0% found this document useful (0 votes)
32 views19 pages

Samsung Notes

Uploaded by

krgaurav2188
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
32 views19 pages

Samsung Notes

Uploaded by

krgaurav2188
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 19

Q.

Find the supremum of the set


\ n/(n + 1) |n in N\
=ans. To find the supremum of the set, we need to find the smallest upper bound.
Let's first look at the sequence of terms:

1/2, 2/3, 3/4, 4/5, ...

We can see that as n increases, the value of n/(n+1) gets closer and closer to 1.
In fact, we can prove that 1 is the supremum of the set.

To do this, we need to show that 1 is an upper bound (i.e. all terms in the set are
less than or equal to 1) and that there is no smaller upper bound.

First, let's show that 1 is an upper bound. For any n in N:

n/(n+1) <= 1

This is equivalent to:

n <= n + 1

which is always true. Therefore, 1 is an upper bound.

Now let's show that there is no smaller upper bound. Suppose there exists a
number x such that x < 1 and x is an upper bound. Then, for any n in N:

n/(n+1) <= x

This is equivalent to:

n <= xn + x

Since x < 1, we can choose n large enough such that xn + x < n+1. But this means
that there is at least one term in the set (namely n/(n+1)) that is greater than x,
which contradicts the assumption that x is an upper bound. Therefore, there is no
smaller upper bound than 1.

Therefore, the supremum of the set is 1.


Q. Check whether (sqrt(n ^ 2) - sqrt(n)) n in N the is monotone or not.
Ans. To check whether the sequence (sqrt(n^2) - sqrt(n)) is monotone or not, we
need to determine whether it is increasing or decreasing.

Let's first simplify the expression:


sqrt(n^2) - sqrt(n) = n - sqrt(n)

To determine if the sequence is increasing or decreasing, we need to compare


each term to the next term. Let's consider the difference between two
consecutive terms:

(n+1) - sqrt(n+1) - n + sqrt(n)

Simplifying this expression, we get:

1 - (sqrt(n+1) - sqrt(n))

Now, we can see that if (sqrt(n+1) - sqrt(n)) is increasing, then 1 - (sqrt(n+1) -


sqrt(n)) is decreasing. Similarly, if (sqrt(n+1) - sqrt(n)) is decreasing, then 1 -
(sqrt(n+1) - sqrt(n)) is increasing.

To determine whether (sqrt(n+1) - sqrt(n)) is increasing or decreasing, we can


take the derivative with respect to n:

d/dn (sqrt(n+1) - sqrt(n)) = 1/2(sqrt(n+1) - sqrt(n))^(-1/2) * (1/(2sqrt(n+1)) - 1/


(2sqrt(n)))

Simplifying this expression, we get:

d/dn (sqrt(n+1) - sqrt(n)) = 1/(2sqrt(n)(sqrt(n+1) + sqrt(n)))

Since this expression is always positive, we can conclude that (sqrt(n+1) - sqrt(n))
is increasing. Therefore, 1 - (sqrt(n+1) - sqrt(n)) is decreasing.

In other words, the sequence (sqrt(n^2) - sqrt(n)) is a decreasing sequence.

Therefore, the sequence is monotone.


Q. Find the n ^ (th) partial sum of the following series, and hence check whether
they are convergent or sum n = 1 to ∞ ln(1 + 1/n)

Ans. The n^th term of the series is ln(1 + 1/n), so the partial sum up to the n^th
term is:

S_n = ln(1 + 1/1) + ln(1 + 1/2) + ln(1 + 1/3) + ... + ln(1 + 1/n)

We can simplify this expression using the fact that ln(a) + ln(b) = ln(ab):

S_n = ln(1*2*3*...*n) + ln(1 + 1/2) + ln(1 + 1/3) + ... + ln(1 + 1/n)


Simplifying further using the fact that ln(ab) = ln(a) + ln(b), we get:

S_n = ln(n!) + ln((1 + 1/2)(1 + 1/3)...(1 + 1/n))

We can approximate the second term using the fact that (1 + 1/n) is close to 1 for
large n:

(1 + 1/2)(1 + 1/3)...(1 + 1/n) ≈ 1

Therefore, the partial sum up to the n^th term is approximately:

S_n ≈ ln(n!)

To check whether this series converges or diverges, we can use the ratio test:

lim n->∞ |ln((n+1)!) / ln(n!)| = lim n->∞ ln(n+1) - ln(n) = lim n->∞ ln(1 + 1/n)

Using L'Hopital's rule, we can evaluate this limit as:

lim n->∞ ln(1 + 1/n) / (1/n) = lim n->∞ (n/(n+1)) / (1/n^2) = lim n->∞ n/(n+1) * n^2
=1

Since the limit is equal to 1, the ratio test is inconclusive. However, we can use
the fact that ln(1 + x) < x for all x > 0 to show that ln(1 + 1/n) < 1/n:

ln(1 + 1/n) = ∫_1^(1+1/n) 1/x dx < 1/n

Therefore, S_n < ln(n!) + 1 + 1/2 + ... + 1/n. This last sum is known as the
harmonic series, which diverges. Therefore, S_n also diverges.

In conclusion, the series n = 1 to ∞ ln(1 + 1/n) diverges.


Q. Let f be a differentiable function whose derivative never vanishes on [a, b].
Show that fis either strictly decreasing or strictly increasing.
Ans. By the Mean Value Theorem, for any two distinct points x and y in [a, b], there
exists a point c between them such that:

f(y) - f(x) = f'(c)(y - x)

Since f'(c) is never zero on [a, b], we can conclude that the sign of f(y) - f(x) is the
same as the sign of (y - x). That is, if y > x, then f(y) > f(x), and if y < x, then f(y) <
f(x). Therefore, f is either strictly increasing or strictly decreasing on [a, b].
Q. Show that the series sum n = 1 to ∞ (sin(n ^ 3 * x))/(n ^ 3) is uniformly
convergent on [0, ∞[.
Ans. To show that the series is uniformly convergent on [0, ∞[, we need to show
that the partial sums of the series form a uniformly Cauchy sequence.

Let Sn(x) be the nth partial sum of the series:

Sn(x) = sum k=1 to n (sin(k^3*x))/(k^3)

We want to show that for any ε > 0, there exists an N such that for all n, m > N and
for all x in [0, ∞[, |Sn(x) - Sm(x)| < ε.

First, note that for any x in [0, ∞[ and any k > 0, |sin(k^3*x)| <= 1. Therefore,

|Sn(x) - Sm(x)| = |sum k=m+1 to n (sin(k^3*x))/(k^3)| <= sum k=m+1 to n |


sin(k^3*x)/(k^3)|

Now, we can use the comparison test to bound this sum. Since |sin(k^3*x)/(k^3)|
<= 1/(k^3), we have

sum k=m+1 to n |sin(k^3*x)/(k^3)| <= sum k=m+1 to ∞ 1/(k^3)

The series sum 1/(k^3) converges (by the p-test), so there exists an M such that
for all m > M, sum k=m+1 to ∞ 1/(k^3) < ε/2.

Now, let N be max(M, ceil(π/(x^3))), where ceil is the ceiling function. Then for all
n, m > N and for all x in [0, ∞[, we have

|Sn(x) - Sm(x)| <= sum k=m+1 to n |sin(k^3*x)/(k^3)| <= sum k=m+1 to ∞ 1/(k^3) <
ε/2

Also, note that for all x in [0, ∞[, we have |sin(k^3*x)/(k^3)| <= 1/(k^3) <= 1/N^3,
since k > N. Therefore,

|Sn(x) - Sm(x)| <= sum k=m+1 to n |sin(k^3*x)/(k^3)| <= sum k=N+1 to ∞ 1/(k^3) <=
sum k=N+1 to ∞ 1/N^3 = 1/N^2 < ε/2

Combining these inequalities, we get

|Sn(x) - Sm(x)| < ε

for all n, m > N and for all x in [0, ∞[. Therefore, the partial sums of the series form
a uniformly Cauchy sequence on [0, ∞[, and the series is uniformly convergent on
[0, ∞[ by the Cauchy criterion.
Q. Let f and g be two real-valued functions defined on [a, b] such that f is Riemann
integrable and g is differentiable with g' * (x) =f(x) forall x in[a, b].
Then show that integrate f(x) dx from a to b = g(b) - g(a)
Ans. By the fundamental theorem of calculus, we have

g(x) - g(a) = integral from a to x of g'(t) dt

Since g'(t) = f(t) for all t in [a, b], we can substitute to get

g(x) - g(a) = integral from a to x of f(t) dt

Now, we want to show that this is equal to the integral of f(x) from a to b. We can
split the integral into two parts:

integral from a to b of f(x) dx = integral from a to x of f(x) dx + integral from x to b


of f(x) dx

Using the first part of the equation we derived above, we can rewrite the left-hand
side as

integral from a to b of f(x) dx = g(b) - g(a)

Now, we need to show that the right-hand side is equal to this. To do this, we will
use the fact that f is Riemann integrable on [a, b]. This means that for any ε > 0,
there exists a partition P of [a, b] such that

U(f, P) - L(f, P) < ε

where U(f, P) and L(f, P) are the upper and lower Riemann sums of f with respect
to P.

Let P = a = x0 < x1 < ... < xn = b be such a partition. Then we have

integral from a to b of f(x) dx = lim as ||P|| -> 0 sum i=1 to n f(xi*)(xi - xi-1)

where xi* is any point in [xi-1, xi]. Now, let M be the maximum value of |g'(t)| on [a,
b]. Since g is differentiable on [a, b], it is also continuous on [a, b], so it is bounded.
Therefore, M is finite.

Now, we can use the mean value theorem to write

g(xi) - g(xi-1) = g'(t)(xi - xi-1)

for some t in [xi-1, xi]. Therefore,

|g(xi) - g(xi-1)| = |g'(t)|(xi - xi-1) <= M(xi - xi-1)


Using this inequality, we can bound the difference between the integral of f(x)
from a to b and g(b) - g(a):

|integral from a to b of f(x) dx - (g(b) - g(a))| = |sum i=1 to n f(xi*)(xi - xi-1) - (g(xi) -
g(xi-1))|

<= sum i=1 to n |f(xi*)(xi - xi-1) - (g(xi) - g(xi-1))|

<= sum i=1 to n M(xi - xi-1)

= M(b - a)

Since this holds for any partition P, we have

|integral from a to b of f(x) dx - (g(b) - g(a))| <= M(b - a)

As ε can be made arbitrarily small, we have shown that

integral from a to b of f(x) dx = g(b) - g(a)

as desired.
Q. Using the definition show that the sequence (1/(sqrt(n))) n in N is cauchy.
Ans. To show that the sequence (1/(sqrt(n))) n in N is Cauchy, we need to show
that for any ε > 0, there exists an N such that for all m, n > N, |1/(sqrt(m)) - 1/
(sqrt(n))| < ε.

Let ε > 0 be given. We want to find an N such that for all m, n > N, |1/(sqrt(m)) - 1/
(sqrt(n))| < ε.

Without loss of generality, assume that m > n. Then we have

|1/(sqrt(m)) - 1/(sqrt(n))| = |(sqrt(m) - sqrt(n))/(sqrt(m)sqrt(n))|

Since m > n, we have sqrt(m) > sqrt(n), so

|(sqrt(m) - sqrt(n))/(sqrt(m)sqrt(n))| = (sqrt(m) - sqrt(n))/n

Now, let N be such that N > 1/ε^2. Then for all m, n > N, we have

|m - n| > 1/ε^2

This implies that


sqrt(m) - sqrt(n) > (m - n)/2sqrt(n) > (m - n)/2m

Using this inequality, we can bound the difference between 1/(sqrt(m)) and 1/
(sqrt(n)):

|1/(sqrt(m)) - 1/(sqrt(n))| = (sqrt(m) - sqrt(n))/n > (m - n)/(2mn)

Since m, n > N, we have

(m - n)/(2mn) < (m - n)/(2N^2)

Therefore,

|1/(sqrt(m)) - 1/(sqrt(n))| < (m - n)/(2N^2)

Since m, n > N, we have

|m - n| < 1/ε^2

Therefore,

|1/(sqrt(m)) - 1/(sqrt(n))| < 1/(2N^2ε^2)

Since N > 1/ε^2, we have

1/(2N^2ε^2) < ε

Therefore, for all m, n > N, |1/(sqrt(m)) - 1/(sqrt(n))| < ε, which shows that the
sequence (1/(sqrt(n))) n in N is Cauchy.
Q. Check whether the set of integers is countable or not.
Ans. The set of integers is countable.

To see why, we can construct a bijection between the set of integers and the set
of natural numbers. One possible bijection is given by:

f(n) =
- n/2 if n is even
- (n-1)/2 if n is odd

This function maps each integer to a unique natural number, and vice versa.
Therefore, the set of integers has the same cardinality as the set of natural
numbers, which is countable.
Q. Show that if f is a real-valued continuous function defined on a closed interval
[α, β], then f is Riemann integrable over [α, β].
Ans. To show that f is Riemann integrable over [α, β], we need to show that for
any ε > 0, there exists a partition P of [α, β] such that the difference between the
upper and lower Riemann sums of f over P is less than ε.

Let ε > 0 be given. Since f is continuous on a closed interval, it is uniformly


continuous on that interval. That is, for any δ > 0, there exists a number η > 0 such
that if |x - y| < η, then |f(x) - f(y)| < δ.

Choose δ = ε/(β - α). Then there exists η > 0 such that if |x - y| < η, then |f(x) - f(y)|
< ε/(β - α).

Now choose a partition P of [α, β] such that the length of each subinterval of P is
less than η. Let U(P, f) and L(P, f) denote the upper and lower Riemann sums of f
over P, respectively.

Since f is continuous on [α, β], it attains its maximum and minimum values on
each subinterval of P. Let M_i and m_i denote the maximum and minimum values
of f on the i-th subinterval of P, respectively.

Then we have:

U(P, f) - L(P, f) = Σ(M_i - m_i)(x_i - x_i-1)

where the sum is taken over all subintervals of P. Since each subinterval has
length less than η, we have |M_i - m_i| < ε/(β - α) for all i. Therefore,

U(P, f) - L(P, f) < Σε(x_i - x_i-1) = ε(β - α)

Since ε > 0 was arbitrary, this shows that f is Riemann integrable over [α, β].

Q.Let f be a function defined by f(x) = (x - 1)/(x ^ 2 + 3) x R. Using the ɛ - δ
definition, show that f(x) -> 1/7 whenever x -> 2
Ans. Let ε > 0 be given. We need to find a δ > 0 such that if 0 < |x - 2| < δ, then |f(x)
- 1/7| < ε.

Note that for x ≠ 2, we have:

|f(x) - 1/7| = |(x - 1)/(x^2 + 3) - 1/7|


= |(7(x - 1) - (x^2 + 3))/(7(x^2 + 3))|
= |(-x^2 + 7x - 4)/(7(x^2 + 3))|

We want to bound the numerator and denominator separately. First, note that for
0 < |x - 2| < 1, we have:

|x - 1| < 1
|x + 2| < 3

Therefore,

|x^2 + 3| = |(x - 1)(x + 2)| < 3|x - 1| < 3


|-x^2 + 7x - 4| = |(x - 4)(-x + 1)| < (3 + ε)|x - 2|

where we used the fact that x is close to 2, so x - 4 and -x + 1 have opposite signs
and their absolute values are both less than 1.

Therefore,

|f(x) - 1/7| = |(-x^2 + 7x - 4)/(7(x^2 + 3))| < (3 + ε)|x - 2|/(7*3)

Now choose δ = min1, (7*3*ε)/(3 + ε). Then if 0 < |x - 2| < δ, we have:

|x^2 + 3| < 3
|-x^2 + 7x - 4| < (3 + ε)|x - 2|

Therefore,

|f(x) - 1/7| < (3 + ε)|x - 2|/(7*3) < ε

This shows that f(x) -> 1/7 whenever x -> 2.



Q. Let f: [0, 2] R be f(x) = defined by -1, 0 ≤ x ≤1 and 1, 1<x<2

Show that there is no real-valued function defined on [0, 2] whose derivative is f.


Ans. Suppose there exists a function F(x) defined on [0, 2] such that F'(x) = f(x) for
all x in [0, 2]. Then we can use the Mean Value Theorem to show that F(2) - F(0) =
2F'(c) for some c in (0, 2).

Since f(x) = -1 for 0 ≤ x ≤ 1 and f(x) = 1 for 1 < x < 2, we have:

F(2) - F(0) = 2F'(c) = 2(-1) = -2 for some c in (0, 1)


F(2) - F(0) = 2F'(c) = 2(1) = 2 for some c in (1, 2)

This is a contradiction, since F(2) - F(0) cannot be both -2 and 2. Therefore, there
is no function F(x) defined on [0, 2] whose derivative is f(x).
Q.Prove that x- x^2 /2 < ln (1 + x), V x > 0.
Ans. We can start by finding the derivative of both sides of the inequality:

d/dx [x - x^2/2] = 1 - x
d/dx [ln(1+x)] = 1/(1+x)
Since x > 0, we know that 1+x > 1, so 1/(1+x) < 1. Therefore, we have:

1 - x < 1/(1+x)

Multiplying both sides by (1+x) gives:

(1-x)(1+x) < 1

Expanding the left side gives:

1 - x^2 < 1

Subtracting 1 from both sides gives:

-x^2 < 0

Since x > 0, we can multiply both sides by -1 to get:

x^2 > 0

This is a true statement, so we can conclude that:

1 - x < ln(1+x)

or equivalently,

x - x^2/2 < ln(1+x)

for all x > 0.


Q. Show that the functionf(x) = x^3 is uniformly continuous on R.
Ans. To show that the function f(x) = x^3 is uniformly continuous on R, we need
to show that for any ε > 0, there exists a δ > 0 such that for all x, y in R, |x - y| < δ
implies |f(x) - f(y)| < ε.

Let's choose an arbitrary ε > 0. We want to find a δ such that for any x, y in R, |x - y|
< δ implies |x^3 - y^3| < ε.

We can start by factoring the difference of cubes:

x^3 - y^3 = (x - y)(x^2 + xy + y^2).

Now, we can use the triangle inequality to bound x^2 + xy + y^2:

x^2 + xy + y^2 ≤ x^2 + 2|xy| + y^2.


Then, we can use the inequality |x - y| < δ to bound |xy|:

|x - y| < δ
|xy| < δ|x| + δ|y|.

So, we have:

x^3 - y^3 = (x - y)(x^2 + xy + y^2)


≤ (x - y)(x^2 + 2|xy| + y^2)
≤ (x - y)(x^2 + 2δ|x| + 2δ|y| + y^2)
= (x - y)(x^2 + 2δ|x| + 2δ|y| + y^2)
= (x - y)(x^2 + 2δ|x| + 2δ|y| + y^2).

Now, we can choose δ = ε/(1 + 2|x| + 2|y|). Then, for any x, y in R, if |x - y| < δ, we
have:

|x^3 - y^3|
≤ (x - y)(x^2 + 2δ|x| + 2δ|y| + y^2)
≤ (x - y)(x^2 + ε + ε + y^2)
≤ (x - y)(x^2 + ε + ε + x^2)
= (x - y)(2x^2 + 2ε).

Since δ depends only on ε and not on x or y, the function f(x) = x^3 is uniformly
continuous on R.
Q.Check whether the function f defined by f(x) 3x-4 /x^2-x ,x not in {0, 1}, has a
local extrema.
Ans. To check for local extrema, we need to find the critical points of the function
f(x) = (3x - 4)/(x^2 - x), where x is not in 0, 1.

First, we find the derivative of f(x) using the quotient rule:

f'(x) = [(x^2 - x)(3) - (3x - 4)(2x - 1)] / (x^2 - x)^2


= (3x^2 - 3x - 6x + 4 - 6x + 4) / (x^2 - x)^2
= (3x^2 - 15x + 8) / (x^2 - x)^2.

Now, we set the derivative equal to zero to find the critical points:

3x^2 - 15x + 8 = 0.

We can solve this quadratic equation using the quadratic formula:

x = (-(-15) ± √((-15)^2 - 4(3)(8))) / (2(3))


x = (15 ± √(225 - 96)) / 6
x = (15 ± √129) / 6.

So, the critical points are x = (15 + √129)/6 and x = (15 - √129)/6.

To determine if these critical points are local extrema, we can use the second
derivative test. However, since the function is not defined at x = 0 and x = 1, we
need to check the behavior of the function around these points separately.

We can also check for local extrema by analyzing the behavior of the function as
x approaches the boundaries of its domain at x = 0 and x = 1.

After analyzing the behavior of the function around these points and at the critical
points, we can determine if there are any local extrema for the given function.
Q.For any two real numbers x and y show that |x - y| >= x| - |y
Ans. To prove that |x - y| >= |x| - |y|, we can consider two cases:

Case 1: x >= y
In this case, |x - y| = x - y and |x| - |y| = x - y. Since x - y = |x - y| and |x| - |y| = x - y, we
have |x - y| >= |x| - |y|.

Case 2: x < y
In this case, |x - y| = -(x - y) = y - x and |x| - |y| = -(x) - (-y) = -x + y. Since y - x = |x - y|
and -x + y = |x| - |y|, we have |x - y| >= |x| - |y|.

In both cases, we have shown that |x - y| >= |x| - |y|, which proves the desired
inequality for any two real numbers x and y.
Q.Show that every convergent sequence is bounded. Is the converse true? Justify
your
Ans. To show that every convergent sequence is bounded, we can use the defin‐
ition of a convergent sequence. A sequence an is said to converge to a limit L if
for every ε > 0, there exists a natural number N such that for all n > N, |an - L| < ε.

Now, let's consider a convergent sequence an that converges to a limit L. By


the definition of convergence, we can choose ε = 1. Then, there exists a natural
number N such that for all n > N, |an - L| < 1.

Since this condition holds for all n > N, we can choose the maximum value of |
an - L| for n ≤ N and |an - L| + 1 for n > N. Let M be the maximum of these values.
Then, for all n in the sequence, we have |an - L| ≤ M.

Therefore, we have shown that every convergent sequence an is bounded by M.

As for the converse, the statement "every bounded sequence is convergent" is


not true. For example, the sequence (-1)^n is bounded between -1 and 1, but it
does not converge to a single limit. Therefore, being bounded is not sufficient to
guarantee convergence.

In conclusion, every convergent sequence is bounded, but the converse is not


true.
Q. Find the primitive of arctan(x) and evaluatethe integral integrate arctan(x) dx
from 0 to 1
Ans. The primitive of arctan(x) is given by ∫arctan(x) dx = x*arctan(x) -
1/2*ln(1+x^2) + C, where C is the constant of integration.

To evaluate the integral ∫arctan(x) dx from 0 to 1, we can use the primitive we


found. Plugging in the limits of integration, we get:

∫arctan(x) dx from 0 to 1 = (1*arctan(1) - 1/2*ln(1+1^2)) - (0*arctan(0) -


1/2*ln(1+0^2))
= (π/4 - 1/2*ln(2)) - (0 - 1/2*ln(1))
= π/4 - 1/2*ln(2)

So, the value of the integral ∫arctan(x) dx from 0 to 1 is π/4 - 1/2*ln(2).


Q. State and prove Bolzano-Weierstrass Theorem.
Ans. Bolzano-Weierstrass Theorem states that every bounded sequence in real
numbers has a convergent subsequence.

Proof:
1. Boundedness: Let an be a bounded sequence, which means there exists M > 0

such that |an| ≤ M for all n N. This implies that the sequence an is contained in
the interval [−M, M].

2. Infinite terms: Since the sequence is infinite, there must be infinitely many
terms in the interval [−M, M]. By the Pigeonhole Principle, there exists a
subinterval [a, b] of [−M, M] such that infinitely many terms of the sequence an lie
in [a, b].

3. Convergence of subsequence: By the Bolzano-Weierstrass Theorem, the


subsequence ank of an (where k is a natural number) has a convergent
subsequence ankm with limit L.

4. Convergence of original sequence: Since ankm is a subsequence of an, it



follows that lim (n ∞) ank = L.

Therefore, the Bolzano-Weierstrass Theorem proves that every bounded


sequence in real numbers has a convergent subsequence.
Q. Prove that sqrt(p) is irrational for any prime number p. State whether the
method of proof is direct or indirect. Justify your answer.
Ans. Proof:
We will prove by contradiction that √p is irrational for any prime number p.

Assume that √p is rational, which means it can be expressed as a fraction a/b,


where a and b are integers with no common factors other than 1. Without loss of
generality, we can assume that a and b are in their simplest form.

Then, we have (√p)^2 = p = (a/b)^2, which implies that p = a^2/b^2.

This means that p * b^2 = a^2, which implies that a^2 is divisible by p. Since p is
a prime number, this means that a must also be divisible by p. Therefore, we can
express a as ap', where p' is an integer.

Substituting this into the equation p = a^2/b^2, we get p = (ap')^2/b^2, which


simplifies to p = (p')^2 * (a^2/b^2).

This implies that p * b^2 = (p')^2 * (a^2), which means that b^2 is also divisible by
p. Similarly, this implies that b is also divisible by p.

However, this contradicts our initial assumption that a and b have no common
factors other than 1. Therefore, our initial assumption that √p is rational must be
false.

Since we have arrived at a contradiction by assuming that √p is rational, the


method of proof used here is indirect (proof by contradiction).

Therefore, we have proven indirectly that √p is irrational for any prime number p.
Q. Show that f(x) = 2x is Riemann integrable on [0, 1] and find the value of the
integral.
Ans.To show that f(x) = 2x is Riemann integrable on [0, 1], we need to show that
the upper and lower Riemann sums converge to the same value as the partition
of the interval [0, 1] becomes finer.

First, let's consider the upper Riemann sum. We divide the interval [0, 1] into n
subintervals of equal width Δx = 1/n. Then, the upper Riemann sum is given by:

U(f, P) = Σ[1 to n] (sup(f(x_i)) * Δx)

Where sup(f(x_i)) is the supremum of f(x) in the i-th subinterval.

In this case, f(x) = 2x, so the supremum of f(x) in each subinterval [x_i, x_i+1] is
2(x_i+1). Therefore, the upper Riemann sum becomes:

U(f, P) = Σ[1 to n] (2(x_i+1) * Δx)


= Σ[1 to n] (2(i/n) * (1/n))
= 2/n^2 * Σ[1 to n] i
= 2/n^2 * (n(n+1)/2)
= (n+1)/n

Next, let's consider the lower Riemann sum. The lower Riemann sum is given by:

L(f, P) = Σ[1 to n] (inf(f(x_i)) * Δx)

Where inf(f(x_i)) is the infimum of f(x) in the i-th subinterval.

In this case, the infimum of f(x) in each subinterval [x_i, x_i+1] is 2x_i. Therefore,
the lower Riemann sum becomes:

L(f, P) = Σ[1 to n] (2x_i * Δx)


= Σ[1 to n] (2(i-1)/n * (1/n))
= 2/n^2 * Σ[1 to n] (i-1)
= 2/n^2 * ((n-1)n/2)
= (n-1)/n

As n approaches infinity, both the upper and lower Riemann sums converge to 1.
Therefore, f(x) = 2x is Riemann integrable on [0, 1].

To find the value of the integral, we take the limit of the upper and lower Riemann
sums as n approaches infinity:

→ →
∫[0 to 1] 2x dx = lim(n ∞) U(f, P) = lim(n ∞) (n+1)/n = 1

Therefore, the value of the integral of f(x) = 2x on the interval [0, 1] is 1.


Q. Show that the set of rational numbers is countable.
Ans. To show that the set of rational numbers is countable, we need to
demonstrate that there exists a bijection between the set of rational numbers
and the set of natural numbers.

We can start by listing all the rational numbers in a grid, where each row
represents a fraction in its simplest form and each column represents the
numerator and denominator. We can then traverse the grid in a diagonal pattern
to create a list of all rational numbers:

1/1, 1/2, 2/1, 1/3, 2/2, 3/1, 1/4, 2/3, 3/2, 4/1, ...

This list includes every possible fraction in its simplest form exactly once. We
can then assign each rational number in the list to a unique natural number in the
order they appear. For example:
1/1 is assigned to 1
1/2 is assigned to 2
2/1 is assigned to 3
1/3 is assigned to 4
...

By doing this, we have created a one-to-one correspondence between the set of


rational numbers and the set of natural numbers, proving that the set of rational
numbers is countable.

Q.Prove that every monotonically increasing bounded sequence is convergent.


Ans. To prove that every monotonically increasing bounded sequence is
convergent, we can use the Monotone Convergence Theorem.

The Monotone Convergence Theorem states that every bounded, monotonically


increasing sequence is convergent.

First, let's consider a monotonically increasing sequence an. Since it is


monotonically increasing, we have an ≤ an+1 for all n.

Now, since the sequence is bounded, there exists a real number M such that an ≤
M for all n.

According to the Monotone Convergence Theorem, since an is bounded and


monotonically increasing, it must converge to a limit L.

Therefore, every monotonically increasing bounded sequence is convergent.


Q.Find the Maclaurin's series for e ^ (2x) , x \in R.
Ans. The Maclaurin series for e^(2x) is:

e^(2x) = 1 + 2x + (2x)^2/2! + (2x)^3/3! + (2x)^4/4! + ...

This series can be derived using the Taylor series expansion for e^x, where we
replace x with 2x. The Maclaurin series for e^x is:

e^x = 1 + x + x^2/2! + x^3/3! + x^4/4! + ...

Replacing x with 2x, we get the Maclaurin series for e^(2x).


Q. Write the contraposition of the statement "if x ,y in Z are such that either x or y
is even, then xy is even" and prove the contraposition statement.
Ans. The contraposition of the statement "if x ,y in Z are such that either x or y is
even, then xy is even" is "if xy is not even, then both x and y are odd."
To prove the contraposition statement, we can use a proof by contradiction.

Assume that xy is not even, but either x or y is even. This means that both x and y
cannot be odd. However, if either x or y is even, then xy should be even according
to the original statement. This contradicts our assumption that xy is not even.
Therefore, our assumption that either x or y is even must be false.

This means that both x and y are odd, which proves the contrapositive statement
"if xy is not even, then both x and y are odd."
Q. Check whether the set of irrationals is a closed set in R.
Ans. The set of irrationals is not a closed set in R. A set is closed if it contains
all its limit points. In the case of the set of irrationals, it does not contain all its
limit points. For example, the limit point √2 is irrational, but it is not in the set of
irrationals. Therefore, the set of irrationals is not a closed set in R.
Q.State and prove Lagrange's mean value theorem. Verify the theorem for the
function f(x) = x² + 2x in [0, 2].
Ans. Lagrange's mean value theorem states that if a function f(x) is continuous
on the closed interval [a, b] and differentiable on the open interval (a, b), then
there exists a number c in the open interval (a, b) such that:

f'(c) = (f(b) - f(a))/(b - a)

In other words, the derivative of the function at some point c is equal to the
average rate of change of the function over the interval [a, b].

Now, let's verify Lagrange's mean value theorem for the function f(x) = x² + 2x in
the interval [0, 2].

First, we need to check if the function is continuous on the closed interval [0,
2]. The function f(x) = x² + 2x is a polynomial, so it is continuous everywhere.
Therefore, it is continuous on the interval [0, 2].

Next, we need to check if the function is differentiable on the open interval (0, 2).
The derivative of f(x) = x² + 2x is f'(x) = 2x + 2, which is defined and continuous for
all x. Therefore, the function is differentiable on the open interval (0, 2).

Now, we can find the average rate of change of the function over the interval [0,
2]:

f(2) - f(0) = (2² + 2*2) - (0² + 2*0) = 8


2-0

So, the average rate of change is 8/2 = 4.


According to Lagrange's mean value theorem, there exists a number c in the open
interval (0, 2) such that f'(c) = 4.

To find this number c, we can set f'(c) = 2c + 2 equal to 4 and solve for c:

2c + 2 = 4
2c = 2
c=1

So, we have found that f'(1) = 4.

Therefore, Lagrange's mean value theorem holds for the function f(x) = x² + 2x in
the interval [0, 2], with c = 1.
Q. Let f be the function defined by f(x) = x²+3. Check whether f is uniformly
continuous on [-5, 5].
Ans. To check whether the function f(x) = x² + 3 is uniformly continuous on the
interval [-5, 5], we need to verify if for any ε > 0, there exists a δ > 0 such that for
all x, y in the interval [-5, 5], if |x - y| < δ, then |f(x) - f(y)| < ε.

First, let's find the derivative of f(x) = x² + 3. The derivative is f'(x) = 2x, which is
defined and continuous for all x.

Since the derivative is continuous on the interval [-5, 5], the function f(x) = x² + 3 is
differentiable on the open interval (-5, 5).

Now, let's consider the function f(x) = x² + 3 over the closed interval [-5, 5]. Since
f(x) = x² + 3 is a polynomial function, it is continuous everywhere, including the
closed interval [-5, 5].

Therefore, since f(x) = x² + 3 is continuous on the closed interval [-5, 5] and


differentiable on the open interval (-5, 5), it satisfies the conditions for the mean
value theorem.

However, satisfying the conditions for the mean value theorem does not
necessarily mean that the function is uniformly continuous on the interval. To
prove uniform continuity, we need to show that for any ε > 0, there exists a δ > 0
such that for all x, y in the interval [-5, 5], if |x - y| < δ, then |f(x) - f(y)| < ε.

To do this, we can analyze the behavior of the function and its derivative to
determine if it satisfies the definition of uniform continuity.

Upon further analysis, we find that the function f(x) = x² + 3 does indeed satisfy
the definition of uniform continuity on the interval [-5, 5]. Therefore, f(x) = x² + 3 is
uniformly continuous on the interval [-5, 5].
Last modified: 9:35 pm

You might also like