You are on page 1of 31

Chapter 1

Introduction to Logic

Before anything else, this course covers material indispensable to any mathematician:
it could even be subtitled “What Every Mathematician Needs to Know.” It includes
material necessary for students to succeed in upper-division mathematics courses, and
more importantly, the analytical tools necessary for thinking like a mathematician.
Mathematics is not a spectator sport. Mastering the topics covered is important; but
more important is enhancing your ability to think. Reading mathematics is a prerequisite
to thinking mathematically. Hence, for each topic, you should:

1. Read the section from beginning to end,

2. Carefully study each example, while working out the details with paper and pencil,

3. Work each practice problem before looking up the solution,

4. Finally, WORK the assigned exercises and problem sets.

1.1 Understanding Proofs


Mathematicians and scientists need the ability to read and understand valid arguments
(or proofs) and to recognize invalid arguments. A basic knowledge of logic is indispensable
for analyzing and constructing proofs.
In all mathematical proofs, there is a collection of statements called hypotheses and
a statement, called the conclusion, which must be proved to follow logically from the
hypotheses. There are many forms a proof can take. However, every proof must be a
finite sequence of statements that are either hypotheses, previously proved statements,
or statements that follow logically from previous statements in the proof. The final
statement in a proof should be the statement to be proved.
Two major features of a proof are the idea behind the proof—the creative part, what
really makes it work—and the written part. The written part must be done so that other
mathematicians can read and understand the proof. To become a successful mathemati-
cian, one must learn to communicate ideas, both verbally and in written form, and also
to understand proofs created by others.
To understand a proof, it is necessary to know what is to be proved. That is, we
must know what the proof is “all about.” For example, is it a proof about sets or about
functions in calculus? A proof about sets will require facts and definitions of set theory.

1
A proof involving functions in calculus will require definitions such as that of a limit of
a function.
When reading any proof, we must always ask:
What is the goal of the proof?
What are the hypotheses?
What definitions are necessary?
What previously proved facts or laws of logic are used in the proof?
To illustrate these ideas, let us consider some examples of proofs. We need the fol-
lowing definitions: an even integer is an integer that can be expressed in the form 2k,
where k is an integer; an odd integer is an integer that can be expressed in the form
2k + 1, where, again, k is an integer. For example, if p and q are integers, then 2 (p + q)
is even, whereas 2 (p2 − q) + 1 is odd. We assume that the following is known:
An integer is either even or odd, but not both.

Example 1. Let n be an odd integer. Prove that n2 is an odd integer.

Proof. Let n be an odd integer.


Hence, n = 2k + 1 for some integer k.
Hence,

n2 = (2k + 1)2
= 4k 2 + 4k + 1
= 2 2k 2 + 2k + 1


Since 2k 2 + 2k is an integer, we conclude that n2 is odd.

Practice Problem 1. Let n be even. Prove that n2 is even.

Example 2. Prove: If n is an integer and n2 is even, then n is even.

Practice Problem 2. Prove: If n is an integer and n2 is odd, then n is odd.



Example 3. Let x and y be nonnegative numbers. Prove that xy ≤ (x + y) /2.

The main goal of the proof writer is to convince others that the statement being
proved follows logically from certain assumptions. A secondary goal is to write an elegant,
concise proof. In some ways, these goals compete with one another because what is clear,
concise, and elegant to some readers will be a terse muddle to others. The less experienced
a person is at reading proofs, the more important it is for that person to do scratch work
in order to absorb the ideas behind the proof.
The creation of a proof sometimes requires that we work backwards from the orig-
inal goal, by means of a needs assessment, to a new goal that is closer to the given
facts. Also, we may have to rewrite (but not alter) the given facts in the hypothesis
so that they are closer to the new goal. This parallel process is sometimes called the
backwards/forwards method.

2
1.2 Introduction to Propositional Logic
Symbolic logic can be described as the analytical study of the art of reasoning. Logic
is important because it forms the basis for proof techniques and, therefore, has special
utility for mathematics.
Consider the following English sentences:

• Where is my orange pen?

• Please give me some oranges.

• Oh, what a beautiful morning!

• Easter Sunday is on 1 April 2018.

• The sun sets in the east.

Let us examine each sentence above. The first asks a question (interrogative), the
second commands (imperative) while the third expresses a feeling (exclamatory). Each
one of these three sentences conveys an idea but it cannot be judged as either true or
false. The last two, on the other hand, declare information (declarative) and hence, may
properly be evaluated as either true or false, but not both. Based on our Roman Catholic
calendar, the fourth sentence is true while the fifth is false, based on common knowledge.
In mathematics, sentences such as the fourth and the fifth are called propositions.
Furthermore, in propositional logic, the actual content of such sentences is unimportant;
of primary interest is their truth or falsity.
We, therefore, define a proposition as a declarative sentence which is either true or
false, but not both. The truth value of a proposition is “true” if the proposition is true
and “false” if the proposition is false.

Practice Problem 3. Which of the following are propositions?

(a) Not all trees have green leaves.

(b) How often do you go to the dentist?

(c) Two points determine one line.

(d) Check your solution.

(e) 4 + 1 = 6

(f) It is raining.

Practice Problem 4. Determine the truth value of each of the following propositions.

(a) Fidel V. Ramos is the 12th President of the Republic of the Philippines.

(b) 2 is not a prime number.

(c) Some students in this class are not Math majors.

(d) A square is a rectangle.

3
We shall denote propositions by small letters as illustrated below:
p: A square is a rectangle.
q: One and only one plane can be drawn to contain two distinct lines.
We read the notations above as:
p is the proposition “A square is a rectangle.”
q is the proposition “One and only one plane can be drawn to contain two distinct
lines.”

1.2.1 Operations on Propositions


A proposition may be classified as simple or compound. A simple or prime proposition
makes only one assertion while a compound or composite proposition makes more than
one assertion. For example, “Elizabeth is pretty.” is a simple proposition, while “Justin
is handsome and kind” is a compound proposition. The latter really means “Justin is
handsome and Justin is kind”.
Suppose we have a set of propositions. Are there operations that we can perform
on this set in order to produce new propositions? There are such operations and we
call them logical connectives or simply connectives. Compound propositions can be
constructed from simple ones (or compound ones) by utilizing the following connectives:
Connectives Symbol Usage Propositions Formed
AND ∧ p∧q Conjunction
OR ∨ p∨q Disjunction
IF... THEN... →, ⇒ p→q Conditional
IF AND ONLY IF ↔, ⇔, ≡ p↔q Biconditional
NOT ¬, ∼ ¬p Negation
The first four of these connectives are called binary connectives because they com-
bine two propositions to make one resulting proposition. The last connective is called a
unary connective.

Example 4. Let us use the above operations on the propositions p: “Two is positive”
and q: “Two is even”.
Symbol Compound Proposition
(a) Conjunction p∧q “Two is positive and two is an even number.”
or “Two is a positive even number.”
(b) Disjunction p∨q “Two is positive or two is an even number.”
(c) Conditional p → q “If two is positive, then two is an even number.”
(d) Biconditional p ↔ q “Two is positive if and only if two is an even number.”
(e) Negation ¬p “Two is not a positive number.”

1.2.2 Syntactics
Syntactics is a set of rules used to determine whether a sequence of symbols is a propo-
sition. The syntactical rules for constructing propositions are as follows:

S1. All simple propositions p, q, r, . . . are propositions.

S2. If p and q are propositions, then (p ∧ q) is a proposition.

S3. If p and q are propositions, then (p ∨ q) is a proposition.

4
S4. If p and q are propositions, then (p → q) is a proposition.

S5. If p and q are propositions, then (p ↔ q) is a proposition.

S6. if p is a proposition, then ¬p is a proposition.

No other sequence of symbols is a proposition.


The main connective of a proposition is the last one used to construct the proposition.
Example 5. Show that ((p ∧ ¬q) ∨ r) is a proposition. What is the main connective?
Solution
1. p, q, and r are propositions by S1.

2. ¬q is a proposition by S6 and Step 1.

3. (p ∧ ¬q) is a proposition by S2 and Steps 1 and 2.

4. ((p ∧ ¬q) ∨ r) is a proposition by S3 and Steps 3 and 1.


The main connective is ∨ since it is the one introduced in Step 4.
In what follows, we often omit outside parentheses, since doing so will not cause any
confusion in reading and interpreting a proposition correctly. Also, when a proposition is
complicated, pairs of brackets, [ and ], may be used in place of some pairs of parentheses.
Practice Problem 5. Show that ¬ (p ∧ q) → (¬p ∧ ¬q) is a proposition. What is the
main connective?
Practice Problem 6. Show that ¬p → (q ∧ r) is a proposition. What is the main
connective?
The sequences of symbols (p∧ → q), (¬ ∧ p), and ¬p ∨ q) are not propositions since
there is no way to construct them from Rules S1–S6.

Remarks
The NOT (¬) connective is applied first to the next letter or to the next parenthesized
expression. The AND (∧) and OR (∨) connectives are applied next. The conditional
symbol (→) is applied only after the NOT (¬), AND (∧), and OR (∨) connectives have
been applied. Finally, the biconditional (↔) is applied.

a. ¬p ∧ q does not stand for ¬ (p ∧ q).

b. ¬p → q does not stand for ¬ (p → q).

c. p ∧ q → r stands for (p ∧ q) → r, not for p ∧ (q → r).

d. ¬p ∨ q → r ∧ q stands for (¬p ∨ q) → (r ∧ q), and not for ¬p ∨ (q → r ∧ q) or any


other way of placing the parentheses.

e. p → q ↔ p ∨ r stands for (p → q) ↔ (p ∨ r), and not for p → (q ↔ p ∨ r) or any


other way of placing the parentheses.

f. p ∨ r ∧ t is ambiguous, since it is not clear whether to apply ∨ or ∧ first.

5
g. p → q → r is ambiguous, since either occurrence of → can be applied first.

Do not omit parentheses when there is any chance of confusion, even though the above
conventions permit it. The hierarchical order of the connectives is as follows:
parenthesized expressions
NOT (¬)
AND (∧), OR (∨)
IMPLIES (→)
IF AND ONLY IF (↔)

Example 6. Determine which of the following is ambiguous or not. If it is unambiguous,


put in all missing parentheses.

(a) p ∧ q → ¬q ∨ r

(b) p → ¬q ∨ r

(c) p ∧ q ∨ r → r ∧ p

(d) r → t ↔ r

(e) (p ∧ q) → s → r

1.2.3 Truth Tables


We are interested here in studying only the truth values of propositions. In particular,
how can we determine the truth value of a compound proposition if we are given the
truth values of all its constituent simple propositions? To answer this question, we need
the truth tables for all the connectives.
p q p∧q p∨q p→q q→p p↔q ¬p
T T T T T T T F
T F F T F T F F
F T F T T F F T
F F F F T T T T

Note that an AND proposition, p ∧ q, is true when both the propositions p and q are
true, and false otherwise. An OR proposition, p ∨ q, is false when both the propositions p
and q are false, and true otherwise. An OR proposition is sometimes called an inclusive-
or proposition, since it is true if one or both of its constituent propositions are true. A
NOT proposition, ¬p, is true when the proposition p is false, and false when p is true.

The Conditional and Its Variants


The truth table for p → q, or p implies q, can be understood by the following example.
Suppose you buy a washing machine, and it has a guarantee. Essentially, the guarantee
says: “If the machine breaks, then a repairman will come to fix it.” (or “The machine
breaks implies that a repairman will come to fix it.”) Then to say that the guarantee is
not valid is to say that the proposition p → q is false. The only way that the guarantee
can be invalid is when the machine breaks (p is true) and no repairman comes to fix it
(q is false). So, p → q is false when, and only when, p is true and q is false.

6
The washing machine example illustrates row 2 of the table. Let us examine, for
example, row 3. It says, in the terminology of our example, “The machine is not broken,
but the repairman comes to fix it.” This in no way invalidates the guarantee, and similarly,
neither do rows 1 and 4.
In the proposition p → q, p is called the hypothesis or antecedent while q is called
the conclusion or consequence. Note that rows 3 and 4 of the truth table define an
implication as being true whenever the hypothesis is false, no matter what truth value
the conclusion has. Because of this, these rows are called the vacuously true cases. To
illustrate this idea with another example, consider the proposition “If I play with my new
tennis racket, then I’ll win the match.” As long as the person making this proposition
plays with an old racket, that person cannot be called a liar, no matter how the match
turns out. A proposition that is true because it satisfies a vacuously true case is said to
be a vacuously true proposition.
When the conditional proposition p → q is always true, we write p ⇒ q.
Furthermore, p is referred to as a sufficient condition for q while q is a necessary
condition for p.
Consider the proposition “If I work hard, I shall have a happy life.” This can be
interpreted as saying that working hard guarantees my having a happy life, so that
working hard is a sufficient condition for having a happy life. On the other hand, having
a happy life is the spontaneous consequence of working hard, so that having a happy life
is the necessary condition for working hard.
Finally, we point out that there are many ways of expressing the proposition “if p,
then q”. For example, the proposition “If I receive a bonus, then I shall have a holiday
in Boracay” is rendered by each of the following:

(i) If I receive a bonus, I shall have a holiday in Boracay.

(ii) I shall have a holiday in Boracay if I receive a bonus.

(iii) I shall have a holiday in Boracay provided that I receive a bonus.

(iv) I receive a bonus only if I shall have a holiday in Boracay.

(v) Receiving a bonus is a sufficient condition for a holiday in Boracay.

(vi) Having a holiday in Boracay is a necessary condition for receiving a bonus.

The conditional p → q has three variations, namely:

(1) converse: q → p

(2) inverse: ¬p → ¬q

(3) contrapositive: ¬q → ¬p

Example 7. Consider the following conditional proposition:


“If x is a prime number greater than 2, then x is an odd number.”
The three variants are:

(1) converse: “If x is an odd number, then x is a prime number greater than 2.”

(2) inverse: “If x is not a prime number greater than 2, then x is not an odd number.”

7
(3) contrapositive: “If x is not an odd number, then x is not a prime number greater
than 2.”

Practice Problem 7. Consider this: “If x is your father, then x is a male.”

Practice Problem 8. Let p: A person is a bigamist and q: A person violates the law.
Give the four variations of the conditional.

The Biconditional
The biconditional proposition p ↔ q is defined as (p → q) ∧ (q → p). Not that the
biconditional is true whenever the truth values of p and q are the same.
Just like the conditional, if p ↔ q is a true proposition, then we write p ⇔ q or p ≡ q,
and the propositions p and q are said to be equivalent.
We sometimes write “iff” as an abbreviation for “if and only if”. The biconditional
proposition can be read as follows:

(a) p if and only if q.

(b) p when and only when q.

(c) The necessary and sufficient conditions for p is q.

(d) The necessary and sufficient conditions for q is p.

Example 8. Let
p: “Today is Thursday.”
q: “Tomorrow is Friday.”
r: “It is raining today.”
s: “Today is Wednesday.”
t: “Tomorrow is Thursday.”
Assume today is a rainy Thursday. Find the truth values of the following propositions:

(a) p ∧ r (g) r → t
(b) q ∨ r (h) s → r
(c) r ∧ s
(i) s → t
(d) ¬t ∧ r
(e) p → q (j) p ↔ r

(f) s ∨ t (k) s ↔ r

Example 9. Construct the truth table for ¬ (p ∧ q) → ¬p ∨ ¬q.

Practice Problem 9. Construct the truth table for ¬ (p ∧ q) → ¬p ∧ ¬q.

8
1.3 Logical Equivalence and Tautologies
Before we proceed, a remark — so far, we have been using small letters like p and q to
denote propositions. Sometimes, and in some books, lowercase letters such as p and q are
used to denote simple propositions only while uppercase letters like P and Q are used
to denote propositions of all types, compound or simple. In our discussion however, we
will only be using small letters, p, q, r, . . . and we take them to mean either simple or
compound propositions.
Compound propositions that have the same truth values in all possible cases are called
logically equivalent. We define this notion as follows.

Definition 1. The propositions p and q are called logically equivalent if p ↔ q is a


tautology. The notation p ≡ q denotes that p and q are logically equivalent.

Remark: Two propositions composed of the same simple propositions are logically
equivalent whenever the main columns of their standard truth tables are identical.
The symbol ≡ is not a logical connective since p ≡ q is not a compound proposition,
but rather is the statement that p ↔ q is a tautology. The symbol ⇔ is sometimes used
instead of ≡ to denote logical equivalence.

Example 10.
p q ¬p ¬q p ∧ q ¬ (p ∧ q) ¬p ∨ ¬q
(1) (2) (3) (4) (5) (6) (7)
T T F F T F F
T F F T F T T
F T T F F T T
F F T T F T T
In the above truth table, we see that columns (6) and (7) are the same. This is to say
that the propositions ¬ (p ∧ q) and ¬p ∨ ¬q have the same truth value under any possible
case. Hence, the given two propositions are equivalent.

Intuitively speaking, two propositions are logically equivalent means that two propo-
sitions are of different forms but they express the same idea. For instance:
p: Today is Monday.
q: It is not true that today is not Monday.
Then, p and q are logically equivalent.
Now, let’s consider one more example:
r: 2 is not greater than 3.
s: 2 is less than or equal to 3.
Then, r and s are also logically equivalent.

Definition 2. A tautology is a proposition which is always true, no matter what truth


values its constituent simple propositions have.

Example 11. q → (p ∨ q)

Every row in the main column of a tautology is T. The symbol T will denote a
proposition that always has truth value T. Hence, q → (p ∨ q) ≡ T .

Definition 3. A contradiction is a proposition which is always false, no matter what


truth values its constituent simple propositions have.

9
Example 12. (¬p ∨ q) ∧ (p ∧ ¬q)
Every row in the final column of a contradiction is F, just like the previous example.
The symbol F will be used to denote a proposition that always has truth value F. Hence,
(¬p ∨ q) ∧ (p ∧ ¬q) ≡ F .
Practice Problem 10. Verify that (p → q) ∧ (q → p) ≡ p ↔ q.
By definition, a proposition is not a tautology whenever at least one row of the main
column of its truth table is F. Accordingly, to show that a proposition is not a tautology,
a row in the main column of the truth table for that proposition must be found with a
truth value of F. The combination of truth values assigned to the simple propositions in
any row that produces an F in the main column of the table is called a counterexample.
Practice Problem 11. Show that q ∧ (p → q) → p is not a tautology.
In fact, the above proposition is a well-known fallacy of logic, called the fallacy
of asserting the conclusion. For example, let p: Today is Saturday and q: It is a
weekend. Then q ∧ (p → q) would read “It is a weekend and if today is a Saturday then it
is a weekend.” However, this may very well be true even if today is a Sunday. Therefore,
it doesn’t follow that p is always true.
Following are some important logical equivalencies called the Boolean laws of logic.

1. Identity laws
p∧T ≡p
p∨F ≡p

2. Domination laws
p∨T ≡T
p∧F ≡F

3. Idempotent laws
p∧p≡p
p∨p≡p

4. Double Negation law


¬ (¬p) ≡ p

5. Commutative laws
p∧q ≡q∧p
p∨q ≡q∨p

6. Associative laws
(p ∧ q) ∧ r ≡ p ∧ (q ∧ r)
(p ∨ q) ∨ r ≡ p ∨ (q ∨ r)
Note: The connectives ∧ and ∨ are both associative, hence parentheses are not
necessary. i.e. p ∧ q ∧ r and p ∨ q ∨ r have no ambiguity. However, the parentheses
cannot be omitted for the distributive law.

10
7. Distributive laws
p ∧ (q ∨ r) ≡ (p ∧ q) ∨ (p ∧ r)
p ∨ (q ∧ r) ≡ (p ∨ q) ∧ (p ∨ r)

8. De Morgan’s laws
¬ (p ∧ q) ≡ ¬p ∨ ¬q
¬ (p ∨ q) ≡ ¬p ∧ ¬q

9. Absorption laws
p ∧ (p ∨ q) ≡ p
p ∨ (p ∧ q) ≡ p

10. Complement laws


p ∧ ¬p ≡ F
p ∨ ¬p ≡ T
¬T ≡ F
¬F ≡ T

We add some more logical equivalences involving implications to this list.

11. Transposition or Contrapositive law


p → q ≡ ¬q → ¬p.

12. Switcheroo law or the Or-form of an implication


p → q ≡ ¬p ∨ q.

13. p ∨ q ≡ ¬p → q

14. p ∧ q ≡ ¬ (p → ¬q)

15. ¬ (p → q) ≡ p ∧ ¬q

16. (p → q) ∧ (p → r) ≡ p → (q ∧ r)

17. (p → r) ∧ (q → r) ≡ (p ∨ q) → r

18. (p → q) ∨ (p → r) ≡ p → (q ∨ r)

19. (p → r) ∨ (q → r) ≡ (p ∧ q) → r
And some more logical equivalences involving biconditionals...

20. p ↔ q ≡ (p → q) ∧ (q → p)

21. p ↔ q ≡ ¬p ↔ ¬q

22. p ↔ q ≡ (p ∧ q) ∨ (¬p ∧ ¬q)

23. ¬ (p ↔ q) ≡ p ↔ ¬q

11
Example 13. Write down the negation of the following:

(a) The sun is spherical and the plane can fly.

(b) London is not the capital of China or the house is made of wood.

Example 14. Let p, q, r, s be four propositions. Find the negation of (p ∧ ¬q)∨(¬r ∧ ¬s).

Example 15. Let p, q be two propositions. Simplify [(p ∧ q) ∨ q] ∧ p.

Example 16. Show that ¬ (p ∨ (¬p ∧ q)) and ¬p ∧ ¬q are logically equivalent.

Example 17. Show that (p ∧ q) → (p ∨ q) is a tautology.

12
1.4 Predicates and Quantifiers
In review, propositions are declarative sentences that have a truth value; in other words,
they are either true or false, but not both. For example, the sentence “−1 < 0” is a
proposition, which happens to be true in the context of the real numbers. Many other
sentences used in mathematics, however, are not propositions. Sentences like “x < 5”
and “sin x = 1” are neither true nor false because they contain a variable that denotes
no particular object.
A variable is used to symbolize an arbitrary element from a given domain or universal
set U . That is, elements from U can be substituted for x in a sentence like “x > 3.”
The sentence “x is greater than 3” has two parts. The first part, the variable x, is the
subject of the sentence. The second part—the predicate, “is greater than 3”—refers to
a property that the subject of the sentence can have. We can denote the sentence “x is
greater than 3” by P (x), where P denotes the predicate “is greater than 3” and x is the
variable. The sentence P (x) is also said to be the value of the propositional function
P at x. Once a value has been assigned to the variable x, the sentence P (x) becomes a
proposition and has a truth value. Consider Example 1.
Example 18. Let P (x) denote the statement “x > 3”. What are the truth values of
P (4) and P (2)?
We can also have statements that involve more than one variable. For instance,
consider the statement “x = y + 3.” We can denote this statement by Q (x, y), where x
and y are variables and Q is the predicate. When values are assigned to the variables x
and y, the statement Q (x, y) has a truth value.
Example 19. Let Q (x, y) denote the statement “x = y + 3.” What are the truth values
of the propositions Q (1, 2) and Q (3, 0)?
Example 20. Let R (z, y, z) denote the statement “x+y=z.” What are the truth values
of the propositions R (1, 2, 3) and R (0, 0, 1)?
In general, a statement involving the n variables x1 , x2 , . . . , xn can be denoted by
P (x1 , x2 , . . . , xn ). A statement of the form P (x1 , x2 , . . . , xn ) is the value of the propo-
sitional function P at the n-tuple (x1 , x2 , . . . , xn ), and P is also called a predicate.
Example 21. Consider the following statements:
1. x · 0 = 0
2. x · 5 = 0
3. x = 0 implies x · 5 = 0
4. x < 5
The last item in the previous example shows us that a solution set depends on both
the predicate and the given universal set.
The sentences in the following example are statement and NOT predicates.
Example 22. The x that seems to be a variable in each of the following statements is
not for substitution (try it) and is frequently called a dummy or bound variable.
1. {x | x2 − 3x + 2 = 0} = {1, 2}
2. lim (3x − 2) = 4
x→2

13
1.4.1 Quantifiers
When all the variables in a propositional function are assigned values, the resulting state-
ment becomes a proposition with a certain truth value. However, there is another impor-
tant way, called quantification, to create a proposition from a propositional function.
Two types of quantification will be discussed here, namely, universal quantification and
existential quantification. The area of logic that deals with predicates and quantifiers is
called the predicate calculus or predicate logic.

Universal Quantifier
Many mathematical statements assert that a property is true for all values of a variable
in a particular domain, called the universe of discourse or the universal set or the
domain. Such a statement is expressed using a universal quantification.

Definition 4. The universal quantification of P (x) is the proposition

“P (x) is true for all values of x in the domain or universe of discourse.”

The notation
∀xP (x)
denotes the universal quantification of P (x). Here, ∀ is called the universal quantifier.
∀x means “for all x”, “for every x”, “for any x”, or “for each x”.

Example 23. Assume that the universe of discourse is R. Then:

1. ∀x (x · 0 = 0) means “for all real numbers x, it is true that x · 0 = 0” or “any real


number multiplied by 0 is equal to 0.”

2. ∀x sin2 x + cos2 x = 1 means “for every real number x, sin2 x+cos2 x = 1” or “the


sum of the squares of the sine and cosine of any real number is equal to 1.” This
kind of assertion is made in every trigonometric identity.

3. ∀x (x + 1 > x) is true.

4. ∀x (x < 2). “Q (x) : (x < 2)” is not true for every real number x, since, for instance,
Q (3) is false. Thus, ∀xQ (x) is false.

5. ∀x (x2 ≥ x). What if U = Z?

Example 24. What does the statement ∀xT (x) mean if T (x) is “x has two parents”
and the domain consists of all people?

To show that a statement of the form ∀xP (x) is false, where P (x) is a propositional
function, we need only find one value of x in the universe of discourse for which P (x) is
false. Such a value of x is called a counterexample to the statement ∀xP (x). This is
shown in Example 23 #4 above.
Hence, the statement ∀xP (x) is false if P (x) is false for at least one x in U .

14
Existential Quantifier
Many mathematical statements assert that there is an element with a certain property.
Such statements are expressed using existential quantification.

Definition 5. The existential quantification of P (x) is the proposition

“There exists an element x in the universe of discourse such that P (x) is true.”

We use the notation


∃xP (x)
for the existential quantification of P (x). Here, ∃ is called the existential quantifier.
∃x is read as “there is (at least) one x”, “there exists (at least) one x”, or “for some x”.

Example 25. Assume that U = R.

1. ∃x (x · 5 = 7) means that x · 5 = 7 has a solution. In this case, there is exactly one


solution, 7/5.

2. ∃x (x > 3). Since “P (x) : (x > 3)” is true—for instance, when x = 4—the existen-
tial quantification is true.

3. ∃x (x = x + 1) is false.

Example 26. Differentiate the following statements:

1. “There is one coin in your pocket.”

2. “There is only one coin in your pocket.”

3. “There is one and only one coin in your pocket.”

Theorem 1. Let P (x) be any predicate, then

1. ¬ (∀x) (P (x)) ≡ (∃x) (¬P (x));


That is, “There exists an x for which P (x) is false.”

2. ¬ (∃x) (P (x)) ≡ (∀x) (¬P (x)).


That is, “P (x) is false for every x.”

Corollary 2. Let P (x) be any predicate, then

1. ¬ (∀x) (¬P (x)) ≡ (∃x) (P (x));

2. ¬ (∃x) (¬P (x)) ≡ (∀x) (P (x)).

The universal quantifier can be considered the generalized “and”, and the existential
quantifier can be considered the generalized “or”. To see this, let U be the two-element
set {0, 1}. Then ∀xP (x) is true if and only if P (0) ∧ P (1) is true, and ∃xP (x) is true
if and only if P (0) ∨ P (1) is true.

Example 27. What are the negations of the statements “There is an honest politician”
and “All Filipinos eat rice”?

15
Example 28. What are the negations of the statements ∀x (x2 > x) and ∃x (x2 = 2)?
Example 29. Express the following statements using predicate and quantifiers.
1. “Every student in this class has studied calculus.”
2. “Some student in this class has visited Baguio.”
3. “Every student in this class has visited either Manila or Davao.”
Example 30. Write these statements in symbols using the predicates:
F (x) : x ends in the digit 5; S (x) : x is a perfect square; P (x) : x is positive; N (x) : x
is negative.
Assume that all variables are integers.

1. Perfect squares are positive.


2. Some negative numbers end in the digit 5.
3. Exactly one number is positive.
4. No positive numbers are negative.
5. A number is a perfect square if and only if it does not end in 5.

1.4.2 Nested Quantifiers


Consider, for example, the proposition “Given any number, say x, there is a number, say
y, such that x + y = 0.” Expressed symbolically, this becomes “∀x∃y (x + y = 0)”. Here,
the main connective is “∀x”; therefore, the given proposition is true if and only if the
proposition “∃x (x + y = 0)” generates only true propositions. For example, substituting
3 for x, we obtain the proposition “∃y (3 + y = 0)”. This proposition is true, of course,
if we substitute −3 for y. Similarly, each of the propositions generated by ∃x (x + y = 0)
can be shown to be true.
Note that if the proposition ∀x∃yP (x, y) is true, it means that the value of y depends
on x, i.e. for any x, there exists a y (we can always find y whose value depends on x)
such that P (x, y) is true.
Next, consider the other proposition “There exists a number, say x, such that for any
number y, x + y = y”. This proposition is symbolized by “∃x∀y (x + y = y)”, and it is
true if and only if the propositional form “∀y (x + y = y)” generates one true proposition.
Well, it is true when x is substituted by 0. Hence, “∃x∀y (x + y = y)” is true.
Again, if the proposition ∃x∀yP (x, y) is true, then it signifies that there is a fixed
x (x may not be unique and the value of x does not depend on y) such that for any y,
P (x, y) is true.
Example 31. Assume U = R. Translate into English the following statements:
1. ∀x∀y (x + y = y + x)
2. ∀x∀y ((x > 0) ∧ (y < 0) → (xy < 0))
Example 32. Translate the following statements into English, where C (x) is “x has a
computer”, F (x, y) is “x and y are friends”, and the universe of discourse for all variables
x, y and z consists of all students in UP.

16
1. ∀x (C (x) ∨ ∃y (C (y) ∧ F (x, y)))

2. ∃x∀y∀z ((F (x, y) ∧ F (x, z) ∧ (y 6= z)) → ¬F (y, z))

Example 33. Express the following as a logical expression involving predicates, quanti-
fiers with a domain consisting of all people, and logical connectives.

1. “If a person is female and is a parent, then this person is someone’s mother.”

2. “Everyone has exactly one best friend.”

Example 34. Translate the statement “The sum of two positive integers is positive”
into a logical expression.

Example 35. Express the definition of a limit using quantifiers.

Definition 6. A proposition in the predicate calculus that is true in all universes is called
a validity.

The following validities are useful in mathematical proofs:

V1 ∃y∀xP (x, y) → ∀x∃yP (x, y)

V2 ¬∀xP (x) ↔ ∃x¬P (x)

V3 ¬∃xP (x) ↔ ∀x¬P (x)

V4 ∀x (P (x) ∧ Q (x)) ↔ ∀xP (x) ∧ ∀xQ (x)

V5 ∃x (P (x) ∨ Q (x)) ↔ ∃xP (x) ∨ ∃xQ (x)

V6 ∃x (P (x) ∧ Q (x)) → ∃xP (x) ∧ ∃xQ (x)

V7 ∀xP (x) ∨ ∀xQ (x) → ∀x (P (x) ∨ Q (x))

17
1.5 Rules of Inference
In this section, we introduce proof techniques and apply them to proving statements from
a collection of premises. We restrict ourselves to statements in propositional logic so that
we can better concentrate on the form that a correct logical argument must take.
An argument is a list of statements called premises followed by a statement called
the conclusion. (We allow the list of premises to be empty.) We say that an argument
is valid if the conjunction of its premises implies its conclusion. In other words, validity
means that if all the premises are true, then so is the conclusion. Validity of an argument
does not guarantee the truth of its premises, and so does not guarantee the truth of its
conclusion. It only guarantees that the conclusion will be true if the premises are.
We could construct a truth table. But this could become very cumbersome as the truth
table can get quite large. Secondly, checking the validity of an argument mechanically
by constructing a truth table is almost completely unenlightening as it gives you no good
idea why an argument is valid.
Hence, we’ll concentrate on an alternative way of showing that an argument is valid,
called a proof, that is far more interesting and tells you much more about what is going
on in the argument.
Hence, a proof of an argument is a list of statements, usually beginning with the
premises, in which each statement that is not a premise must be true if the statements
preceding it are true. In particular, the truth of the last statement, the conclusion,
must follow from the truth of the first statements, the premises. The conclusion of an
argument is that statement which asserts something while a premise serves as evidence
for the assertion. How do we know that each statement follows from the preceding ones?
We cite a rule of inference that guarantees that it is so.
A proof then, is a step-by-step demonstration that a statement (conclusion) can be
derived from a collection of premises. A premise is a statement that is assumed in the
context of a proof. Each step of a proof is either a premise or can be shown to be a
consequence of previous steps using certain rules of inference.
For example,
If I love math, then I will pass this course.
I love math.
Therefore, I will pass this course.
The statement that asserts a fact is “I will pass this course.” This is the conclusion.
The two other statements “If I love math, then I will pass this course,” and “I love math”
support the conclusion. These are the premises.
Suppose we make the following representations concerning the argument above.
p: I love math.
q: I will pass this course.
Then the argument can be written in this form:

p→q Premise (If I love math, then I will pass this course.)
p Premise (I love math.)
∴q Conclusion (Therefore, I will pass this course.)
The premises are separated from the conclusion by a horizontal bar. Any argument
written in the above form is said to be in standard form. Thus, the premises and the
argument conclusion are easily identified when an argument is written in standard form.

18
Example 36. Using symbols, write in standard form the argument stated below.
If two sides of a triangle are equal, then the opposite angles are equal. Two sides of
a triangle are not equal. Therefore, the opposite angles are not equal.
Practice Problem 12. Translate the following argument in symbols and write in standard
form.
If a man is a bachelor, he is unhappy. If a man is unhappy, he dies young. Therefore,
bachelors die young.
So far, we have learned only how to write arguments in standard form without ques-
tioning the validity of these arguments. Now, we shall study how to determine the validity
of arguments.
An argument is said to be valid if and only if the conclusion is true whenever the
premises are assumed to be true. Otherwise, the argument is called a fallacy.
Example 37. Determine whether the argument given in Example 36 is valid or not.
By the definition of a valid argument, the premises can only be assigned the truth
value T and from these assignments, the conclusion must have the truth value T only.
Otherwise, the argument is a fallacy. Observe that in row 3, the premises p → q and
¬p are both true whereas ¬q is false. This means that the premises are true but the
conclusion is false. In fact, we have shown a counterexample. Hence, the argument is
declared a fallacy.
Note that only rows 3 and 4 are used in determining the validity of the argument
because it is in these rows where the premises are both true. Thus, as a general rule in
determining the validity of any argument, observe only the truth value of the conclusion
from the rows where the premises are all assigned the truth value T. If in these rows, the
conclusion only has the truth value T, then the argument is valid.

If an argument is known to be valid, its form can be used to check the validity of
other arguments. In what follows, we introduce some rules of inference and illustrate
how each of them is used to construct proofs.

1.5.1 Modus Ponens or Direct Reasoning


In words: If p implies q, and if p is true, then q must be true.
Modus ponens is based on the tautology [p ∧ (p → q)] → q.
Example 38. Premises: r ∧ t, (r ∧ t) → s. Prove s.
Solution
1. r∧t Premise
2. (r ∧ t) → s Premise
3. ∴s Conclusion (1, 2, Modus Ponens)

Practice Problem 13. Premises: t → r, r → s, t. Prove s.


Example 39. Premises: r ∨ t → (¬r → t), r ∨ t, ¬r. Prove t.
Practice Problem 14. Premises: ¬ (r ∧ t) → ¬r ∨ ¬t, ¬ (r ∧ t), ¬r, ¬r ∨ ¬t → (¬r → t).
Prove t.

19
Notice that in a proof, you do not have to use all of the premises.
Word of Caution: Modus Ponens tells us that, if p → q appears on the list, and if
p also appears on the list, then we can add q to the list of true statements. If p → q
appears on the list, but if p does not appear on the list, then we cannot add q to the list.
Put another way, if p implies q is true, then we cannot conclude that q is true until we
know that p is true.

1.5.2 Modus Tollens or Indirect Reasoning


In words: If p implies q, and q is false, then so is p.
Modus tollens is based on the tautology [(p → q) ∧ ¬q] → ¬p.

Example 40. Once again, we take p: “I love math” and q: “I will pass this course,” we
get
If I love math then I will pass this course; but I know that I will fail it. Therefore, I
must not love math.
In symbols, we have

p→q Premise (If I love math, then I will pass this course.)
¬q Premise (I will not pass this course.)
∴ ¬p Conclusion (Therefore, I do not love math.)
Practice Problem 15. Use either modus ponens or modus tollens to draw the appropriate
conclusion.
If I go to Mars, I will run for office.
I am going to Mars.
Therefore,

If I go to Mars, I will run for office.


I am going to run for office.
Therefore,

If I go to Mars, I will run for office.


I am not going to Mars.
Therefore,

If I go to Mars, I will run for office.


I will not run for office.
Therefore,

1.5.3 Simplification
In words: If both p and q are true, then, in particular, p is true.
Simplification is based on the tautology (p ∧ q) → p and (p ∧ q) → q.
The rule of simplification states that whenever a conjunction is true, then we can
deduce either of its conjuncts. That is, if a conjunction is given to be true, then either
conjunct must be true. For instance, if the statement
“Carla runs away from responsibilities, yet she is not a coward”
is accepted as true, then concluding the statement

20
“Carla is not a coward”
is correct. Or, if our conclusion is the other statement, that is
“Carla runs away from responsibilities”
then the conclusion is still correct.

1.5.4 Addition
In words: If p is true, then we know that either p or q is true.
Addition is based on the tautology p → (p ∨ q).

Example 41. If the sky is blue, then either the sky is blue or some ducks are kangaroos.
In symbols, we have

p Premise (The sky is blue.)


∴p∨q Conclusion (The sky is blue or some ducks are kangaroos.)

1.5.5 Conjunction or Adjunction


In words: If p and q are provable from the same set of premises, then p ∧ q
is provable from that set of premises. or If p and q are any two lines in a
proof, then we can add the line p ∧ q to the proof.
Conjunction is based on the tautology (r → p) ∧ (r → q) ↔ (r → p ∧ q).

Example 42. Premises: p → q, q ∧ (p ∨ r) → s ∧ t, p, p → p ∨ r. Prove s ∧ t.


Solution
1. p Premise
2. p→q Premise
3. q 1, 2, Modus Ponens
4. p→p∨r Premise
5. p∨r 1, 4, Modus Ponens
6. q ∧ (p ∨ r) 3, 5, Adjunction
7. q ∧ (p ∨ r) → s ∧ t Premise
8. ∴s∧t Conclusion (6, 7, Modus Ponens)

Practice Problem 16. Premises: q → p, q, p ∧ q → r ∨ t. Prove r ∨ t.

1.5.6 Absorption
The rule of absorption tells us that any proposition may be conjoined to both sides of
an implication.
In words: If p → q and r is true, then (p ∧ r) → (q ∧ r).
Absorption is based on the tautology (p → q) ∧ r → [(p ∧ r) → (q ∧ r)].

In general, a rule of inference is just an instruction for obtaining additional true


statements from a list of true statements. Think of them as tools for constructing new
statements from old ones; the more tools you have at your disposal, the easier your task
becomes.

21
Problem Set
Problem 1.5.1. Let p represent “sugar is white”, q represent “today is Friday” and r
represent “Talisay is a city”. Rewrite each of the following in words:

(a) ¬p ∧ ¬q

(b) ¬p ∨ q

(c) p ∧ (¬q ∧ ¬r)

(d) ¬ (¬q ∨ r)

(e) ¬p ∧ (q ∨ r)

Problem 1.5.2. Write down the negation of the following statements:

(a) Mr. Gates and Mr. Jordan are both rich.

(b) Mary is a smart girl or Rose wears a pair of glasses.

(c) x is greater than y.

(d) x is neither a prime number nor an even number.

Problem 1.5.3. Translate the following arguments in symbols and write in standard
form.

(a) 2 is even or 5 is not a prime number. But 5 is prime. So 2 is even.

(b) If I study, then I will not fail math. If I do not play basketball, then I will study. I
failed math. Hence, I played basketball.

Problem 1.5.4. Determine whether each of the arguments given in the preceding prob-
lem is valid or not.

Problem 1.5.5. Make correct conclusions for each of the following sets of statements.

(a) If it does not rain, I will go to the beach. I did not go to the beach.

(b) If rain continues and the river rises, then either the bridge is washed out or the
people make a temporary road. But the rain continued and the river rose.

(c) I studied hard, yet I failed the exam.

Problem 1.5.6. Prove each of the given statements from the given premises.

(a) Premises: p ∨ q → s ∧ t, p. Prove s ∧ t.

(b) Premises: (p ∨ q) → (r ∧ ¬s), ¬ (r ∧ ¬s), (p ∨ q) → p. Prove ¬p.

22
Chapter 2

Methods of Proof

A proof of a mathematical statement is a logical argument which establishes the truth of


the statement. The steps of the logical argument are provided by implications. One of
the main aims of this course is to describe a variety of methods of proof so that you can
follow these when you meet them and also construct proofs for yourself.
At this point, you may be quite familiar with some “proofs” and proving techniques.
As a Math student, especially at a Masteral level, you are expected to be able to construct
your own proofs and, as importantly, to write them out carefully so that other people can
understand them – or even so that you can understand them yourself when you come to
look back at your work some months later. One real difficulty is that we do not normally
discover proofs in the polished form in which they are presented. It is important to realize
that you will usually spend time constructing a proof before you then write out a formal
proof. You can think of this as erecting a sort of scaffolding for the purpose of constructing
the proof. When the proof has been constructed, the scaffolding is removed so that the
proof can be admired in all its economical beautiful simplicity! However, one difficulty
for the person encountering the proof for the first time is that it can be hard to make
sense of. To read and understand the proof, we may have to reconstruct the scaffolding
for ourselves from the formal proof. This can be difficult – but not usually as difficult as
thinking of the proof in the first place unless the proof is very badly written. This is a
problem not just for beginning undergraduates but also for professional mathematicians
when they read mathematics.
You may ask why then the scaffolding is not retained. The difficulty is that if every
detail is given then mathematical arguments become enormously long and cluttered. The
aim is to pitch your writing at the level of the expected reader so that there is just enough
information to enable the reconstruction of the scaffolding if necessary but not so much
that it would mask the essence of the argument. Too much detail can make a proof based
possibly on one simple idea appear enormously complicated. To avoid detail it is quite
common to use statements like “it is easy to show that <some statement> is true” or “it
now readily follows that <some statement> is true” and then the reader has to confirm
that this is indeed the case. It is, however, important not to use such phrases as a lazy
way of avoiding thinking about the details.
It is also the case that excessive pedantic precision can sometimes make mathematics
hard to read. Writing mathematics is not like writing a computer program; what is
written will be read by a human being who has much common experience with the writer
and so is able to anticipate to some extent what the writer intends. As Gila Hanna has
written,

23
The student of mathematics has to develop a tolerance for ambiguity. Pedantry can be
the enemy of insight.
Putting this into practice is a matter of fine judgement: ambiguous statements are only
acceptable in contexts which resolve the ambiguity almost immediately. While learning
to write good mathematics it is probably better to err on the side of pedantry.
As novices, arguments will often be presented first with lots of scaffolding (as ‘con-
structing a proof’) and then with the scaffolding removed (as ‘(formal) proofs’). You
should ensure that in each case you do understand why the ‘proof ’ does prove
the result as claimed. When you read most mathematics books you need to work
with pencil in hand reconstructing the detail behind the proofs provided. You cannot
normally read a mathematics book like a conventional novel. (Eccles 1997)

2.1 Direct Proof


Let us begin by thinking about one of the simplest forms of mathematical result. Very
many theorems are of the form P ⇒ Q. How do we set about proving such a statement?
Since the statement is necessarily true (i.e. vacuously true) if P is false, we only need
consider the case when P is true. Then from the truth table, we see that P ⇒ Q is true
so long as Q is also true.
So to prove that P ⇒ Q is true, it is sufficient to assume that P is true and deduce
Q. This is the direct form of proof. Here is an example.
Proposition: For positive real numbers a and b, a < b ⇒ a2 < b2 .
Constructing a proof.
Proof. Given positive real numbers a and b suppose that a < b.
Then a2 < ab (multiplying both sides by a > 0)
and ab < b2 (multiplying both sides by b > 0).
Hence a2 < b2 . (transitivity)
It follows that a < b ⇒ a2 < b2 .
Notice that this proof is written as a sequence of sentences with words like ‘then’,
‘hence’ and ‘it follows’ indicating how the sentences are related. Of course, some symbols
are used but it is a good practice when writing mathematics to read it out aloud to check
that when the symbols are converted into words (‘a is less than b’ in the first sentence
of the above proof for example) what you have written is a sensible piece of prose. We
do sometimes use more symbols so that the above proof might have been written out as
follows.
Proof. For positive integers a and b, a < b ⇒ (a2 < ab and ab < b2 ) ⇒ a2 < b2 . Hence
a < b ⇒ a2 < b2 .

2.1.1 Techniques
(a) propositions having no hypothesis

• Prove that for all sets A and B, A ⊆ A ∪ B. F

24
Proof. Let A and B be arbitrary sets. To prove A ⊆ A ∪ B, let x be an
arbitrarily chosen element of A. [Note: We are assuming that x ∈ A.] We
must prove that x ∈ A ∪ B. By the definition of “union,” this means we must
prove that either x ∈ A or x ∈ B. Since we know x ∈ A, by our assumption,
the desired condition x ∈ A or x ∈ B follows immediately.

• Prove that for all real numbers x, x < x + 1.

(b) propositions having one or more hypotheses

• For all sets A, B, and C, if A ⊆ B, then A ∩ C ⊆ B ∩ C.


• If x ≤ y and u ≤ v, then x + u ≤ y + v.

(c) division into cases

• Prove that |xy| = |x| · |y| for all real numbers x and y.

(d) proving equality of sets


To prove A = B, we have to prove that A ⊆ B and B ⊆ A.

(e) proving the biconditional or proof of equivalence


To prove P ⇔ Q, we have to prove that P ⇒ Q and Q ⇒ P .

(f) proving equivalent statements


To show that these statements are equivalent:

P1 , P2 , . . . , Pn

we must prove the following implications: P1 ⇒ P2 , P2 ⇒ P3 , . . . , Pn−1 ⇒ Pn , and Pn ⇒


P1 . The order may vary but the loop must be satisfied.

(g) choose method


From the proof of F above, our starting point “... assume x ∈ A ...” is an
application of one of the most widely-used approaches to proof-writing, known
as the choose method. The basic approach to deriving a conclusion of the form
∀x [P (x) → Q (x)], is to begin by choosing an arbitrary object (giving it a
specific name such as “x”) for which it is assumed that P (x) is true. Our goal is
to deduce that Q (x) must consequently also be true.
The object x is a fixed, but arbitrarily chosen, element of the universe of discourse
of P (x) and Q (x). We do not assign any specific value to x; rather we give the
name “x” to a generic object [that is assumed to satisfy the propositional function
P (x)] and use that name to keep track of the object as we proceed through the
steps of the proof.
The power of this approach is that any conclusion we draw about “this x” applies
to every object a for which the assumption P (a) is true. This is valid by the rule
of universal generalization.

25
Universal instantiation is the rule of inference used to conclude that P (c) is
true, where c is a particular member of the universe of discourse, given the
premise ∀xP (x).
example: “All women are wise.” Lisa is a member of the universe of discourse
of all women. Therefore, “Lisa is wise”.

∀xP (x)
∴ P (c)
Universal generalization is the rule of inference that states that ∀xP (x) is
true, given the premise that P (c) is true for all elements c in the universe of
discourse. Universal generalization is used when we show that ∀xP (x) is true
by taking an arbitrary element c from the universe of discourse and showing
that P (c) is true. The element c that we select must be an arbitrary, and
not a specific, element of the universe of discourse. choose method Universal
generalization is used implicitly in many proofs in mathematics and is seldom
mentioned explicitly.

P (c) for an arbitrary c


∴ ∀xP (x)
Existential instantiation is the rule that allows us to conclude that there is an
element c in the universe of discourse for which P (c) is true if we know that
∃xP (x) is true. We cannot select an arbitrary value of c here, but rather it
must be a c for which P (c) is true. Usually we have no knowledge of what c
is, only that it exists. Since it exists, we may give it a name “c” and continue
our argument.

∃xP (x)
∴ P (c) for some element c
Existential generalization is the rule of inference that is used to conclude that
∃xP (x) is true when a particular element c with P (c) true is known. That is,
if we know one element c in the universe of discourse for which P (c) is true,
then we know that ∃xP (x) is true.

P (c) for some element c


∴ ∃xP (x)

(h) constructing proofs backwards

• ε − δ limit proofs in Calculus


√ (x + y)
• Prove that xy ≤ .
2
1
• Prove that x + ≥ 2.
x
(i) proving ‘or’ statements

• Prove: If a and b are real numbers, then ab = 0 if and only if a = 0 or b = 0.

26
Proof. (⇒) Suppose that a = 0 or b = 0.
Then, ab = 0 · b = 0 or ab = a · 0 = 0.
Therefore, ab = 0.
(⇐) (Converse)
Suppose that a and b are real numbers such that ab = 0.
We need to prove that a = 0 or b = 0. Remembering the Switcheroo, this is
equivalent to the statement “if a 6= 0, then b = 0.”
So, suppose that a 6= 0.
Then, dividing ab = 0 by a gives us b = 0, as required. (again, this is possible
because a 6= 0)
Hence, a = 0 or b = 0.

We can also try proving this by cases.

27
2.2 Indirect Proof
2.2.1 Proof by Contrapositive
P ⇒ Q ≡ ¬Q ⇒ P

• Prove: If n2 is even, then n is even.

• Let m and n be nonnegative integers. Prove that if m + n > 50, then m > 25 or
n > 25.

2.2.2 Proof by Contradiction


(or reductio ad absurdum)

¬s is always false ¬s ⇒ (r ∧ ¬r)


∴ s is always true ∴ s is always true

• Prove that 2 is an irrational number.

• Prove that there do not exist integers m and n such that

14m + 20n = 101.

Using our rules for quantifiers, we can rewrite this statement as “For all integers
m and n, 14m + 20n 6= 101. Note first that it is difficult to prove this statement
directly for we clearly cannot consider ALL the possibilities for m and n one at a
time. We now proceed with the proof by contradiction.

Proof. Suppose for contradiction that there exist integers m and n such that 14m +
20n = 101.
101 = 14m + 20n = 2 (7m + 10n) is even. contradiction! because 101 is odd.
Hence, our assumption is false, and we conclude that such integers m and n cannot
exist.

Remark: One never assumes the negation of the hypothesis in either a proof of an
implication by contradiction or by contrapositive.

2.3 Proof of Existence


see comments above on Existential Generalization.

2.3.1 Constructive
• Every equation ax + b = 0, where a and b are real numbers and a 6= 0, has a real
number solution.

28
2.3.2 Nonconstructive
One common method of giving a nonconstructive existence proof is to use proof by
contradiction. Many nonconstructive proofs assume the non-existence of the thing whose
existence is required to be proven, and deduce a contradiction.

• Prove that there is a prime number > 3.

2.4 Proof of Uniqueness


Some theorems assert the existence of a unique element with a particular property. In
other words, these theorems assert that there is exactly one element with this property.
To prove a statement of this type we need to show that an element with this property
exists and that no other element has this property. The two parts of a uniqueness proof
are:

Existence: We show that an element x with the desired property exists.

Uniqueness: We show that if y 6= x, then y does not have the desired property.
A common technique in proving uniqueness is to assume that two elements x and
y satisfy the desired property, and then show that x = y.

• Every equation ax+b = 0, where a and b are real numbers and a 6= 0, has a unique
real number solution.

2.5 Disproving an Argument: Counterexample


• Prove or disprove: Every prime integer is odd.

Remark: The introduction to the methods of proofs presented here is far from complete.
What we presented are just the common techniques and rules of logic used in proving.
There is no definite procedure in writing out a proof. The only way to be proficient in
proving theorems is to follow the given proofs of as many theorems as possible and then
use these as patterns to prove similar theorems.

29
2.6 MATHEMATICAL INDUCTION
For our purpose in mathematics, we need a proof by induction that is guaranteed as valid.
To have a basis for such method of proof, we shall utilize the following fact about the set
of positive integers known as the Well-Ordering Principle.
Every non-empty set S of non-negative integers contains a least element; that is, there
is some integer a in S such that a ≤ b for all b in S.

2.6.1 The Induction Principle


Mathematical induction is used to prove propositions of the form ∀nP (n), where the
universe of discourse is the set of positive integers.
A proof by mathematical induction that P (n) is true for every positive integer n
consists of two steps.

BASIS STEP: Prove P (1).

INDUCTIVE STEP: Prove the implication P (k) ⇒ P (k + 1) for all positive integers
k.

Here, the statement P (k) for a fixed positive integer k is called the inductive hy-
pothesis or IH. When we complete both steps of a proof by mathematical induction,
we have proved that P (n) is true for all positive integers n; that is, we have shown that
∀nP (n) is true.
For the purpose of a proof by mathematical induction, we suggest the following for-
mat:
Statement of the Theorem:
In symbols:
Proof. I. Verification or basis step.
Let n = 1:
Let n = 2:

II. Inductive step.


Assume the statement is true for n = k, where k is a natural number.
That is, (inductive hypothesis/IH).
To prove: The statement is true for n = k + 1.
That is, . ?
To do this,
Proof:
<write the proof of ? here>

III. Conclusion.
By the principle of Mathematical Induction, is true for
all natural numbers or all positive integers n.

30
• Use mathematical induction to prove that the sum of the first n odd positive integers
is n2 .

• Use mathematical induction to prove the inequality

n < 2n .

2.6.2 Strong Induction


There is another form of mathematical induction that is often useful in proofs. With this
form, we use the same basis step as before, but we use a different inductive step. We
assume that P (j) is true for j = 1, 2, . . . , k and show that P (k + 1) must also be true
based on the assumption. This is called strong induction (also known as the second
principle of mathematical induction). Here are the two steps.

BASIS STEP: The proposition P (1) is shown to be true.

INDUCTIVE STEP: It is shown that [P (1) ∧ P (2) ∧ · · · ∧ P (k)] ⇒ P (k + 1) is true


for all positive integers k.

The two forms of mathematical induction are equivalent; that is, each can be shown
to be a valid proof technique assuming the other. Although, the strong induction prin-
ciple differs from the first in that it requires trying to show P (k + 1) is true when
P (1) , P (2) , . . . , P (k) are all true; that is, the inductive hypothesis says P (n) is true
not only for n = k, but for all positive integers preceding k or less than k.

• Prove that for every positive integer n, if n ≥ 14, then n can be expressed as a sum
of 3’s and/or 8’s.

31

You might also like