Theory of Numbers
Theory of Numbers
1 Theory of Divisibility 3
1.1 Divisibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Greatest Common Divisor and Least Common Multiple . . . . . . . . . . 5
1.3 Factorization in Prime Numbers . . . . . . . . . . . . . . . . . . . . . . . 8
3 Congruences 23
3.1 Congruences and Its Properties . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Fermat’s Theorem, Euler’s Theorem and Wilson’s Theorem . . . . . . . . 26
4 Solutions of Congruences 31
4.1 Linear Congruences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
5 Quadratic Residuacity 36
5.1 Quadratic Residue and Nonresidues . . . . . . . . . . . . . . . . . . . . . 36
5.2 Legendre Symbol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
1
9 Limit and Continuity of a Function 56
9.1 Limit of a Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
9.2 Continuous Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
10 Differentiation 61
11 Riemann Integration 64
A Reviews 66
2
Chapter 1
Theory of Divisibility
1.1 Divisibility
Definition 1.1.1 (Natural Number). A number used for counting is said to be a nat-
ural number. The set of natural numbers is denoted and defined by N = {1, 2, 3, 4, 5, · · · }.
Definition 1.1.2 (Integer). Any whole number positive, negative, or zero is said to be an
integer. The set of integers is denoted and defined by Z = {· · · , −3, −2, −1, 0, 1, 2, 3, · · · }.
Definition 1.1.3 (Divisibility). A number b is said to be divisible by a number a(a 6= 0)
if there is a number c such that b = ac. We denote this by a | b, read as a divides b. If
no such integer can be found we write a - b, read as a does not divide b.
Example 1.1.1. (1) 3 | 12 because 12 = 3 · 4
0 < a < b, then a is called a proper divisor or factor of b. We have b = a · c = (−a) · (−c).
Thus if a | b, then −a | b. For practical purposes we can limit our attention to only
(ii) a | b and b | c ⇒ a | c
(iii) a | b and b | a ⇒ a = ±b
(iv) a | 1 ⇒ a = ±1
3
(v) a | b and a | c ⇒ a | (bx + cy) for any integers x and y.
The sum, difference, and product of two integers are obviously integers but the quotient
a = qb + r, 0 ≤ r < |b|.
Proof. The integer a lies between two consecutive integers of the sequence
So WLOG we may assume that q|b| ≤ a < (q + 1)|b|. Then a − q|b| ≥ 0 and a − q|b| < |b|.
a = qb + r when b > 0
⇒ |q − q1 ||b| = |r1 − r|
⇒ |q − q1 | < 1
Since q and q1 are both integers, so the above relation hold if q − q1 = 0 ⇒ q = q1 .
Consequently, 0 = r1 − r ⇒ r = r1 .
Definition 1.1.5 (Prime number). Any integer p which exceeds unity is called a prime
if it has no integral divisors except ±p and ±1.
4
Definition 1.1.6 (Composite number). An integer c > 1 which has divisors other than
±c and ±1 is called a composite number.
For example (20, 30) = 10, (−6, 9) = 3. Note that (a, b) ≥ 1, (a, 0) = |a|, (0, 0) = ∞ and,
Definition 1.2.2 (Relatively prime). If (a, b) = 1 then the integers a and b are called
relatively prime or co-prime.
Theorem 1.2.1 ( Euclidean Algorithm(330-275 B.C.)). Let a and b be two positive inte-
gers where b - a. Let r0 = a, r1 = b and apply the division algorithm repeatedly to obtain
a set of remainders r2 , r3 , · · · , rn , rn+1 defined successively by the relations:
r 0 = q1 r 1 + r 2 , 0 < r 2 < r 1
r 1 = q2 r 2 + r 3 , 0 < r 3 < r 2
r 2 = q2 r 3 + r 4 , 0 < r 4 < r 3
··· ··· ··· ··· ···
rn−2 = qn−1 rn−1 + rn , 0 < rn < rn−1
rn−1 = qn rn + rn+1 , where rn+1 = 0.
Then rn , the last nonzero remainder in this process, is (a, b), the g.c.d. of a and b.
which rn+1 = 0. The last relation rn−1 = qn rn shows that rn | rn−1 and hence the
5
Hence rn is the required g.c.d. of a and b.
The following theorem states that the g.c.d. of any two integers can be expressed as
Theorem 1.2.2. If g = (a, b), then there exists integers x and y such that g = ax + by.
Proof. Consider the linear combinations au + bv, where u, v ∈ Z. This set of integers
includes positive and negative values, and 0 by the choice of u = v = 0. Choose x and y
Now we prove that l | a and l | b. Suppose that l - a. Then there exist integers q and
r such that
a = ql + r, 0 < r < l.
Thus r is in the set {au + bv}. This contradicts the fact that l is the least positive integer
Now, since g is the g.c.d. of a and b, we may write a = gA, b = gB, and l = ax + by =
Example 1.2.1. 1. Find the g.c.d. of 1769 and 2378 using Euclidean algorithm
6
2. Now
29 = 551 − 9 · 58
= 551 − 9 · (609 − 1 · 551)
= 10 · 551 − 9 · 609
= 10(1769 − 2 · 609) − 9 · 609
= 10 · 1769 − 29 · 609
= 10 · 1769 − 29 · (2378 − 1 · 1769)
= 39 · 1769 + (−29) · 2378
∴ 29 = 39 · 1769 + (−29) · 2378.
Definition 1.2.3 (Least Common Multiple). The least common multiple(l.c.m. ) of two
or more non zero integers is the smallest positive integer that is divisible by all of them.
If l is the l.c.m. of a and b, we denote it by [a, b] = l.
For example [20, 30] = 60. Note that there is no greatest common multiple of two or
more integers. Also observe that g.c.d. and l.c.m. are unique.
∴ ab | (la, lb)
⇒ ab | l(a, b)
⇒ ab | lg · · · (1)
ab ab ab
On the other hand, since a | g
and b | g
, so g
is a common multiple of a and b. Therefore,
ab
l| g
⇒ lg | ab · · · (2) From (1) and (2) we conclude that lg = ab ⇒ [a, b](a, b) = ab.
¯
¯
Note. [a, b]¯abm where m is an integer,that is, l.c.m. of any two or more integers divides
any common multiple of them.
Theorem 1.2.5. Let p be a prime and a be any integer. Then either p | a or (a, p) = 1.
7
Proof. If p | a, then there is nothing to prove. If p - a, then (a, p) = 1. Since p | ab and
(a, p) = 1, so p | b.
ax + by = 1.
Since a | bc ⇒ bc = am, m ∈ Z.
= acx + amy
= a(cx + my)
Hence a | c
Proof. Let p1 > 1 be the least divisor of N . Then p1 < N . Evidently, p1 is a prime.
proceed with N2 as before. After a finite number of such steps we obtain a factorization
N = p1 p2 p3 · · · pn
N = q1 q2 q3 · · · qr
is a second factorization of N .
8
Then the prime q1 evidently divides one of the primes pi , say p1 . Hence q1 = p1 and
Corollary 1.3.2. An integer n > 1 can be uniquely written as n = pα1 1 pα2 2 pα3 3 · · · pαk k
where p1 , p2 , p3 , · · · , pk are distinct primes and α1 , α2 , α3 , · · · , αk are positive integers.
n = p1 p2 p3 · · · pr
Since the prime factors are not necessarily distinct, by collecting the same primes we can
Motivation:. N = 2 · 3 · 5 = 30
P = N + 1 = 30 + 1 = 31, not divisible by the prime in the composition of P .
Proof. Let there are only a finite number of primes, say p1 , p2 , p3 , · · · , pk , where pk is
N = p1 p2 p3 · · · pk
⇒N + 1 = p1 p2 p3 · · · pk + 1
⇒P = p1 p2 p3 · · · pk + 1, where P = N + 1.
Now if P is a prime then pk is not the largest prime. If P is not a prime, then by
9
Since there are only prime p1 , p2 , p3 , · · · , pk , so p will be one of them, say p = pi , 1 ≤
Example 1.3.1. Show that one of the two consecutive integers is divisible by 2.
Solution: Let the two consecutive integers be n, n + 1. If n be odd, then n + 1 is even
and is divisible by 2. If n is even then it is divisible by 2.
Example 1.3.2. Show that one of the three consecutive integers is divisible by 3.
Solution: Let n, n + 1, n + 2 be any three consecutive integers. Here n must be any
one of the forms 3m, 3m + 1, 3m + 2. When n = 3m, the first integer is divisible by 3,
when n = 3m+1, n+2 = 3m+3 which is divisible by 3, when n = 3m+2, n+1 = 3m+3
which is divisible by 3.
Example 1.3.3. Show that the product of any three consecutive integers is divisible by
6.
Solution: Let n be any integers. Then n(n + 1)(n + 2) is divisible by 6.
a5 −
1 =
Example 1.3.4. If 2p − 1 is a prime, then show that p is itself a prime.
(a−1).
Solution: Suppose that p is not a prime. Then it is composite, say p = ab. Then
(a4 +
2p − 1 = 2ab − 1 = (2a )b − 1 which is divisible by 2a − 1. So, 2ab − 1 or 2p − 1 is a composite 3
a + 1)
number which contradicts our given assumption.
So p must be a prime.
10
Problem Plus 1
2.
11
Chapter 2
12
Definition 2.1.6. The function
X
σk : N → N defined by σk (n) = dk
d|n
d≥1
denotes the sum of the k-th powers of the divisors of n, where n is a positive integer.
Remark. If n is a prime, then n has only two positive divisors, 1 and n itself.
Definition 2.1.7 (Euler’s φ function). Let n ≥ 1 be a positive integer. Then φ(n) denotes
the number of numbers not exceeding n, which are relatively prime to n. This φ(n) is
called Euler’s φ function.
Example 2.1.3. Following table lists τ (n), σ(n) and P (n) for n = 1, 2, 3, · · · , 10.
n 1 2 3 4 5 6 7 8 9 10
τ (n) 1 2 2 3 2 4 2 4 3 4
σ(n) 1 3 4 7 6 12 8 15 13 18
P (n) 1 2 3 8 5 36 7 64 27 100
φ(n) 1 1 2 2 4 2 6 4 6 4
Here [x] and a are respectively called the integral and fractional part of x.
Theorem 2.1.1 (Goldbach Conjecture). This states that every even integer bigger than
2 is the sum of two primes.
13
Example 2.1.5.
4=2+2
6=3+3
8=3+5
10 = 3 + 7
50 = 3 + 47
100 = 29 + 71
It is an unsolved problem. Goldbach verified up to 1,00,000 at least.
Now let us provide the general formulae for τ and σ when n is a composite number
as follows:
Theorem 2.1.2. If n = pα1 1 pα2 2 pα3 3 · · · pαk k where p1 , p2 , p3 , · · · , pk are distinct primes and
α1 , α2 , α3 , · · · , αk are positive integers, then
k
Y
(i) τ (n) = (α1 + 1)(α2 + 1) · · · (αk + 1) = (αi + 1)
i=1
k
pα1 +1 − 1 pα2 2 +1 − 1 pαk +1 − 1 Y pαi i +1 − 1
(ii) σ(n) = 1 · ··· k =
p1 − 1 p2 − 1 pk − 1 i=1
pi − 1
Proof. We shall prove the theorem by induction on k. Let k = 1, that is, n has only
one prime factor, say n = pα , where α is a positive integer. Then the positive divisors of
n are 1, p, p2 , · · · , pα . We have
pα+1 − 1
τ (n) = α + 1 and σ(n) = 1 + p + p2 + · · · + pα =
p−1
Next suppose that (i) and (ii) hold when n has k − 1 distinct prime factors(k ≥ 2).
α
Write n = n0 pαk k where n0 = pα1 1 pα2 2 pα3 3 · · · pk−1
k−1
, (n0 , pαk k ) = 1, and all p’s are distinct
primes. Moreover, any divisor of n0 is a divisor of n. Hence any divisor of n is of the form
d0 ptk , 0 ≤ t ≤ αk , d0 ≥ 1, d0 | n0 .
X
Now τ (n) = 1
d|n
d≥1
X X X X
= 1+ 1+ 1 + ··· + 1
¯
d0 pk k ¯n
d0 |n d0 pk |n d0 p2k |n α
14
By induction hypothesis we have τ (n0 ) = (α1 + 1)(α2 + 1) · · · (αk−1 + 1)
k
Y
∴ (i) τ (n) = (α1 + 1)(α2 + 1) · · · (αk−1 + 1)(αk + 1) = (αi + 1)
i=1
X
Again σ(n) = d
d|n
d≥1
X X X
= d0 + d0 pk + · · · + d0 pαk k
d0 |n0 d0 |n0 d0 |n0
d0 ≥1 d0 ≥1 d0 ≥1
X 0 X 0 X 0 αk
= d+ d pk + · · · + d pk
d0 |n0 d0 |n0 d0 |n0
d0 ≥1 d0 ≥1 d0 ≥1
X 0
= d (1 + pk + · · · + pαk k )
d0 |n0
d0 ≥1
pαk k +1 − 1
0
= σ(n )
pk − 1
By induction hypothesis
α +1
pα1 1 +1 − 1 pα2 2 +1 − 1 p k−1 − 1
σ(n0 ) = · · · · k−1
p1 − 1 p2 − 1 pk−1 − 1
α +1 k
pα1 1 +1 − 1 pα2 2 +1 − 1 p k−1 − 1 pαk k +1 − 1 Y pαi i +1 − 1
∴ (ii) σ(n) = · · · · k−1 · =
p1 − 1 p2 − 1 pk−1 − 1 pk − 1 i=1
pi − 1
k k k−1
Theorem
¡ ¢ 2.1.3. If p is a prime and k is a positive integer, then φ(p ) = p − p =
k 1
p 1− p .
Proof. The positive integers less than or equal to pk which are not relatively prime to pk
are multiples of p, i.e., these are p, 2p, 3p, · · · , pk−1 p. Hence deleting these pk−1 numbers
φ(pk ) = pk − pk−1 ,
15
Corollary 2.1.4. If p is a prime, then φ(p) = p − 1.
Theorem 2.1.5. ( Prof. Dr. Md. Fazlur Rahman [?]) The number theoretic functions
τ, σ and, φ are multiplicative.
Theorem 2.1.6. If a and b are relatively prime positive integers, then φ(ab) = φ(a)φ(b).
a − columns
1 k=2 3 → φ(a) = 2
b − rows 4 5 6
7 8 9
10 11 12
↓
φ(b) = 2
Proof. Considering the product ab, we write the first ab integers in b rows and a
1 2 3 4 ··· k ··· a
a+1 a+2 a+3 a+4 ··· a+k ··· 2a
2a + 1 2a + 2 2a + 3 2a + 4 ··· 2a + k ··· 3a
3a + 1 3a + 2 3a + 3 3a + 4 ··· 3a + k ··· 4a
··· ··· ··· ··· ··· ··· ··· ···
(b − 1)a + 1 (b − 1)a + 2 (b − 1)a + 3 (b − 1)a + 4 ··· 3(b − 1)a + k ··· ba
Now considering the k-th column we find that if (k, a) = 1, then every number in k-th
column is prime to a.
Now the first row contains φ(a) integers prime to a and therefore there are φ(a)
columns in each of which every term is prime to a. Let k-th column be such one. Clearly
this column is in A.P.(Arithmetic Progression). So, the terms of this column when divided
by b will leave the remainders 0, 1, 2, · · · , (b − 1). Hence this column contains φ(b) terms
prime to b.
Thus, there are φ(a) columns of numbers prime to a and each column contains φ(b)
numbers prime to b.
16
Hence in the above arrangements of ab numbers there are φ(a)φ(b) numbers which
are prime to both a and b and so to the product ab. Therefore φ(ab) = φ(a)φ(b). So the
theorem is established.
Corollary 2.1.7. If a, b, c, · · · , l are prime to each other, then φ(abc · · · l) = φ(a)φ(b)φ(c) · · · φ(l).
Corollary 2.1.9. If n = pα1 1 pα2 2 pα3 3 · · · pαk k where pi ’s are relatively prime in pairs, then
φ(n) = n(1 − p11 )(1 − p12 )(1 − p13 ) · · · (1 − p1k ).
µ : N → {−1, 0, 1}
In other words,
1 if n = 1
µ(n) = 0 if n has a square factor or a2 | n for some a ∈ Z
(−1)k if n = p1 p2 · · · pk where p1 , p2 , · · · , pk are distinct prime factors
Example 2.1.7.
X X
µ(d) = µ(d) = µ(1) = 1.
d|n d|1
17
where pi are distinct primes, αi ≥ 1, i = 1, 2, 3, · · · , r.
X
Therefore µ(d) = µ(1) + [µ(p1 ) + µ(p2 ) + · · · + µ(pr )]
d|n
+ · · · + µ(p1 p2 · · · pr )
square factors)
Xr r
X
= µ(1) + µ(pi ) + µ(pi pj ) + · · · + µ(p1 p2 · · · pr )
i=1 i<j
i,j=1
= (1 − 1)r
=0
(
X 1 if n=1
∴ µ(d) =
d|n
0 if n>1
has a square factor, so µ(a) = 0, µ(a)µ(b) = 0 and µ(ab) = 0 and hence µ(ab) = µ(a)µ(b).
and b are distinct. Therefore µ(a) = (−1)k , µ(b) = (−1)l and µ(ab) = (−1)k+l . Now
Note. The mobius function µ is not completely multiplicative since µ(4) = µ(22 ) = 0 but
µ(2) · µ(2) = (−1)2 = 1.
18
2.2 Diophantine Equations
An equation which has more than one unknowns is called an indeterminate equation. A
system of equations is called indeterminate if the number of equations is less than that
of the unknowns. Indeterminate equations are called Diophantine equations, after the
name of Greek mathematician Diophantus about 200 A.D., when we seek their solutions
in integers. In this section we shall only consider the linear diophantine equations in two
The following theorem gives the information when the linear diophantine equation
Theorem 2.2.1. The linear diophantine equation ax + by = c has a solution if and only
if d | c where d = (a, b).
Proof. First suppose that ax + by = c · · · (1) has a solution (x0 , y0 ). Then ax0 +
by0 = c · · · (2) Since (a, b) = d, so ax0 + by0 is a multiple of d. Therefore ax0 + by0 =
c = td ⇒ d | c.
d = au + bv, for some u, v ∈ Z ⇒ td = atu + btv · · · (5). From (4) and (5) we have
c = atu + btv ⇒ a(tu) + b(tv) = c. Comparing this with the given equation ax + by = c
b a
x = x0 + t, y = y0 − t,
d d
where t ∈ Z and d := (a, b) | c.
19
Proof. We first show that the expressions x = x0 + db t, y = y0 − ad t represent solutions
equation ax + by = c is satisfied by
b a
x = x0 + t, y = y0 − t,
d d
where t ∈ Z.
a(x − x0 ) = b(y0 − y)
a b
(x − x0 ) = (y0 − y)
d d
b a
¡ ¢
So d
| x − x0 and d
| y0 − y as ad , db = 1. It follows that x − x0 = db t, y0 − y = ad t and
hence
b a
x = x0 + t, y = y0 − t,
d d
where t ∈ Z.
Corollary 2.2.3. If (a, b) = 1 and (x0 , y0 ) is a particular solution of the linear diophan-
tine equation ax + by = c, then its general solutions are given by x = x0 + bt, y = y0 − at
where t ∈ Z.
Example 2.2.1. Find the positive integral solutions of the linear diophantine equation
172x + 20y = 1000.
Solution: Here a = 172, b = 20, c = 1000, and d := (a, b) = (172, 20) = 4 obtained
by the following Euclidean algorithm:
172 = 20 · 8 + 12
20 = 12 · 1 + 8
12 = 8 · 1 + 4
8=4·2+0
Since d | c, a solution to the equation 172x + 20y = 1000 exists by Theorem 2.2.1.
20
Using steps of Euclid’s algorithm from backward direction d = 4 can be expressed as
a linear combination of a = 172 and b = 20.
4 = 12 + (−1) · 8
= 12 + (−1) · {20 + (−1) · 12}
= 2 · 12 + (−1) · 20
= 2 · {172 + (−8) · 20} + (−1) · 20
= 172 · (2) + 20 · (−17)
∴ 172 · (500) + 20 · (−4250) = 1000
It follows that (x0 , y0 ) = (500, −4250) is a particular solution to the equation 172x+20y =
1000. Hence the general solutions given by Theorem 2.2.2 are:
b a
x = x0 + t, y = y0 − t, with t ∈ Z.
d d
That is, x = 500 + 5t, y = −4250 − 43t · · · (1)
Now the positive integral solutions can be obtained by considering the system of in-
equalities
4x = 500 + 5t > 0 and y = −4250 − 43t > 0
500 4250 36
⇒t>− = −100 and t < − = −98
5 43 43
36
⇒ −100 < t < −98
43
Since t is an integer so t = −99. Hence from (1)
Therefore (x, y) = (5, 7) is the only positive solution of the linear diophantine equation
172x + 20y = 1000.
21
Problem Plus 2
2. Calculate the value of Euler φ function at 30, 50, 100, 120, 240, and 264.
4. Find the positive integral solutions of the linear diophantine equation 18x+5y = 48.
5. Determine the particular solution of the linear diophantine equation 39x + 26y =
105.
6. A man presented a check for Taka 510 to a bank and wished to be paid in 20 and
50 taka notes. In how many ways and means can the cashier meet his request.
7.
22
Chapter 3
Congruences
The concept of congruences was introduced by Carl Friedrich Gauss about 1800, one of
the greatest mathematician of all time. Congruence often arises in every day life. For
instance, if the second of January is Sunday, then 9,16,23 of the same month are all
Sundays, since when they are divided by 7, the remainders are all 2.
be treated in the same manner as equations. Let m be a positive integer. Two integers
a and b leave remainders when divided by m. If the remainders are same we say that a
a ≡ b(mod m) or m | a − b.
Alternatively,
a ≡ b(mod m) or m | a − b.
Definition 3.1.2 (Incongruence). If the difference of two integers a and b is not divisible
by m, we shall say that a is incongruent to b modulo m, and we write
a 6≡ b(mod m) or m - a − b.
23
(i) a + c ≡ b + d(mod m)
(ii) a − c ≡ b − d(mod m)
(iii) ac ≡ bd(mod m)
(i) (a−b)+(c−d) = (a+c)−(b+d) = (t1 +t2 )m = t3 m, say. So, a+c ≡ b+d(mod m).
(ii) (a−c)−(b−d) = (a−b)−(c−d) = (t1 −t2 )m = t4 m, say. So, a−c ≡ b−d(mod m).
ac ≡ bd(mod m).
Theorem 3.1.3. a ≡ b(mod m) ⇒ na ≡ nb(mod m), but the converse need not hold.
Theorem
¡ ¢ 3.1.4 (Cancelation law). If na ≡ nb(mod m) and (n, m) = g, then a ≡
m
b mod g
Corollary 3.1.5. If a, b, g are integers such that na ≡ nb(mod m) and (m, n) = 1, then
a ≡ b(mod m).
Definition 3.1.3. Let a, b(b > 0) be any two integers. Then ∃ integers q, r such that
a = bq + r where 0 ≤ r < b. Then r is called the least residue of a modulo m.
24
r, a + r, 2a + r, 3a + r, 4a + r are
1, 4, 2, 0, 3 respectively.
Theorem 3.1.6 (Least Residue Theorem). If a and b(b > 0) are relatively prime and if
r is any integer, then the least residues modulo b of
r, a + r, 2a + r, 3a + r, · · · , (b − 1)a + r · · · (1)
are
0, 1, 2, 3, · · · , (b − 1) · · · (2)
in some rearranged order.
Proof. By definition, the least residue modulo b of any integer is one of the numbers
(2). There are b numbers (1) and b numbers (2). The theorem will be established if we
can show that no two of the numbers (1) have the same least residue (2). Any two of the
If they have the same least residue modulo b, they are congruent. Then
sa + r ≡ ta + r(mod b)
sa ≡ ta(mod b)
But s − t is positive and less than b so that it is not divisible by b. This contradiction
shows that no two distinct numbers (1) have the same least residue modulo b. Hence the
theorem is proved.
Corollary 3.1.7. If (a, b) = 1, then a, 2a, 3a, · · · , (b − 1)a are congruent modulo b to
1, 2, 3, · · · , (b − 1) in some order.
25
3.2 Fermat’s Theorem, Euler’s Theorem and Wil-
son’s Theorem
Theorem 3.2.1 (Fermat’s Theorem). If p is a prime and (a, p) = 1, then ap−1 ≡
1(mod p).
This theorem was conjectured by Fermat and was first proven by Euler in 1736. It is
Proof. By the least residue theorem the integers a, 2a, 3a, · · · , (p − 1)a are congruent
ap−1 1 · 2 · 3 · · · · · p − 1 ≡ 1 · 2 · 3 · · · · · p − 1(mod p)
Note. The converse of the Fermat’s theorem is not true, i.e. if ap−1 ≡ 1(mod p), p need
not be prime.
34 .1.3.5.7 ≡ 3.1.7.5(mod m)
⇒34 ≡ 1(mod m), since (1.3.5.7, m) = 1
∴ 3φ(8) ≡ 1(mod m)
⇒aφ(m) ≡ 1(mod m) where (a, m) = 1.
Theorem 3.2.3 (Euler’s Theorem). If m be any positive integer and a be any integer
prime to m, i.e. (a, m) = 1, then aφ(m) ≡ 1(mod m).
Proof. Let a1 , a2 , a3 , · · · , aφ(m) be the positive integers, in ascending order, less than m
and prime to m. Now consider the products aa1 , aa2 , aa3 , · · · , aaφ(m) . Since (a, m) = 1,
26
it follows that aa1 , aa2 , aa3 , · · · , aaφ(m) are congruent modulo m, not necessarily in order
of appearance to a1 , a2 , a3 , · · · , aφ(m) .
Then
aa1 ≡ a01 (mod m)
≡ a1 · a2 · a3 · · · · · aφ(m) (mod m)
¡ ¢
⇒ aφ(m) a1 a2 a3 · · · aφ(m) ≡ a1 a2 a3 · · · aφ(m) (mod m)
¡ ¢
Since (ai , m) = 1 for each i, we have a1 a2 a3 · · · aφ(m) , m = 1 and hence by the Cancela-
Proof. Since (a, p) = 1, by Euler’s theorem aφ(p) ≡ 1(mod p). But if p is a prime, then
Proof. Let a denote any one of the numbers 1, 2, 3, · · · , p − 1. Clearly (a, p) = 1, since
p is a prime. Then by the corollary of the least residue theorem 3.1.7 the integers
1, 2, 3, · · · , p − 1 · · · (2)
27
in some order. Hence there exists one and only one number in (1), say ka which will be
ka ≡ 1(mod p) · · · (3).
For all other values of a, k differ from a. This means that for every a belonging to
the p − 3 numbers
2, 3, · · · , p − 2 · · · (4)
there exists one and only one number k also belonging to (4) but different from a such
that
ka ≡ ak ≡ 1(mod p).
p−3
Thus, the numbers in (4) can be arranged into 2
pairs such that the product of
the integers of each pair is congruent to 1(mod p). Taking product of all these pairs we
obtain
2 · 3 · · · · p − 2 ≡ 1(mod p)
⇒1 · 2 · 3 · · · (p − 2) · (p − 1) ≡ (p − 1)(mod p)
have xp−1 − 1 ≡ 0(mod p). This equation has p − 1 roots with modulo p and the roots
are
1, 2, 3, · · · , p − 1
28
Since p is a prime, so either p = 2 or p is an odd prime.
−1 ≡ −1(mod 2)
−1 + 2 ≡ −1(mod 2)
(2 − 1)! ≡ −1(mod 2)
(p − 1)! ≡ −1(mod p)
Case-I: If p is an odd prime, then (−1)p−1 = 1 and hence (1) yields that (p − 1)! ≡
−1(mod p).
Proof. Suppose that p is not a prime. Then p is composite and so there exist a divisor
can write
(p − 1)! ≡ 0(mod d)
(p − 1)! 6≡ −1(mod d)
(p − 1)! 6≡ −1(mod p)
This is a contradiction to the converse statement. It follows that p must be a prime.
29
Problem Plus chapter 3
30
Chapter 4
Solutions of Congruences
Theorem 4.1.1. The linear congruence ax ≡ b(mod m) has a unique solution if (a, m) =
1.
Therefore
aφ(n)−1 ax ≡aφ(n)−1 b(mod m) (multiplying (1) by aφ(n)−1 )
31
Uniqueness: Let x1 and x2 be two solutions of the linear congruence (1). Then
⇒a(x1 − x2 ) ≡ 0(mod m)
⇒x1 ≡ x2 (mod m)
(i) no root if g - b
m m
(ii) exactly g incongruent roots if g | b. The roots are x0 , x0 + g
, x0 +2· g
,··· , x0 +
(g − 1) · mg where x0 is the unique root of ag x ≡ gb (mod mg ).
ax ≡ b(mod m)
⇒gAx ≡ gB(mod gM )
⇒Ax ≡ B(mod M ) · · · (1) which has a unique solution, say x0 , by (i) since (A, M ) = 1
⇒M | (Ax − B)
⇒m | (ax0 − b)
32
Suppose x0 is any other solution of ax ≡ b(mod m). Then ax0 ≡ b(mod m) · · · (3)
⇒m | (ax0 − ax0 )
⇒gM | gA(x0 − x0 )
⇒M | A(x0 − x0 )
So x0 = x0 + M (gq + r) = x0 + gM q + M r
m
⇒x0 = x0 + mq + r
g
m
⇒x0 ≡ x0 + r(mod m)
g
Thus, if ax ≡ b(mod m) admits a solution x0 , then it has g incongruent solutions,
namely:
m m m
x0 , x 0 + , x0 + 2 · , · · · , x0 + (g − 1) · .
g g g
1
If a ≡ b(mod m) then b ≡ a(mod m)
33
Example 4.1.2. Solve 18x ≡ 30(mod 42)
Solution: Here a = 18, b = 30, m = 42, g = (18, 42) = 6 and g | b. Therefore the given
congruence has exactly 6 solution. Now
34
Problem Plus chapter 4
35
Chapter 5
Quadratic Residuacity
à !à ! à !
a b ab
(ii) p p
= p
à ! à !
a b
(iii) a ≡ b(mod p) =⇒ p
= p
36
à ! à ! à ! à ! à !
p−1
a2 a2 b b 1 −1
(iv) p
= 1, p
= p
, p
= 1, p
= (−1) 2
Proof.
37
Problem Plus chapter 5
38
Chapter 6
6.1 Sets
Definition 6.1.1 (Set). A set is a well-defined list, collection, or class of objects.
The set operations are binary operations and are some useful tools to obtain new
sets from the old ones through combining(union), intersecting(intersection), and taking
differences(difference or complement).
Definition 6.1.2 (Union of two sets). The union of two sets A and B is mathematically
defined as
A ∪ B = {x : x ∈ A or x ∈ B or x ∈ both}.
Definition 6.1.3 (Intersection of two sets). The intersection of two sets A and B is
mathematically defined as
A ∩ B = {x : x ∈ A, x ∈ B}.
Definition 6.1.4 (Difference of two sets). The difference of two sets A and B is
mathematically defined as
A ∼ B = {x ∈ A : x ∈
/ B}.
39
6.1.1 Cartesian Product Sets and their visualization
Definition 6.1.6 (Cartesian Product set). The cartesian product of two nonvoid sets
A and B is mathematically defined as
A × B = {(a, b) : a ∈ A, b ∈ B}.
• The field properties of R, often called the algebraic properties, are based on the two
The historical development of real number system R was from the positive inte-
gers, called the natural numbers N to the rational numbers Q to the full real numbers
erty of R). The existence of irrationals is guaranteed by any of the three theorems called
40
(A5) existence of an additive inverse, called negative element, in R
The real number system R is a FIELD because it obeys the field axioms.
Definition 6.2.2 (Order Properties of R). On the field of real numbers there is a
binary relation denoted by <. For all x, y, z ∈ R this relation satisfies the following
properties:
(O4) compatibility of < and · : 0 < z and x < y implies x.z < y.z
The real number system R is an ORDERED FIELD because it obeys the order
axioms.
√ √ m
Proof. If 2 be a rational number, assume 2 = n
where (m, n) = 1 and m, n ∈
a rational number.
Definition 6.2.3 (Intervals). Open interval: (a, b) = {x ∈ R : a < x < b} and closed
interval: [a, b] = {x ∈ R : a ≤ x ≤ b}.
41
Definition 6.2.4 (Bounded Set). Let S ⊂ R. Then u ∈ R is said to be an upper bound
of S if s ≤ u for all s ∈ S and l ∈ R is said to be a lower bound of S if l ≤ s for all
s ∈ S. Set S is bounded above if it has an upper bound and bounded below if it has a
lower bound. It is said to be bounded if it has both upper bound and lower bound. It is
unbounded if it lacks either an upper bound or a lower bound.
Definition 6.2.5 (Suprema and Infima). If S is bounded above, then an upper bound
u ∈ R is said to be a least upper bound (lub) or supremum of S if no number smaller than
u is an upper bound of S. Similarly greatest lower bound(glb) or infimum is defined.
Theorem 6.2.2. An upper bound u of a nonempty set S ⊆ R is the supremum of S if
and only if for each ² > 0 there exists an s² ∈ S such that u − ² < s² .
Example 6.2.1. • If S = (0, 1), the sup S = 1, and inf S = 0.
• If S = [0, 2] the sup S = 2, greatest element of S and inf S = 0, the least element
os S.
• If S = { n1 : n ∈ N}, the sup S = 1, greatest element of S and inf S = 0. But the set
has no least element.
Example 6.2.2. Since every real number is an upper bound for the empty set, so the
empty set has no supremum. Similarly it has no infimum.
Theorem 6.2.3 (Completeness Axiom of R/The Supremum Property of R).
Every nonempty set of real numbers that has an upper bound has a supremum in R.
Theorem 6.2.4 (Completeness Axiom of R/The Infimum Property of R). Every
nonempty set of real numbers that has a lower bound has an infimum in R.
Theorem 6.2.5. There exists a positive real number x such that x2 = 2.
Proof. Let S = {s ∈ R : 0 ≤ s, s2 < 2}. Since 1 ∈ S, the set is not empty. Also,
Supremum Property implies that the set S has a supremum in R and let x = sup S(here
√
sup S = 2).
Thus after some mathematical reasoning we can prove that x is a real number.
Definition 6.2.6. A cut or Dedekind cut of the real line is a pair of nonempty subsets
L and R of R such that L ∪ R = R and for every x ∈ L and y ∈ R, x < y. In terms of
geometry, L is the left-hand set and R is the right-hand set.
Example 6.2.3. If (−∞, 1] ∪ (1, ∞) = R, then L = (−∞, 1] and R = (1, ∞)
Theorem 6.2.6 (Completeness Axiom of R). Let L and R define the cut of the real
line. Then there is one and only one real number z such that for every x ∈ L and y ∈ R,
x ≤ z ≤ y.
42
√
Example 6.2.4.√ We have
√ earlier proved that √2 is not√a rational
√ number and observe
here that (−∞, 2)
√ ∪ ( 2, ∞) 6= R, but (−∞, 2) ∪ { 2} ∪ ( 2, ∞) = R. So by the
Dedekind theorem 2 is a real number.
Thus, from the above discussion we may conclude that R is complete BUT Q is not.
6.3 Functions
Definition 6.3.1 (Function). A function f from a set A into a set B, denoted by
f : A → B, is a rule that assigns to each element x ∈ A a unique element y ∈ B, and we
write y = f (x).
is a subset f of the cartesian product A × B such that for each x ∈ A there exists a
unique y ∈ B, and we write y = f (x). Set A is called the domain of f and the set
Problem Plus 6
1.
2.
43
Chapter 7
7.1 Sequence
Definition 7.1.1 (Sequence). A sequence in R is a function from the set N of natural
numbers into the set R of real numbers. Thus, if X : N → R is a sequence, then the value
of X at n ∈ N is denoted by xn rather than X(n). We denote sequence by the notations:
44
Example 7.1.3. Show that lim(1/n) = 0.
Solution: For given ² > 0, 1/² > 0. Hence by Archimedean Property there exists a
natural number K := K(²) such that 1/² < K, then for any n ∈ N such that n ≥ K =⇒
1/² < K ≤ n so that 1/n < ². Therefore, if n ≥ K, then | n1 − 0| = n1 < ².
Theorem 7.1.3. The sum, difference, scalar multiple, product, and division (not always)
of two convergent sequences are convergent.
Example 7.1.4. Use the above theorem to show that lim( sinn n ) = 0
Theorem 7.1.6 (Ratio Test). Let xn be a sequence of positive real numbers such that
L = lim(xn+1 /xn ) exists. If L < 1, then xn converges and lim(xn ) = 0.
Example 7.1.7. Consider the sequence xn+1 = (xn + a/xn )/2 where a > 0, x1 > 0.
(This calculates the square roots of a). Here xn − xn+1 ≥ 0 =⇒ xn+1 ≤ xn for all
n ≥ 2. It follows from Monotone
√ Convergence Theorem that x = lim(xn ) exists. So
x = (x + a/x)/2 =⇒ x = a.
45
7.2 Sequence of Functions
What we have studied so far is the sequence of constants i.e., sequence of constant
functions. In this section we shall discuss about the sequence of arbitrary functions.
Therefore, the sequence of function is a sequence whose terms are functions rather
than real numbers. Sequences of functions arise naturally in real analysis and are espe-
cially useful in obtaining approximations to a given function and defining new functions
Note that the limit of the sequence of functions depend on x that is the sequence of
functions converges pointwise. Also note that the functions in the sequence are continuous
Here note that the limit of the sequence of functions does not depend on x that is the
sequence of functions converges uniformly. Also note that the functions in the sequence
Therefore according to these examples the following two definitions are in order:
In other words, fn → f (pointwise) on A if for every ² > 0 and for each x ∈ A, there
exists a natural number K := K(², x) such that if n ≥ K
46
Definition 7.2.2 (Uniform Convergence). Let < fn > be a sequence of functions on
B ⊆ R to R, let A ⊆ B, and f : A ⊆ B → R. Then the sequence < fn > converges
uniformly on A to f if, for ALL x ∈ A, limn→∞ fn (x) = f (x).
In other words, fn → f (uniformly) on A if for every ² > 0 and for all x ∈ A, there
exists a natural number K := K(²) (depending on ² but not on x) such that if n ≥ K
Notations:
47
Example 7.2.5. Discuss the uniform convergence of the sequence < fn > where f : A =
x
[0, 1] → ∞ is defined by fn (x) = 1+nx , n ∈ N.
Solution: First find the pointwise convergence and then the uniform convergence. Clearly,
fn (x) → f = 0 on A. We claim that this convergence is uniform as well. To this end, let
² > 0, then 1/² > 0, by Archimedean theorem, ∃K ∈ N such that 1/² < K ⇒ 1/K < ².
Now, if n ≥ K, then
¯ ¯ ¯ x ¯ ¯x¯
¯ ¯ ¯ ¯ ¯ ¯
¯fn (x) − f (x)¯ = ¯ − 0¯ < ¯ ¯ = 1/n for all x ∈ A
1 + nx nx
< 1/K < ²
Therefore, fn ⇒ f on A
Lemma 7.2.1. A sequence of functions fn on B ⊆ R does not converge uniformly on
A ⊆ B to a function f : A → R iff for some ²0 > 0 there is a subsequence fnk of fn and
a sequence xk in A such that
¯ ¯
¯ ¯
¯fnk (xk ) − f (xk )¯ ≥ ²0 for all k ∈ N.
Example 7.2.6. Discuss the uniform convergence of the sequence < fn > where fn :
nx
A = [0, ∞) → R is defined by fn (x) = 1+n 2 x2 , n ∈ N.
Solution: First find the pointwise convergence and then the uniform convergence. Clearly,
fn (x) → f = 0 on A. We claim that this convergence is not uniform on A. To this end,
let ² > 0, and let nk = n, xk = 1/n, then
¯ ¯ ¯ ¯
¯ ¯ ¯ ¯
¯fnk (xk ) − f (xk )¯ = ¯fn (1/n) − f (1/n)¯ = 1/2 ≥ ² for all n ∈ N.
Theorem 7.2.4 (Interchange of Limit and Integral). Let < fn > be a sequence of
functions that are (Riemann) integrable on A = [a, b] and suppose that < fn > converges
Rb
uniformly on A to a function f . Then f is (Riemann) integrable on A and a f (x)dx =
Rb
limn→∞ a fn (x)dx i.e.,
Z b Z b
lim fn (x)dx = lim fn (x)dx
n→∞ a a n→∞
.
48
Theorem 7.2.5 (Bounded Convergence Theorem). Let < fn > be a sequence of
(Riemann) integrable functions on A = [a, b] and suppose that < fn > converges on A
to a (Riemann) integrable function f . Suppose also that there exits M > 0 such that
|fn (x)| ≤ M for all x ∈ A. Then
Z b Z b
f (x)dx = lim fn (x)dx.
a n→∞ a
49
Problem Plus 7
1.
2.
50
Chapter 8
8.1 Series
Definition 8.1.1
P (Series).
PSum of the terms of an infinite sequence is called a series.
Given a series ∞ x
n=1 n = xn = x1 + x2 + x3 + · · · , let sn denote its n-th partial sum:
n
X
sn = xk = x1 + x2 + x3 + · · · + xn .
k=1
If the sequence
P sn is convergent, i.e. if x is a real number such that lim(sn ) = x, then the
series xn is called convergent and we write
n
X ∞
X
lim xk = xk = x1 + x2 + x3 + · · · = x.
n→∞
k=1 k=1
The number x is called the sum of the series. Otherwise, the series is divergent.
P∞ Pn
Note that k=1 xk = limn→∞ k=1 xk .
51
P
Theorem 8.1.2. The p-series ∞ 1
n=1 np is convergent if p > 1 and divergent if p ≤ 1.
When p = 1 the series is called the harmonic series.
P
Theorem 8.1.3. If the series ∞ n=1 xn is convergent , then lim(xn ) = 0. But the converse
is not true in general, e.g., harmonic series.
Theorem
P∞ 8.1.4 (The Test for Divergence). If xn → ∞ or xn 9 0 then the series
n=1 xn is divergent.
P n2
Example 8.1.2. Test whether the series ∞ n=1 5n2 +4 is convergent or divergent.
2
Solution: Here xn = 5nn2 +4 → 1/5 6= 0 so The Test for Divergence implies that the series
is divergent.
Theorem 8.1.5. The sum, difference, and scalar multiple of two convergent series are
convergent.
P
Example 8.1.3. Find the sum of the series ∞ 3 1
n=1 ( n(n+1) + 2n ).
Solution:
P∞ 1 Here second series is a geometric
P∞ series with a = Pn1/2 and r = 1/2, so
a 3 1
n=1 2n = 1−r
= 1. The first series is n=1 n(n+1) = 3 limn→∞ i=1 i(i+1) = limn→∞ (1−
1
n+1
) = 3 · 1 = 3. Therefore the sum of the given series is 3 + 1 = 4.
P
Theorem 8.1.6 (Cauchy Criterion for Series). A series xn in R is convergent iff
for each ² > 0 there is a natural number K := K(²) such that if m > n ≥ K, then
Theorem 8.1.9 (Limit Comparison Test). Let xn and yn be positive real sequences
and L = lim(xn /ynP ) P
(a) If L 6= 0, thenP xn is convergent iff Pyn is convergent.
(b) If L = 0 and yn is convergent, then xn is convergent.
52
P 2
Example 8.1.5. Test the convergence of the series ∞ 2n
√ +3n
n=1 5+n5 .
Solution: Note that the dominant
√ part of the numerator is 2n2 and the dominant
2 2n2
part of the denominator is n . This suggests taking xn = 2n
5 √ +3n ,
5+n 5 y n = √
n5
=
2 2 1/2 P P
n1/2
, limn→∞ xynn = limn→∞ 2n√ +3n · n
5+n5 2
= 1 Since yn = 2 1/n1/2 is divergent
(p-series with p = 1/2 < 1), the given series diverges by the Limit Comparison Test.
Theorem 8.1.10 (Root Test). Let xn be a sequence in R and
r := lim(|xn |1/n ).
P
Then xn is absolutely convergent if r < 1 and divergent if r > 1.
P ¡ 2n+3 ¢n
Example 8.1.6. Test the series ∞ n=1
¡ 2n+3 ¢n 3n+2 p 2n+3
Solution: Root Test with xn = 3n+2 gives: n |xn | = 3n+2 → 2/3 < 1 Thus, the given
series converges by the Root Test.
Theorem 8.1.11 (Ratio Test). Let xn be a sequence in R and
r := lim(|xn+1 |/|xn |).
P
Then xn is absolutely convergent if r < 1 and divergent if r > 1.
P n n3
Example 8.1.7. Test the series ∞ n=1 (−1) 3n for absolutely convergence.
3
Solution: Ratio Test with xn = (−1)n 3nn gives: | an+1
an
| = | 31 ( n+1
n
)3 | = | 13 (1 + 1/n)3 | →
1
3
< 1 Thus, by the Ratio Test, the given series is absolutely convergent and therefore
convergent.
Theorem 8.1.12 (Raabe’s Test). Let xn be a sequence of nonzero real numbers and
³ ³ |xn+1 | ´
r := lim n 1 − )
|xn |
P
Then xn is absolutely convergent if r > 1 and is not absolutely convergent if r < 1.
Theorem 8.1.13 (Integral Test). RLet f be a continuous, positive decreasing function P
∞
on [1, ∞) and xn = f (n). Then if 1 f (x)dx is convergent (divergent), then xn is
convergent (divergent).
P
Example 8.1.8. Test the series ∞ lnn
n=1 n for convergence.
Solution: The function f (x) = (lnx)/x is positive and continuous for x > 1 because the
logarithm function is continuous. But it is not clear whether or not f is decreasing, so
f 0 (x) = 1−lnx
x2 R
. Thus, f 0 (x) < 0 when 1 − lnx < 0, i.e., x > e. So f is decreasing when
∞ R t lnx
x > e. Here 1 lnx x
dx = lim t→∞ 1 x
dx = ∞ Therefore, by the Integral Test the given
series is divergent.
53
8.2 Series of Functions
We have studied the series of constants(i.e. constant functions). He we shall study the
Definition 8.2.1 (Pointwise and Uniform Convergence of Series). Let P the function
f and the sequence of functions < fn P > be defined on B ⊆ R and sn = nk=1 fk be the
sequence of partial sums of the series ∞ n=1 fn in B ⊆ R. P∞
Then if sn → f (pointwise)
P∞ on B (i.e., s n → f on B), then n=1 fn (x) → f (x) for
x ∈ B and we write n=1 fn (x) = f (x)(pointwise) P∞
Then if sn → f (uniformly)
P on B (i.e., s n ⇒ f on B), then n=1 fn (x) ⇒ f (x) for
x ∈ B and we write ∞ f
n=1 n (x) = f (x)(unif ormly)
Theorem 8.2.2 (Interchange of Sum and Derivative). (J. R. Marsden [?]) IfP < fn >
0 ∞
P∞on a0 set A = (a, b) ⊆ R, the < fn > are continuous, n=1 fn
are differentiable functions
converge pointwise, and n=1 fn converges uniformly, then
³X
∞ ´0 ∞
X
fn = fn0 .
n=1 n=1
Theorem 8.2.3 (Interchange of Integral and Sum). Let < fn > be a sequence of
functions that are (Riemann) integrable on A = [a, b] and suppose that < fn > converges
uniformly on AThen we may interchange the order of integration and summation:
Z b³X
∞ ´ ∞ ³Z b
X ´
fn (x) dx = fn (x)dx .
a n=1 n=1 a
54
Problem Plus 8
1.
2.
55
Chapter 9
Vδ (c) − {c} ∩ A 6= φ
Vδ (c) ∩ A 6= φ
Example 9.1.1.
(a) Let A = (0, 1) ∪ {3}. Then the set of cluster points of A is [0, 1] and the set of limit
points of A is [0, 1] ∪ {3}.
(e) Let S = {x ∈ R : x ∈ [0, 1] and x is rational}. Then all points of [0, 1] are accumu-
lation points of S.
Definition 9.1.3 (² − δ : Definition of Limit of a Function). Let A ⊆ R, and let
c be a cluster point of A. We say that a real number L is a limit of f at c, written
limx→c f (x) = L or f (x) → L as x → c if, given any ²−neighborhood V² (L) of L, there
exists a δ−neighborhood Vδ (c) of c such that if x 6= c is any point of Vδ (c) ∩ A, then f (x)
belongs to V² (L).
56
Example 9.1.2. Apply ² − δ definition of limit to illustrate that
lim (2x − 2) = 6
x→4
In other words, f is continuous at a point c ∈ A, written limx→c f (x) = f (c) if, given
any ² > 0 there exists a δ := δ(², x) > 0 such that if |x − c| < δ, then |f (x) − f (c)| < ²
57
Example 9.2.1.
Example 9.2.2.
(b) f (x) = 1/x is continuous on A = (0, ∞) but not bounded on A. f (x) is not even
bounded when restricted to the set B = (0, 1).
Example 9.2.3. 1. Continuous function on a set A does not necessarily have an ab-
solute maximum or an absolute minimum on the set.
2. f (x) = 1/x has neither an absolute maximum nor an absolute minimum on the set
A = (0, ∞).
3. The same function has neither an absolute maximum nor an absolute minimum
when restricted to the open set B = (0, 1).
4. While the same function has BOTH an absolute maximum and an absolute mini-
mum when restricted to the closed set C = [0, 1].
58
5. The function f (x) = x2 on A = [−1, +1] has two points x = ±1 giving the absolute
maximum and a single point x = 0 yielding the absolute minimum on A.
6. The constant function f (x) = c is such that for x ∈ R is such that EVERY point
of R is BOTH an absolute maximum and an absolute minimum point for f .
59
Problem Plus 9
1.
2.
60
Chapter 10
Differentiation
The Mean Value Theorem relates the values of a function to values of its derivative
One can easily deduce Roll’s Theorem from Mean Value Theorem. Now we shall
present the Taylor’s Theorem which is simply the generalization of the Mean Value
Theorem.
61
Theorem 10.0.13 (Taylor’s Theorem). Let f : I = [a, b] → R be a function such that
f and its derivatives up to order n are continuous on a closed interval [a, b] and f (n+1)
exists on the open interval (a, b). If x0 ∈ I, then for any x in I there exists a point c
between x and x0 such that
f 00 (x0 ) f (n) (x0 ) f (n+1) (c)
f (x) = f (x0 )+f 0 (x0 )(x−x0 )+ (x−x0 )2 +· · ·+ (x−x0 )n + (x−x0 )(n+1)
2! n! (n + 1)!
where the last term
f (n+1) (c)
Rn (x) = (x − x0 )(n+1)
(n + 1)!
is called the remainder.
Theorem 10.0.14 (Taylor’s Inequality or The Remainder Estimation Theo-
rem). If |f (n+1) (x)| ≤ M for all x ∈ I, and a number x0 ∈ I, then
M
|Rn (x)| ≤ |x − x0 |n+1 .
(n + 1)!
Theorem 10.0.15. If n → ∞ then Rn (x) → 0.
x2 −4
• Find limx→2 x−2
62
Problem Plus 10
1.
2.
63
Chapter 11
Riemann Integration
Riemann Integration
64
Problem Plus 11
1.
2.
65
Appendix A
Reviews
Reviews
(A.1) a + b + c
+d+e+f
+g+h+i
j+k+l
stack-out
1. Shah Noor
2 Happy
66
1
2 % 1.simple newton
3
4 function [root] = simple newton(x0,tol);
5 % Finds the root of floating sphere cubic using Newton's method
6 % Input:
7 % x0 − initial guess for root
8 % tol − desired absolute error
9 % Output:
10 % root − approximate root
11 format long e;
12 error = 1e10;
13 x0=0.5; tol = 1e−10;
14 xk=x0;
15 k=0;
16 while error > tol
17 k=k+1;
18 xk = xk − (xkˆ2 − (xkˆ3+1)/3)/(2*xk − xkˆ2);
19 error = abs(xkˆ2−(xkˆ3+1)/3);
20 disp(['Iteration' sprintf('%2i',k) ' approx root ' ...
21 sprintf('%15.13f',xk) ' abs error ' sprintf('%3.2e',error)] )
22 end
8 for k = 1 : n
9 f1 = exp(−2*x(k)) − 2*ycal(k) ;
10 f2 = exp(−2*(x(k)+h/2)) − 2*(ycal(k) +h*f1/2) ;
11 f3 = exp(−2*(x(k)+h/2)) − 2*(ycal(k) +h*f2/2) ;
12 f4 = exp(−2*(x(k)+h)) − 2*(ycal(k) +h*f3) ;
13 sf = f1 + 2 * f2 + 2 * f3 + f4 ;
14 ycal(k+1) = ycal(k) + h*sf/6 ;
15 end yexact = (0.1 + x).*exp(−2*x) ; a = [x' ycal' yexact'] ;
16 b=a(:,1)
17 %plot(x,ycal,'r', x,yexact,'b'); grid on;
18 plot(a(:,1),a(:,2),'r', a(:,1),a(:,3),'b'); grid on;
19 xlabel('x');ylabel('ycal and yexact') return
67
Problem Plus A
1.
2.
68
Index
Goldbach Conjecture, 13
greatest common divisor(g.c.d.), 5
incongruence, 23
69
Bibliography
[2] Prof. Dr. Md. Fazlur Rahman, “Theory of Numbers”, Third Edition,
Titas Publications, Bangladesh, (2009)
70