(x) ⇒ B(x))
≡ ∃ x¬(B(x) ∨ ¬A(x))

3
≡ ∃ x (¬B(x) ∧ A(x)).
Thus to prove the original statement is not true, we have to find an x such that (¬B(x) ∧
A(x)) is true.
(a) Example: Prove or disprove: for all positive integetr n, n2 − n + 41 is prime.
Solution: Let us disprove by counterexample. If the statement is not true then we
have to find a positive integer n such that n2 − n + 41 is not a prime.
Let n = 41. Then n2 − n + 41 is equal to 1681, which is not a prime.
(b) Example: Prove or disprove: for all positive inetegrs n, 2n + 1 is a prime.
Solution: For n = 1, 2n + 1 = 3, which is prime.
For n = 2, 2n + 1 = 5, which is prime.
For n = 3, 2n + 1 = 9, which is not a prime.
6. Existence Proofs: An existence proof is a proof of a statement of the form ∃ x P (x).
Such proofs are generally fall into one of the following two types:
(a) Constructive Proof: Establish P (x0 ) for some x0 in the domain of P .
i. Example: Prove that If f (x) = x3 + x − 5, then there exists a positive real

number x0 such that f (x0 ) = 7.
′ √
Proof: Find f (x) = 7, this gives x0 = 2.
(b) Nonconstructive Proof: Assume no x0 exists that makes P (x0 ) true and derive
a contradiction. In other words, use a proof by contradiction.
i. Example: Pigeonhole Principle: If n+1 pigeons are distributed into n holes,
then some hole must contain at least 2 of the pigeons.
Proof: Assume n + 1 pigeons are distributed into n boxes. Suppose the boxes
are labeled B1 , B2 , . . . , Bn , and assume that no box contains more than 1 object.
Let ki denote the number of objects placed in Bi . Then ki ≤ 1 for i = 1, . . . , n,
and so k1 + k2 + . . . + kn ≤ 1 + 1 + . . . + 1 ≤ n. But this contradicts the fact
that k1 + k2 + . . . + kn = n + 1, the total number of objects we started with.
7. Proof by Induction: There are two form of mathematical induction. One is weak form
and another is strong form. We discuss them separately.
(a) Weak Form of Mathematical Induction: Let P (n) be a statement on positive
integer n such that
1: P (1) is true,
2: for all k ≥ 1, P (k + 1) is true whenever one assumes that P (k) is true.
Then P (n) is true for all positive integer n.
n(n+1)
i. Example: Prove that 1 + 2 + . . . + n = 2
.

4
Proof: Let P (n) = 1 + 2 + . . . + n. Then P (n) holds for n = 1.
k(k+1)
Suppose P (n) holds for n = k, that is, P (k) = 1 + 2 + . . . + k = 2
. Now we
show that P (n) is true for n = k + 1.
k(k+1) (k+1)(k+2)
P (k + 1) = 1 + 2 + . . . + k + (k + 1) = 2
+ (k + 1) = 2
. Thus P (n)
holds for every n.
n(n+1)(2n+1)
ii. Exercise: Prove that 12 + 22 + . . . + n2 = 6
.
iii. Exercise: Prove that for any positive integer n, 1 + 3 + . . . + (n − 1) = n2 .
iv. Exercise: Let n ∈ N and suppose we are given real numbers a1 ≥ a2 ≥ . . . ≥
1
an ≥ 0. Then Arithmetic mean (AM) = a1 +a22+...an ≥ (a1 a2 . . . an ) n = GM
(Geometric mean).
v. Exercise: Fix a positive integer n and let A be a set with |A| = n. Let P (A)
denote the power set of A. Then show that |P (A)| = 2n .
Corollary of weak form of mathematical induction: Let P (n) be a statement
on positive integer n such that for some fixed positive integer n0
1: P (n0 ) is true,
2: for all k ≥ n0 , P (k + 1) is true whenever one assume that P (k) is true.
Then P (n) is true for all positive integer n ≥ n0 .
(b) Strong Form of the Principle of Mathematical Induction: Let P (n) be a
statement on positive integer n such that
1: P (1) is true,
2: P (k + 1) is true whenever one assumes that P (m) is true, for all m, 1 ≤ m ≤ k.
Then P (n) is true for all positive integer n.
Corollary of strong form of mathematical induction: Let P (n) be a statement
on positive integer n such that for some fixed positive integer n0 ,
1: P (n0 ) is true,
2: P (k + 1) is true whenever one assume that P (m) is true, for all m, n0 ≤ m ≤ k.
Then P (n) is true for all positive integer n ≥ n0 .

5
Lecture-(10-12)
(Counting Techniques)

How do you count the number of people in a crowded room? You could count heads, since
for each person there is exactly one head. Alternatively, you could count ears and divide by
two. Of course, you might have to adjust the calculation if someone lost an ear in a pirate raid
or someone was born with three ears. The point here is that you can often count one thing
by counting another, though some fudge factors may be required. This is a central theme of
counting, from the easiest problems to the hardest.
Let us note that every counting problem comes down to determining the size of some
set.
We first present basic counting rules. Then we will show how they can be used to solve many
different counting problems.
The product rule: Suppose a task has n ∈ N compulsory parts and the i-th part can be
completed in mi ∈ N ways for i = 1, 2, . . . , n. Then the task can be completed in m1 m2 . . . mn
ways.
In terms of sets, if A1 , A2 , . . . , An are sets, then

|A1 × A2 × · · · × An | = |A1 |.|A2 |. . . . .|An |.

1. How many three digit natural numbers can be formed using digits 0, 1, . . . , 9?
Solution: 9 × 10 × 10 ways.
2. The chairs of an auditorium are to be labeled with an uppercase English letter followed
by a positive integer not exceeding 100. What is the largest number of chairs that can be
labeled differently?
Solution: 100 × 26 ways.
3. There are 32 computers in a computer center. Each microcomputer has 24 ports. How
many different ports to a computer in the center are there?
Solution: 32 × 24 ports.
4. How many functions are there from a set with m elements to a set with n elements?
Solution: n × n × . . . × n(m times), that is, nm .
5. How many one-to-one functions are there from a set with m elements to one with n
elements?
Solution: n × (n − 1) × (n − 2) × . . . × (n − m + 1).
6. Let |S| = n. |P (S)| = 2n .
Solution: consider the one-to-one correspondence between subsets of S and bit strings
(each element takes on a value of 0 or 1) of length |S|. A susbset of S is associated with

1
the bit string with a 1 in the ith position if the ith element in the list is in the subset. By
the multiplication rule, there are 2|S| bit strings of length |S|. Therefore, |P (S)| = 2|S| .
The sum rule: Suppose a task consists of n alternative parts (either parts), and the i-th part
can be completed in mi ways, i = 1, . . . , n. Then the task can be completed in m1 +m2 +. . .+mn
ways. Following examples illustrate the rule.
In terms of sets, if A1 , A2 , . . . , An are disjoint sets, then

|A1 ∪ A2 ∪ · · · ∪ An | = |A1 | + |A2 | + · · · + |An |.

1. Suppose that either a member of the mathematics faculty or a student who is a mathe-
matics major is chosen as a representative to a university committee. How many different
choices are there for this representative if there are 37 members of the mathematics faculty
and 83 mathematics majors and no one is both a faculty member and a student?
Solution: 37+83= 110.
2. How many three digit natural numbers with distinct digits can be formed using digits
1, . . . , 9 such that each digit is odd or each digit is even?
Solution: The task has two alternative parts. Part 1: form a three digit number with
distinct digits using digits from {1, 3, 5, 7, 9}. Part 2: form a three digit number with
distinct digits using digits from {2, 4, 6, 8}. Observe that Part 1 is a task having three
compulsory subparts. Using multiplication rule, we see that Part 1 can be done in 5×4×3
ways. Part 2 is a task having three compulsory subparts. So, it can be done in 4 × 3 × 2
ways. Since our task has alternative parts, addition rule implies 60 + 24 = 84.
The subtraction rule: If a task can be done in either n1 ways or n2 ways, then the number
of ways to do the task is n1 + n2 minus the number of ways to do the task that are common to
the two different ways.
In terms of sets, if A and B are finite sets, then

|A ∪ B| = |A| + |B| − |A ∩ B|.

The subtraction rule is also known as the principle of inclusion-exclusion, especially when it is
used to count the number of elements in the union of two sets.
1. How many bit strings of length eight either start with a 1 bit or end with the two bits
00?
Solution: 27 + 26 − 25 .
2. A computer company receives 350 applications from computer graduates for a job. Sup-
pose that 220 of these applicants majored in computer science, 147 majored in business,
and 51 majored both in computer science and in business. How many of these applicants
majored neither in computer science nor in business?
Solution: Let A1 be the set of students who majored in computer science and A2 the
set of students who majored in business. By the subtraction rule, the number of students

2
who majored either in computer science or in business equals A ∪ B
|A1 ∪ A2 | = |A1 | + |A2 | − |A1 ∩ A2 | = 220 + 147 − 51 = 316. Thus, 350 − 316 = 34 of the
applicants majored neither in computer science nor in business.
The division rule: There are n/d ways to do a task if it can be done using a procedure that
can be carried out in n ways, and for every way w, exactly d of the n ways correspond to way
w.
In terms of functions, if f : A → B is k−to−1, then |A| = k|B|. A k−to−1 function maps
exactly k elements of the domain to every element of the codomain.
1. How many different ways are there to seat four people around a circular table, where two
seatings are considered the same when each person has the same left neighbor and the
same right neighbor?
Solution: There are 4! seatings. Now any seating is similar to four seatings. For example
A − B − C − D is similar to A − B − C − D, B − C − D − A, C − D − A − B, D − A − B − C.
Thus total number of different seatings is 4!/4 = 6.
Bijection Rule: If there is a bijection f : A → B between A and B, then |A| = |B|.

The Subset Rule: The number of k-element subsets of an n-element set is


 
n n!
= .
k k!(n − k)!

Let A be an n-element set and k1 , k2 , . . . , km be nonnegative integers whose sum is n. A


(k1 , k2 , . . . , km )-split of A is a sequence

(A1 , A2 , . . . , Am ),

where the Ai are disjoint subsets of A and |Ai | = ki for i = 1, 2 . . . , m.

Subset Split Rule: The number of (k1 , k2 , . . . , km )-splits of an n element set is


 
n n!
= .
k1 , k2 , . . . , km k1 ! k2 ! . . . km !

The Bookkeeper Rule: Let l1 , . . . , lm be distinct elements. The number of sequences with
k1 occurrences of l1 , and k2 occurrences of l2 , . . . , and km occurrences of lm is

(k1 + k2 + · · · + km )!
.
k1 ! k2 ! . . . km !
Example: Suppose you are planning a 20-mile walk, which should include 5 northward miles,
5 eastward miles, 5 southward miles, and 5 westward miles. How many different walks are
possible?
Solution: There is a bijection between such walks and sequences with 5 N’s, 5 E’s, 5 S’s, and

3
5 W’s. By the Bookkeeper Rule, the number of such sequences is:

(5 + 5 + 5 + 5)! 20!
= .
5! 5! 5! 5! (5!)4

Pigeonhole Principle: If n + 1 pigeons (resp. objects) are distributed into n holes (resp.
boxes), then some hole (box) must contain at least 2 of the pigeons (objects).
Proof: Assume n + 1 pigeons are distributed into n boxes. Suppose the boxes are labeled
B1 , B2 , . . . , Bn , and assume that no box contains more than 1 object. Let ki denote the number
of objects placed in Bi . Then ki ≤ 1 for i = 1, . . . , n, and so k1 +k2 +. . .+kn ≤ 1+1+. . .+1 ≤ n.
But this contradicts the fact that k1 + k2 + . . . + kn = n + 1, the total number of objects we
started with.
1. Among any group of 367 people, there must be at least two with the same birthday,
because there are only 366 possible birthdays.
2. In any group of 27 English words, there must be at least two that begin with the same
letter, because there are 26 letters in the English alphabet.
3. How many students must be in a class to guarantee that at least two students receive the
same score on the final exam, if the exam is graded on a scale from 0 to 100 points?
Solution: There are 101 possible scores on the final. The pigeonhole principle shows
that among any 102 students there must be at least 2 students with the same score.
The Generalized Pigeonhole Principle: If N objects are placed into k boxes, then there
is at least one box containing at least ⌈N/k⌉ objects, where ⌈.⌉ denotes the ceiling function.
In terms of functions, If |X| > k|Y |, then every function f : X → Y maps at least k +1 different
elements of X to the same element of Y.

1. Among 100 people there are at least ⌈100/12⌉ = 9 who were born in the same month.
2. Show that among any n + 1 positive integers not exceeding 2n there must be an integer
that divides one of the other integers.
Solution: Let us write each of the n + 1 integers a1 , a2 , . . . , an+1 as a power of 2 times
an odd integer. In other words, let aj = 2kj qj for j = 1, 2, . . . , n + 1, where kj is a non-
negative integer and qj is odd. The integers q1 , q2 , . . . , qn+1 are all odd positive integers
less than 2n. Because there are only n odd positive integers less than 2n, it follows from
the pigeonhole principle that two of the integers q1 , q2 , . . . , qn+1 must be equal. Therefore,
there are distinct integers i and j such that qi = qj . Let q be the common value of qi and
qj . Then, ai = 2ki q and aj = 2kj q. It follows that if ki < kj , then ai divides aj , while if
ki > kj , then aj divides ai .

A permutation of a set of distinct objects is an ordered arrangement of these objects. An ordered


arrangement of r elements of a set is called an r-permutation. The number of r-permutations
n!
of a set with n elements is denoted by P (n, r) = n(n − 1)(n − 2) . . . (n − r + 1) = (n−r)! .

4
1. Let S = a, b, c. The 2-permutations of S are the ordered arrangements a, b; a, c; b, a; b, c;
c, a; c, b. Consequently, there are six 2-permutations of this set with three elements. We
see that P (3, 2) = 3.2 = 6.
2. How many permutations of the letters ABCDEFGH contain the string ABC ?
Solution: Because the letters ABC must occur as a block, we can find the answer
by finding the number of permutations of six objects, namely, the block ABC and the
individual letters D, E, F , G, and H . Because these six objects can occur in any order,
there are 6! = 720 permutations of the letters ABCDEFGH in which ABC occurs as a
block
An r-combination of elements of a set is an unordered selection of r elements from the set. The
number of r-combinations of a set with n distinct elements is denoted by C(n, r). Note that
n
C(n, r) = r = r!(n−r)! , Here nr is called a binomial coefficient.
n!


Example: How many poker hands of five cards can be dealt from a standard deck of 52 cards?
Solution: Because the order in which the five cards are dealt from a deck of 52 cards does not
matter, there are
52!
C(52, 5) = .
5! 47!

Corollary: Let n and r be nonnegative integers with r ≤ n. Then C(n, r) = C(n, n − r).
The Binomial Theorem: Let x and y be variables, and let n be a nonnegative integer. Then
n  
n
X n
(x + y) = xn−j y j .
j=0
j

Corollary: Let n be a nonnegative integer. Then


n   n  
X n n
X
k n
= 2 and (−1) = 0.
k=0
k k=0
k

Pascal’s Identity: Let n and k be positive integers with n ≥ k. Then


     
n+1 n n
= + .
k k−1 k

Vandermonde’s Identity: Let m, n, and r be nonnegative integers with r not exceeding


either m or n. Then   X r   
m+n m n
= .
r k=0
r − k k

2n Pn n 2
 
Corollary: If n is a nonnegative integer, then n
= k=0 k .

5
Lecture-(13-15)

Theorem (Permutations with Repetition): The number of r−permutations of a set of


n objects with repetition allowed is nr .
Proof: There are n ways to select an element of the set for each of the r positions in the
r-permutation when repetition is allowed, because for each choice all n objects are available.
Hence, by the product rule there are nr r-permutations when repetition is allowed.
Example: How many strings of length r can be formed from the uppercase letters of the
English alphabet?
Solution: By the product rule( or by the above theorem), because there are 26 uppercase
English letters, and because each letter can be used repeatedly, we see that there are 26r
strings of uppercase English letters of length r.
Theorem (Combinations with Repetition): There are C(n+r−1, r) = C(n+r−1, n−1)
r-combinations from a set with n elements when repetition of elements is allowed.
Proof: Each r-combination of a set with n elements when repetition is allowed can be
represented by a list of n − 1 bars and r stars. The n − 1 bars are used to mark off n different
cells, with the ith cell containing a star for each time the ith element of the set occurs in the
combination.
For instance, a 6-combination of a set with four elements is represented with three bars and
six stars. Here
∗ ∗ | ∗ || ∗ ∗∗
represents the combination containing exactly two of the first element, one of the second
element, none of the third element, and three of the fourth element of the set.
As we have seen, each different list containing n − 1 bars and r stars corresponds to an
r-combination of the set with n elements, when repetition is allowed. The number of such
lists is C(n − 1 + r, r), because each list corresponds to a choice of the r positions to place
the r stars from the n − 1 + r positions that contain r stars and n − 1 bars. The number of
such lists is also equal to C(n − 1 + r, n − 1), because each list corresponds to a choice of the
n − 1 positions to place the n − 1 bars.
Example: How many solutions does the equation

x1 + x2 + x3 = 11

have, where x1 , x2 , and x3 are nonnegative integers?


Solution: To count the number of solutions, we note that a solution corresponds to a way
of selecting 11 items from a set with three elements so that x1 items of type one, x2 items
of type two, and x3 items of type three are chosen. Hence, the number of solutions is equal
to the number of 11-combinations with repetition allowed from a set with three elements.
From the above theorem, it follows that there are

C(3 + 11 − 1, 11) = C(13, 11) = C(13, 2) = 78

1
solutions.
Generating Functions: The generating function for the sequence a0 , a1 , . . . , ak , . . . of real
numbers is the infinite series

X
k
G(x) = a0 + a1 x + · · · + ak x + · · · = ak x k .
k=0

Example: What is the generating function for the sequence 1, 1, 1, 1, 1, 1?


Solution: The generating function of 1, 1, 1, 1, 1, 1 is 1 + x + x2 + x3 + x4 + x5 .
Example: The function f (x) = 1/(1 − x) is the generating function of the sequence
1, 1, 1, 1, . . . .

P∞ P∞
Theorem: Let f (x) = k=0 ak xk and g(x) = k=0 bk x
k
. Then


X ∞ X
X k 
f (x) + g(x) = (ak + bk )xk and f (x)g(x) = aj bk−j xk .
k=0 k=0 j=0

Extended Binomial Coefficient: Let u be a real number and k a nonnegative integer.


Then the extended binomial coefficient uk is defined by
 
u
= u(u − 1) . . . (u − k + 1)/k! if k > 0,
k

and uk = 1 if k = 0.


The Extended Binomial Theorem: Let x be a real number with |x| < 1 and let u be a
real number. Then ∞  
u
X u k
(1 + x) = x .
k=0
k
Example(Counting Problems and Generating Functions): Find the number of so-
lutions of x1 + x2 + x3 = 17, where x1 , x2 , and x3 are nonnegative integers with 2 ≤ x1 ≤
5, 3 ≤ x2 ≤ 6, and 4 ≤ x3 ≤ 7.
Solution: The number of solutions with the indicated constraints is the coefficient of x1 7
in the expansion of (x2 + x3 + x4 + x5 )(x3 + x4 + x5 + x6 )(x4 + x5 + x6 + x7 ).
Example (Using Generating Functions to Solve Recurrence Relations): Solve the
recurrence relation ak = 3ak−1 for k = 1, 2, 3, . . . and initial condition a0 = 2.
Solution: Let G(x) be the generating function for the sequence {ak }, that is, G(x) =
P∞ k
k=0 ak x Now, let us observe that


X ∞
X
k+1
xG(x) = ak x = ak−1 xk .
k=0 k=1

With the help of recurrence relation, we have

2
G(x)−3xG(x) = ∞
P k
P∞ k
P∞ k
k=0 a k x −3 k=1 a k−1 x = a 0 + k=1 (ak −3ak−1 )x = 2, because a0 = 2
and ak = 3ak−1 . Further, we have that G(x) = 2/(1 − 3x). Using the identity

X
1/(1 − ax) = ak x k ,
k=0

we obtain ∞ ∞
X X
k k
G(x) = 2 3 x = 2.3k xk ,
k=0 k=0

hence, ak = 2.3k .
Example(Proving Identities via Generating Functions): Use generating functions to
show that n
X
C(n, k)2 = C(2n, n),
k=0

whenever n is a positive integer.


Solution: Using the Binomial theorem, we have that C(2n, n) is the coefficient of xn in
(1 + x)2n . However, we also have

(1 + x)2n = [(1 + x)n ]2 = [C(n, 0) + C(n, 1)x + C(n, 2)x2 + · · · + C(n, n)xn ]2 .

The coefficient of xn in this expression is


n
X
C(n, 0)C(n, n)+C(n, 1)C(n, n−1)+C(n, 2)C(n, n−2)+· · ·+C(n, n)C(n, 0) = C(n, k)2 ,
k=0
Pn
because C(n, n − k) = C(n, k). Because both C(2n, n) and k=0 C(n, k)2 represent the
coefficient of xn in (1 + x)2n , they must be equal.
Exercise: Prove Pascal’s identity and Vandermonde’s identity using generating functions.

Recurrence Relations: A linear homogeneous recurrence relation of degree k with constant


coefficients is a recurrence relation of the form

xn = c1 xn−1 + c2 xn−2 + · · · + ck xn−k ,

where c1 , c2 , . . . , ck are real numbers, and ck ̸= 0.

Theorem: Let c1 and c2 be real numbers. Suppose that r2 − c1 r − c2 = 0 has two


distinct roots r1 and r2 . Then the sequence {xn } is a solution of the recurrence relation
xn = c1 xn−1 + c2 xn−2 if and only if xn = α1 r1n + α2 r2n for n = 0, 1, 2, . . . , where α1 and α2
are constants.

Proof: Here we will do two things to prove the theorem. First, we will show that if r1 and r2
are the roots of the characteristic equation, and α1 and α2 are constants, then the sequence

3
{xn } with xn = α1 r1n + α2 r2n is a solution of the recurrence relation. Second, we will show
that if the sequence {xn } is a solution, then xn = α1 r1n + α2 r2n for some constants α1 and α2 .
Now we will show that if xn = α1 r1n + α2 r2n , then the sequence {xn } is a solution of the
recurrence relation. Because r1 and r2 are roots of r2 − c1 r − c2 = 0, it follows that
r12 = c1 r1 + c2 , r22 = c1 r2 + c2 . From these equations, we see that

c1 xn−1 + c2 xn−2 = c1 α1 r1n−1 + α2 r2n−1 + c2 α1 r1n−2 + α2 r2n−2


 

= α1 r1n−2 (c1 r1 + c2 ) + α2 r2n−2 (c1 r2 + c2 )


= α1 r1n−2 r12 + α2 r2n−2 r22
= α1 r1n + α2 r2n
= xn
This shows that the sequence {xn } with xn = α1 r1n + α2 r2n is a solution of the recurrence
relation.
To show that every solution {xn } of the recurrence relation xn = c1 xn−1 + c2 xn−2 has xn =
α1 r1n + α2 r2n for n = 0, 1, 2, . . ., for some constants α1 and α2 , suppose that {xn } is a solution
of the recurrence relation, and the initial conditions x0 = C0 and x1 = C1 hold. It will be
shown that there are constants α1 and α2 such that the sequence {xn } with xn = α1 r1n +α2 r2n
satisfies these same initial conditions. This requires that

x0 = C0 = α1 + α2 ,
x1 = C1 = α1 r1 + α2 r2 .

We can solve these two equations for α1 and α2 . From the first equation it follows that
α2 = C0 − α1 . Inserting this expression into the second equation gives

C1 = α1 r1 + (C0 − α1 ) r2 .

Hence,
C1 = α1 (r1 − r2 ) + C0 r2 .
This shows that
C1 − C0 r2
α1 =
r1 − r2
and
C 1 − C 0 r2 C0 r1 − C1
α2 = C0 − α1 = C0 − = ,
r1 − r2 r1 − r2
where these expressions for α1 and α2 depend on the fact that r1 ̸= r2 . (When r1 = r2 ,
this theorem is not true.) Hence, with these values for α1 and α2 , the sequence {xn } with
α1 r1n + α2 r2n satisfies the two initial conditions.
We know that {xn } and {α1 r1n + α2 r2n } are both solutions of the recurrence relation xn =
c1 xn−1 + c2 xn−2 and both satisfy the initial conditions when n = 0 and n = 1. Because
there is a unique solution of a linear homogeneous recurrence relation of degree two with two
initial conditions, it follows that the two solutions are the same, that is, xn = α1 r1n + α2 r2n
for all nonnegative integers n. We have completed the proof by showing that a solution of

4
the linear homogeneous recurrence relation with constant coefficients of degree two must be
of the form xn = α1 r1n + α2 r2n , where α1 and α2 are constants.

Example: Find the solution to the recurrence relation

xn = 6xn−1 − 11xn−2 + 6xn−3

with the initial conditions x0 = 2, x1 = 5, and x2 = 15.


Solution:
(1) The characteristic polynomial of this recurrence relation is r3 − 6r2 + 11r − 6.
(2) The characteristic roots are r = 1, r = 2, and r = 3.
(3) The solutions to this recurrence relation are of the form

xn = α1 1n + α2 2n + α3 3n .

(4) Using intial conditions, we have α1 = 1, α2 = −1, and α3 = 2.


(5) After putting values of αi ’s in Step-3, we have xn = 1 − 2n + 2.3n . This completes the
task.
Theorem: Let c1 , c2 , . . . , ck be real numbers. Suppose that the characteristic equation

rk − c1 rk−1 − · · · − ck = 0

has t distinct roots r1 , r2 , . . . , rt with multiplicities m1 , m2 , . . . , mt , respectively, so that mi ≥


1 for i = 1, 2, . . . , t and m1 + m2 + · · · + mt = k. Then a sequence {xn } is a solution of the
recurrence relation
xn = c1 xn−1 + c2 xn−2 + · · · + ck xn−k
if and only if
xn =(α1,0 + α1,1 n + · · · + α1,m1 −1 nm1 −1 )r1n
+ (α2,0 + α2,1 n + · · · + α1,m2 −1 nm2 −1 )r2n +
· · · + (αt,0 + αt,1 n + · · · + αt,mt −1 nmt −1 )rtn
for n = 0, 1, 2, . . . , where αi,j are constants for 1 ≤ i ≤ t and 0 ≤ j ≤ mi − 1.

Example: Find the solution to the recurrence relation xn = −3xn−1 − 3xn−2 − xn−3 with
initial conditions x0 = 1, x1 = −2, and x2 = −1.
Solution:
(1) The characteristic polynomial of this recurrence relation is r3 +3r2 +3r+1 = (r+1)3 = 0
(2) The characteristic roots are r = −1, −1, −1. Here multiplicity of −1 is three.
(3) The solutions to this recurrence relation are of the form

5
(4) Using intial conditions, we have α1,0 = 1, α1,1 = 3, and α1,2 = −2.
(5) After putting values of the constants in Step-3, we have xn = (1 + 3n − 2n2 )(−1)n .
This completes the required job.

6
Lecture-16
(Graph Theory)

1 Basic Terminologies
Definition 1.1 A graph G consists of two sets V and E, where V is non-empty set, called the
vertex set, and E is called the edge set. A graph G is also denoted as G = (V, E). We define
some terminologies in a graph G as follows.
1. Let u, v ∈ V . An edge e ∈ E joining them is denoted as e = uv. In this cases, u and v
are called adjacent vertices, also called the end vertices, of the edge e. We also say that
e is incident at u and v. Two edges e1 , e2 ∈ E are called adjacent if they have a common
end vertex.
2. Let v ∈ V . The neighborhood of v, denoted as N (v), is the set of all the adjacent vertices
to v. Similarly, if A ⊆ V , then N (A) is the set of all the vertices which are adjacent to
at least one vertex of A, that is, N (A) = ∪v∈A N (v).
3. The degree of a vertex v ∈ V , denoted as deg(v), is the number of edges incident at v.
A vertex v is called isolated vertex if deg(v) = 0 and pendent vertex if deg(v) = 1. The
minimum degree of a vertex in G is denoted by δ(G) and the maximum degree of a vertex
in G is denoted by ∆(G).
4. A set of vertices or edges is said to be independent if no two of them are adjacent. The
maximum size of an independent vertex set is called the independence number of G,
denoted α(G).
5. If the end vertices of an edge e ∈ E are same then the edge is called loop. If e1 , e2 are
two edges such that they have same end points, then the edges are called parallel edges
or multiple edges. A graph is called simple if it has no loops or multiple edges.
In these notes, unless stated otherwise, all our graphs are simple graphs with finite number of
vertices (and hence finite number of edges).
P
Lemma 1.0.1 [Handshake Lemma:] Let G = (V, E) be a graph. Then v∈V deg(v) = 2|E|,
where |E| denote the number of edges in E.
Proof: The proof is based on induction on the cardinality of edge set, that is, |E|. Clearly, the
result holds for |E| = 1.
Suppose the result holds for any graph G with |E| = k.
Let G be a graph with |E| = k + 1. Then consider a graph G′ = (V, E ′ ), where E ′ = E \ uv.
Then the number of edges in G′ is k and therefore the result holds for G′ , that is

P
v∈V deg(v) = 2|E |.

the sum of degree counts the total number of times an edge is incident on a vertex. An edge is
incident on exactly 2 vertex . 1
Now, add the removed edge back to G′ . Because this edge is indecent on two vertices, we add
two to the previous sum, that is, v∈V deg(v) + 2 = 2|E ′ | + 2. Thus v∈V deg(v) = 2|E|.
P P

Corollary 1 Let G = (V, E). Then the number of odd degree vertices is even.
Proposition 1 In a graph G = (V, E) with |V | = n ≥ 2, there are two vertices of equal degree.
Proof: If G has two or more isolated vertices, then we are done. Suppose G has one isolated
vertex. Then the remaining n − 1 vertices have degree between 1 to (n − 2) and hence by PHP
the result holds. Otherwise, G has no isolated vertices. Then there are n vertices whose degrees
lies between 1 to n − 1. Again by PHP, the result holds.

2 Some Special Simple Graphs n(n-1)/2=e

1. Complete graph: A graph G = (V, E) with |V | = n is called a complete graph if each


pair of vertices form an edge. We denote the G by Kn .
2. Cycle: A cycle Cn , (n ≥ 3), consists of n vertices v1 , v2 , . . . , vn and the edges v1 v2 ,
v2 v3 , . . . , vn−1 vn , v1 . e=n

3. Wheel Graph: A wheel graph Wn is obtained from cycle Cn , (n ≥ 3), when each vi of
Cn is adjacent to another vertex v. an additional vertex v is added and each previous vertex is
connected to v . |V|=n+1 and |E|=2n
4. Bipartite Graph: A graph G = (V, E) is called bipartite graph if the vertex set V can
be partitioned into two disjoint sets V1 and V2 such that each edge e ∈ E has one end
vertex in V1 and other in V2 . If |V1 | = m and |V2 | = n, then we denote the graph G by
Km,n .
5. Complete Bipartite Graph: A bipartite graph G = (V, E), with partition sets V1 and
V2 of V , is called complete bipartite graph if each pair {u, v}, where v ∈ V1 and u ∈ V2
forms and edge.
e=mn
Theorem 2.1 A simple graph is bipartite if and only if it is possible to assign one of two
different colors to each vertex of the graph so that no two adjacent vertices are assigned the
same color.
Proof: First assume that G = (V, E) is bipartite graph with bipartite subsets V1 and V2 of V .
Then assign one color to each vertex of V1 and a second color to each vertex of V2 will give the
desired condition.
Conversely, let it is possible to assign one of two different colors to each vertex of the graph so
that no two adjacent vertices are assigned the same color. Let V1 be the set of vertices assigned
one color and V2 be the set of vertices assigned the other color. Then, V1 and V2 are disjoint
and V = V1 ∪ V2 . 2

2
a single vertex of G is a subgraph of G
a single edge along with its end vertices in G is a subgraph of G
Lecture-(17-18)

every simple graph of n vertices is a subgraph of complete graph Kn

1 New Graphs from Old Graph condition- each edge in G' must have the same end
points as in G
1. Subgraph: A subgraph G′ of a graph G = (V, E) is a graph G′ = (V ′ , E ′ ) such that
V ′ ⊆ V and E ′ ⊆ E.
2. Spanning subgraph: A subgraph G′ = (V ′ , E ′ ) of G = (V, E) is called spanning sub-
graph of G if V = V ′ .
3. Induced subgraph: A subgraph G′ = (V ′ , E ′ ) of G = (V, E) is called an induced
subgraph of G if for every u, v ∈ V ′ , e = uv ∈ E ′ whenever e = uv ∈ E.
4. If v ∈ V , then the graph G − v, called the vertex deleted subgraph, is obtained from
G by deleting v and all the edges that are incident with v.
5. If e ∈ E, then the graph G − e = (V, E \ {e}) is called the edge deleted subgraph.
6. If u, v ∈ V , then G + uv = (V, E ∪ {uv}) is called the graph obtained by edge addition.
7. The complement G of a graph G is defined as (V , E), where V = V and E = {uv | u ̸=
v, uv ̸∈ E}.
Definition 1.1 Let G = (V (G), E(G)) and H = (V (H), E(H)) be two graphs.
1. Then their intersection, denoted G ∩ H, is defined as (V (G) ∩ V (H), E(G) ∩ E(H)).
2. Then their union, denoted G ∪ H, is defined as (V (G) ∪ V (H), E(G) ∪ E(H)).

2 Representing Graph and Graph Isomorphism


Definition 2.1 Let G = (V, E) be a graph with V = {v1 , v2 , . . . , vn }. Then the adjacency
matrix with respect to the ordering v1 , v2 , . . . , vn of V is the matrix AG = [aij ], where


1 if vi vj is an edge of G
aij =
0 otherwise.
Definition 2.2 Let G = (V, E) be a graph. Suppose that v1 , v2 , . . . , vn are the vertices and
e1 , e2 , . . . , em are the edges of G. Then the incidence matrix with respect to this ordering of V
and E is the n × m matrix M = [mij ], where


1 when edge ej is incident with vi ,
mij =
0 otherwise.
Definition 2.3 Let G1 = (V1 , E1 ) and G2 = (V2 , E2 ) be two graphs. Then G1 is said to
be isomorphic to G2 , denoted as G1 ∼ = G2 , if there is bijective map f : V1 → V2 such that
{f (u), f (v)} ∈ E2 if and only if {u, v} ∈ E1 .

1
Definition 2.4 A graph G is called self-complementary if G ∼
= G.
Definition 2.5 A graph invariant is a function which assigns the same value to isomorphic
graphs. Observe that some of the graph invariants are: |V |, |E|, ∆(G) (maximum degree),
δ(G) (minimum degree), ω(G) (clique number), r(G) (radius), e(G) (eccentricity).
Proposition 1 Let G and H be graphs and let f : G → H be an isomorphism. For any
v ∈ V (G) , G − v ∼
= H − f (v).
Proof: Consider the bijection g : V (G − v) → V (H − f (v)) described by g = fV (G − v).

3 Connectedness
Definition 3.1 Let G = (V, E) be a graph.
1. A u-v walk W is a finite sequence of vertices (u = v1 , v2 , . . . , vn = v) such that vi vi+1 ∈ E
for all i = 1, 2, . . . , n − 1.
2. The length of a walk W = (u = v1 , v2 , . . . , vn = v) is the number of edges in W , that is n.
3. A walk W = (v1 , v2 , . . . , vn ) is called a trail if all the edges are distinct.
4. A walk W = (u = v1 , v2 , . . . , vn = v) is called a path P if all the vertices, and hence the
edges, are distinct. we call u and v as the end vertices of P and the remaining vertices
on P as the internal vertices.
path is also a trail
5. A walk W = (u = v1 , v2 , . . . , vn = v) is called closed if u = v.vertex same
6. A cycle C is a closed path. trail
special trail in which origin and terminating vertex are same but internal vertex are distinct
7. The graph G is called connected if for any vertices u, v ∈ V , there is a u-v path.

4 Some Graph Invariants


Let G = (V, E) be a graph. Then we define some terms associated to G.
Definition 4.1 Let u, v ∈ V . The distance between u and v is denoted as d(u, v) and is defined
as the length of shortest path between u and v. If there is no such path then d(u, v) = ∞.
Definition 4.2 Let u ∈ V . The eccentricity of u, denoted as e(u), is defined as e(u) =
max{d(u, v) | v ∈ V }. If G is disconnected then eccentricity of each vertex is ∞.
Definition 4.3 The radius of G, denoted as r(G), is defined as r(G) = min{e(v) | v ∈ V }.
Since the eccentricity of every vertex in a disconnected graph is infinity, hence the radius of a
disconnected graph will be infinity.
Definition 4.4 The diameter of G, denoted as diam(G), is defined as diam(G) = max{e(v) | v ∈
V }. The diameter of a disconnected graph is ∞.
Definition 4.5 A point u ∈ V is called a center point of (G) if e(u) = r(G). A collection of
all the center points is called the center of G and is denoted as C(G).

2
Definition 4.6 The number of edges in the longest cycle of G is called as the circumference
of G.
Definition 4.7 The number of edges in the shortest cycle of G is called its girth and is denoted
as g(G). If G has no cycle then g(G) = ∞.
Definition 4.8 A complete subgraph of G is called a clique of G. The maximum order of a
clique is called the clique number of G and is denoted as ω(G).
Definition 4.9 A graph which is not connected is called disconnected. If G is a disconnected
graph, then a maximal connected subgraph of G is called a component or connected component
of G.
Proposition 2 Let P and Q be two different u-v paths in G. Then, P ∪ Q contains a cycle.
condition for
disconnecte
Proposition 3 Every graph G containing a cycle satisfies g(G) ≤ 2 diam (G) + 1. d graph
Proposition 4 Let G = (V, E) be a non-null graph. Then G is disconnected if and only if the
vertex set V can be partitioned into two parts, say V1 , V2 , such that if e = uv ∈ E then either
both u, v ∈ V1 or both u, v ∈ V2 .
Proof: Suppose V can be partitioned into two parts V1 and V2 , satisfying the stated condition
in the proposition. Since, V1 and V2 are non-empty, let u ∈ V1 and v ∈ V2 . Let P be path
joining u and v. There there is an edge e = xy such that x ∈ V1 and y ∈ V2 . This contradicts
the assumption that either both x, y ∈ V1 or x, y ∈ V2 . Hence no such P exists.
Conversely, let us assume that G is disconnected. Now fix a vertex u ∈ V and consider
V1 = {x ∈ V | d(u, v) = ∞}. Since G is disconnected, V1 is a proper subset of V and hence the
set V2 = V \ V1 is non-empty subset of V . Clearly, V1 and V2 give a partition of V fulfilling the
given condition. This completes the proof.

3
Lecture-(19-21)

1 Trees
Definition 1.1 A connected graph G is called a tree if it has no cycles. A collection of trees
is called a forest.
We now prove that the following statements that characterize trees are equivalent.
Theorem 1.1 Let G = (V, E) be a graph on n vertices and m edges. Then the following
statements are equivalent for G.
1. G is a tree.
2. Let u, v ∈ V . Then there is a unique path from u to v.
3. G is connected and n = m + 1.
Proof: 1 implies 2: Since G is connected, for each u, v ∈ V , there is a path from u to v. On
the contrary, let us assume that there are two distinct paths P1 and P2 that join the vertices u
and v. Since P1 and P2 are distinct and both start at u and end at v, there exist vertices, say
u0 and v0 , such that the paths P1 and P2 take different edges after the vertex u0 and the two
paths meet again at the vertex v0 (note that u0 can be u and v0 can be v). In this case, we
see that the graph G has a cycle consisting of the portion of the path P1 from u0 to v0 and the
portion of the path P2 from v0 to u0 . This contradicts the assumption that G is a tree (it has
no cycle).
2 implies 3: Since for each u, v ∈ V , there is a path from u to v, the connectedness of G
follows. We need to prove that n = m + 1. We prove this by induction on the number of
vertices of a graph. The result is clearly true for n = 1 or n = 2. Let the result be true for
all graphs that have n or less than n vertices. Now, consider a graph G on n + 1 vertices that
satisfies the conditions of Item 2. The uniqueness of the path implies that if we remove an edge,
say e ∈ E, then the graph G will become disconnected. That is, G \ e will have exactly two
components. Let the number of vertices in the two components be n1 and n2 . Then n1 , n2 ≤ n
and n1 + n2 = n + 1. Hence, by induction hypothesis, the number of edges in G − e equals
(n1 −1)+(n2 −1) = n1 +n2 −2 = n−1 and hence the number of edges in G equals n−1+1 = n.
Thus, by the principle of mathematical induction, the result holds for all graphs that have a
unique path from each pair of vertices.
3 implies 1: It is already given that G is a connected graph. We need to show that G has no
cycle. So, on the contrary, let us assume that G has a cycle of length k. Then this cycle has
k vertices and k edges. Now, consider the n − k vertices that do not lie of the cycle. Then for
each vertex (corresponding to the n − k vertices), there will be a distinct edge incident with it
on the smallest path from the vertex to the cycle. Hence, the number of edges will be greater
than or equal to k + (n − k) = n. A contradiction to the assumption that the number of edges
equals n − 1. Thus, the required result follows.

1
As a next result in this direction, we prove that a tree has at least two pendant (end) vertices.
Theorem 1.2 Let G be a non-trivial tree. Then G has at least two vertices of degree 1.
Proof: Let G = (V, E) with |V | = n ≥ 2. Then, by above theorem, |E| = n − 1. Also, by
handshake lemma, we know that 2|E| = v∈V deg(v). Thus, 2(n − 1) = ni=1 deg(vi ). Now, G
P P

is connected implies that deg(v) ≥ 1, for all v ∈ V and hence the above equality implies that
there are at least two vertices for which deg(v) = 1. This ends the proof of the result.

2 Eulerian Graphs
Definition 2.1 Let G be a graph. A closed trail (v0 , v1 , . . . , vk , v0 ) is called Eulerian trail if it
contains all the edges of the graph. A graph G is said to be Eulerian if it has an Eulerian trail.
Theorem 2.1 Let G be a connected graph. Then the following statements are equivalent.
1. G is Eulerian.
2. Every vertex of G has even degree.
3. The set of edges can be partitioned into cycles.
Proof: 1 ⇒ 2: Let G be Eulerian graph. Then G has an Eulerian trail T . Note that for each
vertex v the trail enters through an edge and departs v from another edge. Thus at each stage,
the process of coming in and going out contribute 2 to the degree of v. Since each edge appears
exactly once, the degree of v is even.
2 ⇒ 3: Since G is connected and the degree of each vertex even, the graph is not a tree. So there
is at least one cycle C1 in G. If C1 is not G. Let G1 be the subgraph (possibly disconnected) of
G after deleting the edges in C1 . Since each vertex in a cycle has degree 2, the degree of each
vertex in G1 has even and as before has a cycle C2 . Let G2 = G1 − C2 . We repeat the process
of identifying the cycles until we get the graph Gk = G − C1 − C2 − · · · − Ck with no edges.
Thus the set of edges of these cycles gives the required partition.
3 ⇒ 1: Suppose the set of edges in a connected graph G is the disjoint union of k cycles.
Consider any one of these cycles, say cycle C1 . Since the graph is connected, there is a cycle,
say C2 , such that the two cycles have a vertex v1 in common. Let Q12 be the circuit that
consists of all the edges in these two cycles. As before, there is a cycle C3 such that this cycle
and the circuit Q12 have no edge common but do have vertex v2 in common. Let Q123 be the
circuit that contains all the edges of these three edge-disjoint cycles. We repeat this process
until we get a circuit that contains all the edges of the graph. Thus graph is Eulerian.
Theorem 2.2 Let G be a connected graph with exactly two vertices of odd degree. Then,
there is an Eulerian walk starting at one of those vertices and ending at the other.
Proof: Let x and y be the two vertices of odd degree and let v be a symbol such that v ∈
/ V (G).
Then, the graph H with V (H) = V (G) ∪ {v} and E(H) = E(G) ∪ {xv, yv} has each vertex of
even degree and hence by Theorem 9.5.2, H is Eulerian. Let Γ = (v, v1 = x, . . . , vk = y, v) be
an Eulerian tour. Then, Γ − v is an Eulerian walk with the required properties.

2
3 Hamiltonian Graph
Definition 3.1 Let G be a graph. A cycle in G is said to be Hamiltonian if it contains all
vertices of G. If G has a Hamiltonian cycle, then G is called a Hamiltonian graph.
Theorem 3.1 (A necessary condition for a graph to be Hamiltonian). If G = (V, E)
is Hamiltonian and if W is any nonempty subset of V , the graph G − W has at most |W |
components.
The converse of the above theorem does not hold always. Consider the complete bipartite graph
K2,n for n ≥ 2.
Theorem 3.2 (Ore’s Theorem: A sufficient condition for a graph to be Hamilto-
nian). A simple graph with n vertices (where n > 2) is Hamiltonian if the sum of the degrees
of every pair of non-adjacent vertices is at least n.
Theorem 3.3 (Dirac’s Theorem: A sufficient condition for a graph to be Hamilto-
nian). A simple graph with n vertices (where n > 2) is Hamiltonian if the degree of every
vertex is at least n/2.
Proof: If each degree is at least n/2, the sum of every pair of vertices is at least n. Then the
proof follows by Ore’s theorem.

4 Planar Graph
A graph is said to be embedded on a surface S when it is drawn on S so that no two edges
intersect, except at end points. A graph is said to be planar if it can be embedded on the plane.
Examples:
1. A tree is embeddable on a plane.
2. Any cycle Cn , n ≥ 3 is planar.
3. The K4 is planar.
4. The K2,3 is planar.
5. Draw a planar embedding of K5 − e, where e is any edge.
6. Draw a planar embedding of K3,3 − e, where e is any edge.
Definition 4.1 Consider a planar embedding of a graph G. The regions on the plane defined
by this embedding are called faces/regions of G. The unbounded face/region is called the
exterior face.
Theorem 4.1 Let G be a connected planar graph with v number of vertices, e number of edges
and f number of faces. Then v − e + f = 2.
Proof: We use induction on f . Let f = 1. Then G cannot have a subgraph isomorphic to
a cycle. For, if G has a subgraph isomorphic to a cycle, then in any planar embedding of G,
f ≥ 2. Therefore, G is a tree, and hence v − e + f = v − (v − 1) + 1 = 2. Assume that the

3
equation is true for all plane connected graphs having 2 ≤ f < n. Let G be a connected planar
graph with f = n. Choose an edge that is not a cut-edge, say e. Then, G − e is still a connected
graph. Notice that the edge e is incident with two separate faces. So, its removal will combine
the two faces, and hence G − e has only n − 1 faces. Thus, using the induction hypothesis
v − e + f = 2. Hence the required result follows.
Corollary 1 If G is a connected planar simple graph with e edges and v vertices, where v ≥ 3,
then e ≤ 3v − 6.
Proof: Each face has at least 3 edges and each edge is participated in two faces. So, the
number of edges e ≥ 3f /2. Now the proof follows by v − e + f = 2.
Exercise: Show that K5 is non planar.
Corollary 2 If a connected planar simple graph has e edges and v vertices with v ≥ 3 and no
circuit of length three, then e ≤ 2v − 4. e>=4f/2

Proof: Note that e ≥ 2f . Then the proof follows by v − e + f = 2.


Exercise: K3,3 is non planar.
Definition 4.2 If a graph is planar, so will be any graph obtained by removing an edge uv
and adding a new vertex w together with edges uw and wv. Such an operation is called an
elementary subdivision. The graphs G1 = (V1 , E1 ) and G2 = (V2 , E2 ) are called homeomorphic
if they can be obtained from the same graph by a sequence of elementary subdivisions.
Theorem 4.2 (Kuratowski’s Theorem) A graph is non planar if and only if it contains a
subgraph homeomorphic to K3,3 or K5 .

4
Lecture-22-24

Definition: A group is a pair (G, ∗), where G is a set, ∗ is a binary operation and the
following axioms hold:
1. (The associative law)

(a ∗ b) ∗ c = a ∗ (b ∗ c) for all a, b, c ∈ G.

2. (Existence of an identity) There exists an element e ∈ G with the property that


e ∗ a = a and a ∗ e = a for all a ∈ G.
3. (The existence of an inverse) For each a ∈ G there exists an element b ∈ G such that

a ∗ b = b ∗ a = e.

Remark: Notice that ∗ : G × G → G is a binary operation and thus the ‘closure axiom’:
a, b ∈ G =⇒ a ∗ b ∈ G is implicit in the definition.
Definition: We say that a group (G, ∗) is abelian or commutative if a ∗ b = b ∗ a for all
a, b ∈ G.
Proposition:
• (Uniqueness of the Identity:) The identity e is the unique element in G : To see
this suppose we have another identity f. Using the fact that both of these are identities
we see that
f = f ∗ e = e.
We will usually denote this element by 1 (or by 0 if the group operation is commutative).
• (Uniqueness of Inverses:)The inverse b ∈ G of a ∈ G is unique. To see this suppose
that c is another inverse to a. Then

c = c ∗ e = c ∗ (a ∗ b) = (c ∗ a) ∗ b = e ∗ b = b.

We call this unique element b, the inverse of a. It is often denoted a−1 (or −a when
the group operation is commutative). For simplicity, we write ab for a ∗ b.
• (Cancellation:) In a group G, the right and left cancellation laws hold; that is,
ba = ca implies b = c, and ab = ac implies b = c.
Proof: Suppose ba = ca. Let a−1 be an inverse of a. Then, multiplying on the right
by a−1 yields (ba)a−1 = (ca)a−1 . Associativity yields b(aa−1 ) = c(aa−1 ). Then, be = ce
and, therefore, b = c as desired. Similarly, one can prove that ab = ac implies b = c by
multiplying by a−1 on the left.
• (Socks-Shoes Property:) For group elements a and b, (ab)−1 = b−1 a−1 .
Proof: Since (ab)(ab)−1 = e and (ab)(b−1 a−1 ) = a(bb−1 )a−1 = aea−1 = aa−1 = e, we
have by uniqueness of inverses that (ab)−1 = b−1 a−1 .

1
Examples:
• The group of integers (Z, +) and Q, R, C with respect to addition are abelian groups.
• The set R∗ of nonzero real numbers is a group under ordinary multiplication. The
if n-j is identity is 1. The inverse of a is 1/a.
inverse of j
then • The set Zn = {0, 1, . . . , n} for n ≥ 1 is a group under addition modulo n. For any
(j+(n-j)) j ∈ Zn , the inverse of j is n − j. This group is usually referred to as the group of
%n=e==0
but 0 is not integers modulo n. (a+b)%n=a...b=e
the identity if b=0, n%n=0!=n....e???
element for • The set {1, 2, . . . , n − 1} is a group under multiplication modulo n if and only if n is
a=n prime. e=1 , inverse of a will be 1/a
• The subset {1, −1, i, −i} of the complex numbers is a group under complex multi-
plication. Note that −1 is its own inverse, whereas the inverse of i is −i, and vice
versa.
• Let X be a set and let Sym(X) be the set of all bijective maps from X to itself. Then
Sym(X) is a group with respect to composition, ◦, of maps. This group is called the
symmetric group on X and we often refer to the elements of Sym(X) as permutations
of X. When X = {1, 2, 3, . . . , n} the group is often denoted Sn and called the symmetric
group on n letters.
• The set of all n × n matrices with determinant 1 with entries from Q (rationals), R
det=1
thats (reals), C (complex numbers), or Zp (p a prime) is a non-Abelian group under matrix
why multiplication. This group is called the special linear group of n × n matrices over
inverse Q, R, C, or Zp , respectively. commutative property doesn't holds here e=Identity matrix and
exist inverse of A is A^-1
• The set of all 2 × 2 matrices with real number entries is not a group under the matrix
multiplication operation. Inverses do not exist when the determinant is 0.
• The set {0, 1, 2, 3} is not a group under multiplication modulo 4. Although 1 and 3
have inverses, the elements 0 and 2 do not.
• The set of integers under subtraction is not a group, since the operation is not asso-
ciative.
Subgroup: Let G be a group with a subset H. We say that H is a subgroup of G if the
following two conditions hold:
• e ∈ H,
• If a, b ∈ H then ab, a−1 ∈ H.
Note: One can replace the above conditions with the more economical:
• H ̸= ∅,
• If a, b ∈ H then a−1 b ∈ H.
Remark: It is not difficult to see that one could equivalently say that H is a subgroup of G
if H is closed under the group multiplication ∗ and that H with the induced multiplication

2
of ∗ on H is a group in its own right.

Order of a Group: The number of elements of a group (finite or infinite) is called its order.
We will use |G| to denote the order of G.
Example: The group Z of integers under addition has infinite order, whereas the group
U (10) = {1, 3, 7, 9} under multiplication modulo 10 has order 4.

Order of an Element: The order of an element g in a group G is the smallest positive


integer n such that g n = e. (In additive notation, this would be ng = 0.) If no such integer
exists, we say that g has infinite order. The order of an element g is denoted by |g|.
Example:
• Consider Z10 under addition modulo 10. Since 2 + 2 = 4, 2 + 2 + 2 = 6, 2 + 2 + 2 + 2 =
8, 2 + 2 + 2 + 2 + 2 = 0, we know that |2| = 5. Similar computations show that
|0| = 1, |7| = 10, |5| = 2, |6| = 5.
Cyclic Group: A group G is called cyclic if there is an element a in G such that G = {an :
n ∈ Z}. Such an element a is called a generator of G.

Coset of H in G: Let G be a group and let H be a subset of G. For any a ∈ G, the set
{ah : h ∈ H} is denoted by aH. Analogously, Ha = {ha : h ∈ H} and aHa−1 = {aha−1 :
h ∈ H}. When H is a subgroup of G, the set aH is called the left coset of H in G containing
a, whereas Ha is called the right coset of H in G containing a. In this case, the element
a is called the coset representative of aH (or Ha). We use |aH| to denote the number of
elements in the set aH, and |Ha| to denote the number of elements in Ha.

Properties of Cosets: Let H be a subgroup of G, and let a and b belong to G. Then,


1. a ∈ aH, e belongs to H -> a*e belongs to aH

2. aH = H if and only if a ∈ H,
3. aH = bH if and only if a ∈ bH
4. aH = bH or aH ∩ bH = ∅,
5. aH = bH if and only if a−1 b ∈ H,
6. |aH| = |bH|,
7. aH = Ha if and only if H = aHa−1 ,
8. aH is a subgroup of G if and only if a ∈ H.
Suppose G is a group with a subgroup H. We define a relation on G as follows:

xRy iff x−1 y ∈ H.

y E xH
x E yH
-> if xRy this means xH=yH
->reflexive- aRa iff a^-1*a E H .....when e E H (always 3as H is a subgroup)
->symmetric - xRy-> xH=yH, yH=xH means y^-1x E H therefore yRx
->transmitive- xRy-> xH=yH, yRz->yH=zH......therefore xH=zH which means x^-1z E H....therefore xRz
x^-1y E H..... y E xH......xRy so y is the equivalence class of x

This relation is an equivalence relation. Notice that xRy if and only if x−1 y ∈ H if and only
if y ∈ xH. Hence the equivalence class of x is [x] = xH, the left coset of H in G.
for aES , [a] = {x : (a, x) E R}

Lagrange’s Theorem: Let G be a finite group with a subgroup H. Then |H| divides |G|.
Proof: Using the equivalence relation above, G gets partitioned into pairwise disjoint equiv-
alence classes, say
G = a1 H ∪ a2 H ∪ · · · ∪ ar H
and adding up we get

|G| = |a1 H| + |a2 H| + · · · + |ar H| = r|H|.

Notice that the map from G to itself that takes g to ai g is a bijection (the inverse is the map
i g) and thus |ai H| = |H|.
g → a−1

Definition: (Normal Subgroup) A subgroup H of G is said to be a normal subgroup if

g −1 Hg ⊆ H ∀g ∈ G.

Definition: Let G be a group with a subgroup H. The number of left cosets of H in G is


called the index of H in G and is denoted [G : H].
Examples: abelian->commutative law holds->Hg=gH->
g^-1Hg = g^-1gH = H
• Every subgroup N of an abelian group G is normal.
if H=e then g^-1Hg =
• The trivial subgroup {e} and G itself are always normal subgroups of G. g^-1eg = g^-1g = e

• If H is a subgroup of G such that [G : H] = 2 then H is normal subgroup of G.


Definition: Let (G, ∗), (H, ◦) be groups. A map Φ : G → H is a homomorphism if

Φ(a ∗ b) = Φ(a) ◦ Φ(b)

for all a, b ∈ G. Furthermore Φ is an isomorphism if it is bijective.


Example: Let R+ be the set of all the postive real numbers. There is a (well-known)
isomorphism Φ : (R, +) → (R+ , .) given by Φ(x) = ex .

4
Lecture-25

Definition: A ring R is a set with two binary operations, addition (denoted by a + b) and
multiplication (denoted by ab), such that for all a, b, c in R :
1. a + b = b + a.
2. (a + b) + c = a + (b + c).
3. There is an additive identity 0. That is, there is an element 0 in R such that a + 0 = a
for all a in R.
4. There is an element −a in R such that a + (−a) = 0.
5. Associative Property: a(bc) = (ab)c.
6. Distributive Property: a(b + c) = ab + ac and (b + c)a = ba + ca.
The above can be summarize as follows: a ring is an Abelian group under addition, also
having an associative multiplication that is left and right distributive over addition.
Definition: We say that a ring (R, +, .) is commutative if a.b = b.a for all a, b ∈ R.
Definition: A unity (or multiplicative identity) in a ring is a nonzero element that is an
identity under multiplication. A nonzero element of a commutative ring with unity need not
have a multiplicative inverse.

Theorem: (Rules of Multiplication)- Let a, b, and c belong to a ring R. Then


• a0 = 0a = 0.
• a(−b) = (−a)b = −(ab).
• (−a)(−b) = ab.
• a(b − c) = ab − ac and (b − c)a = ba − ca.
Furthermore, if R has a unity element 1, then
• (−1)a = −a.
• (−1)(−1) = 1.
Examples:
• The sets Z, Q, R and C with respect to usual addition and usual multiplication are
rings.
• The set Zn = {0, 1, . . . , n − 1} for n ≥ 1 under addition and multiplication modulo n
is a commutative ring with unity 1.
• The set Z[x] of all polynomials in the variable x with integer coefficients under ordinary
addition and multiplication is a commutative ring with unity f (x) = 1.

1
• The set 2Z of even integers under ordinary addition and multiplication is a commutative
ring without unity.
• The set M2 (Z) of 2 × 2 matrices with integer entries is a noncommutative ring with
unity.
Subring: A subset S of a ring R is a subring of R if S is itself a ring with the operations
of R.
Theorem: (Subring Test) A nonempty subset S of a ring R is a subring if S is closed
under subtraction and multiplication that is, if a − b and ab are in S whenever a and b are
in S.
Examples:
• {0} and R are subrings of any ring R. {0} is called the trivial subring of R.
• For each positive integer n, the set nZ = {0, ±n, ±2n, ±3n, . . . } is a subring of the
integers Z.
• The set of Gaussian integers Z[i] = {a + bi : a, b ∈ Z} is a subring of the complex
numbers C.
Definition: A field F , containing at least two elements, is a set with two binary operations,
addition (denoted by a + b) and multiplication (denoted by ab), such that for all a, b, c in F :
1. a + b = b + a.
2. (a + b) + c = a + (b + c).
3. There is an additive identity 0. That is, there is an element 0 in R such that a + 0 = a
for all a in R.
4. There is an element −a in R such that a + (−a) = 0.
5. (Associativity of multiplication) a(bc) = (ab)c.
6. (Distributivity of multiplication) a(b + c) = ab + ac and (b + c)a = ba + ca.
7. (Commutativity of multiplication) ab = ba.
8. (Existence of a multiplicative identity) There is an element 1 ∈ F, such that 1 ̸= 0 and
a.1 = a.
9. (Existence of a multiplicative inverses) If x ̸= 0, then there is an element x−1 ∈ F such
that xx−1 = 1.
Examples:
• The sets Q, R and C with respect to usual addition and usual multiplication are fields.
• The set Zp = {0, 1, . . . , p − 1} for p ≥ 2 under addition and multiplication modulo p is
a field, where p is a prime number.
• The set Z of integers is not a field.

You might also like