Handouts RealAnalysis II
Handouts RealAnalysis II
Handouts RealAnalysis II
Pakistan
Dr. Malik has published several research articles in international journals and
conferences. His area of research includes the study of dierential equations with
nonlocal operators and their applications to image processing. He is also interested
in inverse problems related to reaction-diusion equations with nonlocal integro-
dierential operators and boundary conditions. These models have numerous appli-
cations in anomalous diusion/transport, biomedical imaging and non-destructive
testing.
Course Information
Title and Course Code: Real Analysis II (MTH631)
Course Objective: Real Analysis II is the follow up course of Real Analysis I and in
general an advanced course related to mathematical analysis. The topics of the Real
Analysis II are linked with its rst course namely Real Analysis I, indeed, we will
extend the ideas of Real Analysis I to Euclidean space Rn , we will discuss sequences
and series of functions, limits and continuity of functions of several variables, partial
derivatives their applications, multiple integrals etc. Upon completion of this course
students will be able to
• Apply Dirichlet's test for uniform convergence, series of product of two func-
tions, interchange of sum and intgeration (LO5).
• Represent and study the function which could be written as power series,term
by term integral and derivative of a power series, (LO6). item Understand the
concept of equicontinuous function, The Stone-Weierstrass Theorem (LO7).
• Apply the best approximation theorem and understand the Euler gamma func-
tion and the beta function and their properties (LO9)
• Vector valued functions and their calculus, Bounded functions and several
results about vector valued functions (LO11)
Reference books:
[4] A. N. Kolmogorov and S. V. Fomin, Introductory Real Analysis, Revised English
Edition Translated and Edited by R. A. Silverman, Dover Publication, Inc. New
York.
• Integral Calculus
Contents
1 Sequences and Series of Functions 2
1.1 Informal way . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Pointwise Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Norm Dened Over a Set . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Uniform Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Cauchy's Uniform Convergence Criterion . . . . . . . . . . . . . . . . 8
1.6 Properties Preserved by Uniform Convergence . . . . . . . . . . . . . 10
1.6.1 Continuity of the Limit Function at a Point . . . . . . . . . . 10
1.6.2 Interchange of Limit and Integration . . . . . . . . . . . . . . 12
1.6.3 Under What Conditions We May Have F ′ = limn→∞ Fn′ . . . 14
1.7 Series of Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.8 Convergence of Series of Functions . . . . . . . . . . . . . . . . . . . 15
1.8.1 Cauchy's criterion for functional series . . . . . . . . . . . . . 17
1.8.2 Dominated Series of Real Numbers for Series of Functions . . 18
1.8.3 Weierstrass M-test/dominated Convergence Test . . . . . . . 18
1.9 Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.10 The Taylor's Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
1.11 The Abel's Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
1.12 Pointwise and Uniform Bounded Functions . . . . . . . . . . . . . . . 47
1.13 Equicontinuous Functions on a Set . . . . . . . . . . . . . . . . . . . 48
1.14 The Stone-Weierstrass Theorem . . . . . . . . . . . . . . . . . . . . . 51
1.15 Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
1.15.1 Periodic Functions . . . . . . . . . . . . . . . . . . . . . . . . 53
1.15.2 Periodic Extension . . . . . . . . . . . . . . . . . . . . . . . . 55
1.15.3 Trigonometric Polynomials . . . . . . . . . . . . . . . . . . . 56
1.16 The space E . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
1.16.1 Fourier Series of Even and Odd Functions . . . . . . . . . . . 61
1.17 Fourier Series for Arbitrary Periodic Function . . . . . . . . . . . . . 66
1.18 Best Approximation Theorem . . . . . . . . . . . . . . . . . . . . . . 69
If the sequence of values {Fn (x)} converges for each x in some subset S of D, then
{Fn } denes a limit function on S.
( )
Figure 1.1: Plot of Fn (x) = 1
n+x , n ≥ 1, for n = 1, 2, 4, 8, 100
( )
Figure 1.2: Plot of Fn (x) = x
n+x , n ≥ 1, for n = 1, 2, 4, 8, 20, 100
( )
nx n/2
Fn (x) = 1 − , n ≥ 1,
n+1
∞, x < 0,
lim Fn (x) = 1, x = 0,
n→∞
0, 0 < x ≤ 1.
{
1, x = 0,
F (x) =
0, 0 < x ≤ 1.
Fn (x) = xn e−nx , x ≥ 0, n ≥ 1.
1.2. Pointwise Convergence 4
( )
1
Fn (x) = , n ≥ 1,
n+x
Example: For each positive integer n, let Sn be the set of numbers of the form
x = p/q , where p and q are integers with no common factors and 1 ≤ q ≤ n.
Dene {
1, x ∈ Sn ,
Fn (x) =
0, x ∈
̸ Sn .
Ifx is irrational, then x ̸∈ Sn for any n, so Fn (x) = 0, n ≥ 1. If x is rational, then
x ∈ Sn and Fn (x) = 1 for all suciently large n.
Therefore,
{
1 if x is rational,
lim Fn (x) = F (x) =
n→∞ 0 if x is irrational.
1.3. Norm Dened Over a Set 5
y =F (x) +
y =F (x)
y =F (x) −
x
a b
If S = [a, b] and F is the function with graph shown in then (1.2) implies that the
graph of
y = Fn (x), a ≤ x ≤ b,
lies in the shaded band
.
Example: The sequence {Fn } dened by
Fn (x) = xn e−nx , n ≥ 1,
converges uniformly to F ≡ 0.
We have
∥Fn − F ∥S = ∥Fn ∥S = e−n ,
so
∥Fn − F ∥S < ε
if n > − log ε. For these values of n, the graph of
y = Fn (x), 0 ≤ x < ∞,
1. {Fn } converges pointwise to F on S if and only if there is, for each ε > 0 and
x ∈ S , an integer N (which may depend on x as well as ε) such that
The converse is false; that is, pointwise convergence does not imply uniform conver-
gence.
1.4. Uniform Convergence 7
The sequence {Fn } of converges pointwise to F ≡ 0 on (−∞, ∞), but not uniformly.
Because ( ) ( )
1 −1
∥Fn − F ∥(−∞,∞) = Fn = Fn = n,
n n
so
lim ∥Fn − F ∥(−∞,∞) = ∞.
n→∞
Sρ = (−∞, ρ] ∪ [ρ, ∞)
The convergence is not uniform on S. To see this, suppose that 0 < ε < 1. Then
Therefore,
1 − ε ≤ ∥Fn − F ∥S ≤ 1
for all n ≥ 1. Since ε can be arbitrarily small, it follows that
{
1, x = 1,
F (x) =
0, 0 ≤ x < 1.
∥Fn − F ∥[0,ρ] = ρn
and limn→∞ ρn = 0. Another way to say the same thing: {Fn } converges uniformly
on every closed subset of [0, 1).
∥Fm − F ∥S ≤ 2ε if m ≥ N.
Since ε is an arbitrary positive number, this implies that {Fn } converges uniformly
to F on S.
Solution: We rst note that if u and v are any two real numbers, then (1.5) and
the mean value theorem imply that
Recalling (1.6) and applying this inequality with u = Fn−1 (x) and v = 0 shows that
so
∥Fn+1 − Fn ∥S ≤ r∥Fn − Fn−1 ∥S .
By induction, this implies that
If n > m, then
rN
Therefore, if ∥F1 − F0 ∥S < ε,
1−r
then ∥Fn − Fm ∥S < ε if n, m ≥ N .
|F (x) − F (x0 )| ≤ |F (x) − Fn (x)| + |Fn (x) − Fn (x0 )| + |Fn (x0 ) − F (x0 )|
≤ |Fn (x) − Fn (x0 )| + 2∥Fn − F ∥S .
(1.9)
1.6. Properties Preserved by Uniform Convergence 11
∫ b ∫ b
F (x) dx = lim Fn (x) dx,
a n→∞ a
is true?
∫b ∫b
Example: a F (x) dx = limn→∞ a Fn (x) dx, is not true generally.
Then the sequence {Fn } converges pointwise to F (x) = 0 on [0, 1] and it is not
uniformly convergent. We have
∫ 1 ∫ 1/n ∫ 1 ∫ 1
Fn (x) dx = n dx + 0dx = 1 But F (x) dx = 0
0 0 1/n 0
∫ b ∫ b
lim Fn (x) dx ̸= lim Fn (x) dx,
a n→∞ n→∞ a
1.6. Properties Preserved by Uniform Convergence 12
Proof : Consider ∫ b ∫ b
Fn (x) dx − F (x) dx
a a
∫ b ∫ b ∫ b
Fn (x) dx − F (x) dx ≤ |Fn (x) − F (x)| dx
a a a
≤ (b − a)∥Fn − F ∥S
Theorem: Suppose that {Fn } converges uniformly to F on S = [a, b]. Assume that
F and all Fn are integrable on [a, b].
Then ∫ b ∫ b
F (x) dx = lim Fn (x) dx.
a n→∞ a
∫ b ∫ b
F (x) dx = lim Fn (x) dx.
a n→∞ a
holds.
2. If the sequence {∥Fn ∥[a,b] } is bounded and F is integrable on [a, b], then
∫ b ∫ b
F (x) dx = lim Fn (x) dx.
a n→∞ a
holds.
1.6. Properties Preserved by Uniform Convergence 13
Remark: Part (1) of this theorem shows that it is not necessary to assume that F
is integrable on [a, b], since this follows from the uniform convergence. Part (2) is
known as the bounded convergence theorem. Neither of the assumptions of (2) can
be omitted.
In this example it is clear that ∥Fn ∥[a,b] = 1 for every nite interval [a, b], Fn is
integrable for all n ≥ 1, and F is nonintegrable on every interval.
1
Fn (x) = xn sin .
xn−1
The sequence of functions converges {Fn } converges uniformly to F ≡0 on [r1 , r2 ]
if 0 < r1 < r2 < 1 (or, equivalently, on every compact subset of (0, 1)).
However,
1 1
Fn′ (x) = nxn−1 sin − (n − 1) cos ,
xn−1 xn−1
so {Fn′ (x)} does not converge for any x in (0, 1).
1.6. Properties Preserved by Uniform Convergence 14
while
F+′ (a) = lim Fn′ (a+) and F−′ (b) = lim Fn′ (b−). (1.13)
n→∞ n→∞
Proof : Since Fn′ is continuous on [a, b], due to fundamental theorem of calculus, we
can write ∫ x
Fn (x) = Fn (x0 ) + Fn′ (t) dt, a ≤ x ≤ b. (1.14)
x0
Since Fn′ is continuous and {Fn′ } converges uniformly to G on [a, b], G is contin-
uous on [a, b].
∫b
Therefore,
∫b
(1.14) and using the fact we have proved
a F (x) dx =
limn→∞ Fn (x) dx (with F and Fn replaced by G ′
and Fn ) imply that {Fn } con-
a
verges pointwise on [a, b] to the limit function
∫ x
F (x) = L + G(t) dt.
x0
∫ x
F (x) = L + G(t) dt. (1.16)
x0
The convergence is actually uniform on [a, b], since subtracting (1.14) from (1.16)
yields
∫
x
|F (x) − Fn (x)| ≤ |L − Fn (x0 )| + |G(t) − Fn′ (t)| dt
x0
≤ |L − Fn (x0 )| + |x − x0 | ∥G − Fn′ ∥[a,b] .
Consequently,
∑
n
Fn = fj , n ≥ k.
j=k
If{Fn }∞
k converges pointwise to a function F on a subset S of D, we say that
∑∞
f
j=k j converges pointwise to the sum F on S , and write
∞
∑
F = fj , x ∈ S.
j=k
∑∞
{Fn } converges uniformly to F
If on S , we say that j=k fj converges uniformly
to F on S .
∑
n
Fn = fj , n ≥ k.
j=k
fj (x) = xj , j ≥ 0,
∞
∑
xj
j=0
on D = (−∞, ∞).
Fn (x) = 1 + x + x2 + · · · + xn ,
1
F (x) =
1−x
if |x| < 1 and diverges if |x| ≥ 1.
Hence, we write
∞
∑ 1
xj = , −1 < x < 1.
1−x
j=0
∥F − Fn ∥(−1,1) = ∞,
1
F (x) =
1−x
if |x| < 1 and diverges if |x| ≥ 1.
Neither is it uniform on any interval (−1, r] with −1 < r < 1, since
1
∥F − Fn ∥(−1,r) ≥
2
for every n on every such interval.
Uniform convergence: The series does converge uniformly on any interval [−r, r]
with 0 < r < 1, since
rn+1
∥F − Fn ∥[−r,r] =
1−r
and limn→∞ rn = 0. Put another way, the series converges uniformly on closed
subsets of (−1, 1).
Uniform convergence (using ε): See video lectures.
∑∞
Remark: A necessary condition for j=0 fj (x) to converge on S is that fj (x) → 0
for each x ∈ S.
∑
Theorem: A series fn converges uniformly on a set S if and only if for each
ε>0 there is an integer N such that
∑
Proof : Apply Cauchy's convergence criterion to the partial sums of fn , observing
that
fn + fn+1 + · · · + fm = Fm − Fn−1 .
∑
Theorem: A series fn converges uniformly on a set S if and only if for each
ε>0 there is an integer N such that
∑
Corollary: If fn converges uniformly on S, then limn→∞ ∥fn ∥S = 0. Setting
m = n.
∑∞
Example: We have proved that the series j=0 fj (x), where
fj (x) = xj , j ≥ 0,
is uniformly convergent on any compact subset of (−1, 1) say [−r, r], where 0<r<
1.
Let us apply Cauchy's criterion for functional series, recall that we have
1 − xn+1
Fn (x) = 1 + x + x2 + ... + xn = .
1−x
Consider
We have
2|rn+1 |
∥Fm − Fn ∥[−r,r] ≤ .
1 − |r|
Since
2|rn+1 |
→0 as n → ∞,
1 − |r|
there is an integer N (ε) can be found for which
2|rn+1 |
< ε, when n > N (ε).
1 − |r|
We have
∥Fm − Fn ∥[−r,r] ≤ ε,
∑∞ j
hence by Cauchy's criterion the series j=0 x , is uniformly convergent on [−r, r].
n=1 Mn .
∑∞
Example: Consider
∑
Fn = 1
x2 +n2
and the series of functions n=1 Fn is dominated
by the series 1/n2 because
1
|Fn | < =: Mn .
n2
∑
We know that 1/n2 < ∞.
∥fn ∥S ≤ Mn , n ≥ k, (1.20)
∑
where Mn < ∞.
Proof : From Cauchy's convergence criterion for series of constants, there is for each
ε>0 an integer N such that
Mn + Mn+1 + · · · + Mm < ε if m ≥ n ≥ N.
∑
• 1
x2 +n2
.
∑ sin nx
• n2
.
Solution: We have
1 1 sin nx 1
≤ 2,
x2 + n2 n n2 ≤ n2 .
∑ 1
< ∞.
n2
Due to Weierstrass M-test, we can conclude
∑ 1 ∑ sin nx
and
x + n2
2 n2
converge uniformly on (−∞, ∞).
∑ ∑ ( x )n
fn (x) = .
1+x
Solution: The given series converges uniformly on any set S such that
x
1 + x ≤ r < 1, x ∈ S. (1.21)
−r r
≤x≤ , x ∈ S,
1+r 1−r
this means that the series converges uniformly on any compact subset of (−1/2, ∞).
∑ ∑ ( x )n
fn (x) = .
1+x
1.8. Convergence of Series of Functions 20
∑
Recall: If fn converges uniformly on S, then limn→∞ ∥fn ∥S = 0. The series
does not converge uniformly on S = (−1/2, b) with b < ∞ or on S = [a, ∞) with
a > −1/2, because in these cases ∥fn ∥S = 1 for all n.
∑
Absolute
∑
convergence: A series of functions fn is said toconverge absolutely
on S
∑
if |fn | converges pointwise on S, and converges absolutely uniformly on S if
|fn | converges uniformly on S.
Remarks:
• The condition of absolutely convergence (pointwise or uniform) is stronger
than the usual convergence (pointwise or uniform).
∑
• In our proof of Weierstrass's M-test, we actually proved that fn converges
absolutely uniformly on S.
Proof : Let
Gn = gk + gk+1 + · · · + gn ,
∑∞
and consider the partial sums of n=k fn gn :
Which we rewrite as
or
Hn = Jn−1 + fn Gn , (1.24)
where
Jn−1 = (fk − fk+1 )Gk + (fk+1 − fk+2 )Gk+1 + · · · + (fn−1 − fn )Gn−1 . (1.25)
∞
∑
(fj − fj+1 )Gj . (1.26)
j=k
so
∑
∑
m
m
(fj − fj+1 )Gj
≤ M
|fj − fj+1 |
.
j=n
j=n
S S
∑
Now suppose that ε > 0. Since (fj − fj+1 ) converges absolutely uniformly on S ,
Cauchy's convergence criterion implies that there is an integer N such that the right
side of the last inequality is less than ε if m ≥ n ≥ N . The same is then true of the
left side, so Cauchy's convergence criterion implies that (1.26) converges uniformly
on S.
We have now shown that {Jn } as dened in (1.25) converges uniformly to a limit
function J on S. Returning to (1.24), we see that
Hn − J = Jn−1 − J + fn Gn .
Hence, we have
∥gk + gk+1 + · · · + gn ∥S ≤ M, n ≥ k,
1.8. Convergence of Series of Functions 22
We have
1
|Gn (x)| ≤ , n ≥ 1, n ̸= 2kπ (k = integer).
| sin(x/2)|
Therefore, {∥Gn ∥S } is bounded, and the series converges uniformly on any set S on
which sin x/2 is bounded away from zero.
Since
∑ sin nx
n = ∞, x ̸= kπ.
1
fn (x) = , gn = (−1)n , G2m = 0, and G2m+1 = −1.
n + x2
Therefore, the series converges uniformly on (−∞, ∞). This result cannot be ob-
tained by Weierstrass's test, since
∑ 1
=∞
n + x2
for all x.
∑∞
Theorem: If n=k fn converges uniformly to F on S and each fn is continuous at
a point x0 in S, then so is F. Similar statements hold for continuity from the right
and left.
∑∞
Recall the following: Theorem: If n=k fn converges uniformly to F on S and
each fn is continuous at a point x0 in S, then so is F. Similar statements hold for
continuity from the right and left.
∞ (
∑ )n
x
F (x) =
1+x
n=0
∞
∑ sin nx
G(x) =
n
n=1
∑∞ sin nx
We have seen that the series n=1 is uniformly convergent by applying Dirich-
n
let's Test for Uniform Convergence except at xk = 2kπ (k = integer).
• F is bounded, because
ε
S(P, Fn ) − s(P, Fn ) < .
3
For each x ∈ [a, b] with n=N
ε
|Fn (x) − F (x)| < , for all x ∈ [a, b], n > N,
3(b − a)
implies that
ε ε
Fn (x) − < F (x) < Fn (x) + .
3(b − a) 3(b − a)
Therefore,
ε ε
s(P, Fn ) − < s(P, F ) ≤ S(P, F ) < S(P, Fn ) +
3 3
Hence F is integrable. Finally, for n ≥ N and for each t ∈ [a, b], we have
∫ t ∫t ∫t
Fn (x)dx − F (x)dx ≤ |Fn (x) − F (x)|dx
a a a
ε(b − a)
≤ , for all x ∈ [a, b], n > N.
3(b − a)
1.8. Convergence of Series of Functions 25
∑∞
Theorem: Suppose that n=k fn converges uniformly to F on S = [a, b]. Assume
that F and fn , n ≥ k, are integrable on [a, b]. Then
∫ b ∞ ∫
∑ b
F (x) dx = fn (x) dx.
a n=k a
∑∞
We say in this case that n=k fn can be integrated term by term over [a, b].
Example: Consider the {Fn } dened by
x
Fn (x) = , x ∈ [a, b] ⊂ R.
1 + nx2
∑
Then Weieretrass's M-test shows that Fn converges uniformly on [a, b]
∑ ∞
1
= xn , −1 < x < 1.
1−x
n=0
The series converges uniformly, and the limit function is integrable on any closed
subinterval [a, b] of (−1, 1).
Hence,
∫ b ∑ ∞ ∫ b
dx
= xn dx.
a 1−x
n=0 a
Consequently,
∞ n+1
∑ b − an+1
log(1 − a) − log(1 − b) = .
n+1
n=0
∞ n+1
∑ b − an+1
log(1 − a) − log(1 − b) = .
n+1
n=0
1.8. Convergence of Series of Functions 26
∞
∑ xn+1
log(1 − x) = − , −1 < x < 1.
n+1
n=0
∑
n
Fn = x(ex − 1)e−kx dx.
k=1
e−x (1 − e−nx )
Fn (x) = x(ex − 1) ,
1 − e−x
Solution: For the function xe−nx , we have seen that it attains its maximum at
x = 1/n, we have
∥Fn (x) − x∥ = sup |Fn (x) − x|
x≥0
1
∥Fn (x) − x∥ = sup |xe−nx | = .
x≥0 en
So, as n → ∞, we have ∥Fn (x) − x∥ → 0.
∞
∑
x(ex − 1)e−nx dx,
n=1
1.8. Convergence of Series of Functions 27
Example: Consider
x
Fn (x) = , x ∈ R.
1 + nx2
x |x| 1
|Fn (x)| = ≤ √ = √ .
1 + nx2 2 n|x| 2 n
Fn (x) is uniformly convergent to F (x) = 0 on R. We have
1 − nx2
Fn′ (x) =
(1 + nx2 )2
When x = 0, we have limn→∞ Fn′ (x) = 0 and for x ̸= 0 limn→∞ Fn′ (x) = 1.
• Fn converges uniformly to F on S.
• F is dierentiable on S.
• There exists x∈S with F ′ (x) ̸= limn→∞ Fn′ (x), because Fn′ (0) → 1 ̸= F ′ (0).
∫x
g(t)dt = lim (fn (x) − fn (x0 )).
n→∞
x0
∫x
g(t)dt + lim fn (x0 ) = lim fn (x), on [a, b].
n→∞ n→∞
x0
∫x
g(t)dt + lim fn (x0 ) = F (x), on [a, b].
n→∞
x0
∑∞
Remark: The series n=k fn can be dierentiated term by term on [a, b].
∑∞
• Then dierentiate n=k fn term by term. If the resulting series converges
uniformly. Then term by term dierentiation was legitimate.
∞
∑ 1 x
(−1)n+1 2
sin (1.28)
n n
n=1
∞
∑ xn x2 x3
E(x) = =1+x+ + + ··· . (1.29)
n! 2! 3!
n=0
The series converges uniformly on every interval [−r, r] by Weierstrass's test, because
|x|n rn
≤ , |x| ≤ r,
n! n!
∑ rn
< ∞
n!
for all r, by the ratio test.
Dierentiating the right side of (1.30) term by term yields the series
∞
∑ ∞
∑
xn−1 xn
= ,
(n − 1)! n!
n=1 n=0
∞
∑ xn x2 x3
E(x) = =1+x+ + + ··· . (1.30)
n! 2! 3!
n=0
1.9. Power Series 30
Therefore, the dierentiated series is also uniformly convergent on [−r, r] for every
r, so the term by term dierentiation is legitimate and
Remark: Failure to verify that the given series converges at some point can lead
to erroneous conclusions.
∞
∑ x
cos (1.31)
n
n=1
term by term.
We have
∞
∑ 1 x
− sin .
n n
n=1
Since
1
sin ≤ |x| ≤ r ,
x
|x| ≤ r,
n n n2 n2
∑
and 1/n2 < ∞. which converges uniformly on [−r, r] for every r,
We cannot conclude from this that (1.31) converges uniformly on [−r, r]. In fact,
it diverges for every x.
∞
∑
an (x − x0 )n , (1.32)
n=0
∑
Theorem: The radius of convergence of an (x − x0 )n is given by
1 an+1
= lim
R n→∞ an
∞
∑
f (x) = an (x − x0 )n
n=0
∞
∑
f ′ (x) = (n + 1)an+1 (x − x0 )n . (1.35)
n=0
Proof : Since
∞
∑
f ′ (x) = (n + 1)an+1 (x − x0 )n ,
n=0
∞
∑
f ′ (x) = (n + 1)an+1 (x − x0 )n ,
n=0
∞
∑
f (x) = an (x − x0 )n
n=0
with positive radius of convergence R has derivatives of all orders in its interval of
convergence, which can be obtained by repeated term by term dierentiation. That
is,
∞
∑
f (k) (x) = n(n − 1) · · · (n − k + 1)an (x − x0 )n−k . (1.36)
n=k
Proof : ∞
∑
f (k)
(x) = n(n − 1) · · · (n − k + 1)an (x − x0 )n−k .
n=k
∞
∑
f (k) (x) = (n + k)(n + k − 1) · · · (n + 1)an+k (x − x0 )n , |x − x0 | < R.
n=0
Dening
bn = (n + k)(n + k − 1) · · · (n + 1)an+k . (1.37)
We rewrite this as
∞
∑
f (k)
(x) = bn (x − x0 )n , |x − x0 | < R.
n=0
∞
∑
f (k+1)
(x) = nbn (x − x0 )n−1 , |x − x0 | < R.
n=1
∞
∑
f (k+1)
(x) = (n + k)(n + k − 1) · · · (n + 1)nan+k (x − x0 )n−1 .
n=1
∞
∑
f (k+1)
(x) = n(n − 1) · · · (n − k)an (x − x0 )n−k−1 , |x − x0 | < R,
n=k+1
1.9. Power Series 33
∑ ∞
1
= xn , |x| < 1.
1−x
n=0
∞
∑
k!
= n(n − 1) · · · (n − k + 1)xn−k
(1 − x)k+1
n=k
∑∞
= (n + k)(n + k − 1) · · · (n + 1)xn , |x| < 1,
n=0
∑∞ ( )
1 n+k n
= x , |x| < 1.
(1 − x)k+1 k
n=0
∞
∑ ∞
∑
x2n+1
n x2n
S(x) = (−1) and C(x) = (−1)n
(2n + 1)! (2n)!
n=0 n=0
∞
∑
′ xn
S (x) = (−1)n = C(x)
(2n)!
n=0
and
∞
∑ ∑ ∞
′ x2n−1 x2n+1
C (x) = (−1) n
=− (−1)n = −S(x).
(2n − 1)! (2n + 1)!
n=1 n=0
Theorem: If
∞
∑
f (x) = an (x − x0 )n , |x − x0 | < R,
n=0
then
f (n) (x0 )
an = .
n!
1.9. Power Series 34
Proof : We have
∞
∑
f (k)
(x) = n(n − 1) · · · (n − k + 1)an (x − x0 )n−k .
n=k
Theorem: If
∞ ∞
∑ ∑
an (x − x0 )n = bn (x − x0 )n (1.38)
n=0 n=0
an = bn , n ≥ 0. (1.39)
∑∞ ∑∞
Proof : Let f (x) = n=0 an (x − x0 )n and g(x) = n=0 bn (x − x0 )n .
From previous result, we have
Theorem (Recall the following): For the power series, dene R in the extended
real numbers by
1
= lim sup |an |1/n . (1.41)
R n→∞
In particular, R = 0 if lim supn→∞ |an |1/n = ∞, and R = ∞ if lim supn→∞ |an |1/n =
0.
Then the power series converges
1. only for x = x0 if R = 0;
∑∞
Theorem (Recall the following): Suppose that n=k fn converges uniformly to F
on S = [a, b]. Assume that F and fn , n ≥ k, are integrable on [a, b].
Then
∫ b ∞ ∫
∑ b
F (x) dx = fn (x) dx.
a n=k a
∞
∑
f (x) = an (x − x0 )n ,
n=0
then
∫ ∞
∑
x2
an [ ]
f (x) dx = (x2 − x0 )n+1 − (x1 − x0 )n+1 ;
x1 n+1
n=0
that is, a power series may be integrated term by term between any two points in
its interval of convergence.
∞
∑
f (x) = an (x − x0 )n
n=0
with positive radius of convergence R has derivatives of all orders in its interval of
convergence, which can be obtained by repeated term by term dierentiation; thus,
∞
∑
f (k) (x) = n(n − 1) · · · (n − k + 1)an (x − x0 )n−k . (1.42)
n=k
1.10. The Taylor's Series 36
∞
∑
f (x) = an (x − x0 )n , |x − x0 | < R,
n=0
then
f (n) (x0 )
an = .
n!
∑
m
f (n) (x0 )
Tm (x) = (x − x0 )n ,
n!
n=0
f (x) =
0, x = 0.
the function f is innitely times dierentiable on (−∞, ∞) and f (n) (0) = 0 for
n ≥ 0. So its Maclaurin series is identically zero.
f (n+1) (cn )
f (x) − Tn (x) = (x − x0 )n−1 , (1.44)
(n + 1)!
where cn is between x and x0 .
Therefore,
∞
∑ f (n) (x0 )
f (x) = (x − x0 )n
n!
n=0
for an x in (a, b) if and only if
f (n+1) (cn )
lim (x − x0 )n+1 = 0.
n→∞ (n + 1)!
1.10. The Taylor's Series 37
Remark: It is not always easy to check this condition, because the sequence {cn }
is usually not precisely known, or even uniquely dened; however, the next theorem
is suciently general to be useful.
rn (n)
lim ∥f ∥I = 0. (1.45)
n→∞ n!
∞
∑ f (n) (x0 )
(x − x0 )n
n!
n=0
converges uniformly to f on
Ir = I ∩ [x0 − r, x0 + r].
f (n+1) (cn )
f (x) − Tn (x) = (x − x0 )n−1 ,
(n + 1)!
rn+1 rn+1
∥f − Tn ∥Ir ≤ ∥f (n+1) ∥Ir ≤ ∥f (n+1) ∥I ,
(n + 1)! (n + 1)!
so (1.45) implies the conclusion.
rn
lim = 0, 0<r<∞
n→∞ n!
∞
∑ x2n+1
sin x = (−1)n , −∞ < x < ∞,
(2n + 1)!
n=0
∞
∑ x2n
cos x = (−1)n , −∞ < x < ∞,
(2n)!
n=0
we conclude that
∞
∑ xn
ex = , −∞ < x < ∞,
n!
n=0
( ) ( )
f (n) (x) q f (n) (0) q
= (1 + x)q−n , so = . (1.46)
n! n n! n
is called the binomial series . We saw in Analysis I that this series equals (1 + x)q
for all x if q is a nonnegative integer.
∞ ( )
∑ q n
x = f (x) = (1 + x)q , 0 ≤ x < 1. (1.47)
n
n=0
Since ( )/( )
q q q − n
lim = lim = 1,
n→∞ n + 1 n n→∞ n + 1
the radius of convergence of the series in (1.47) is 1.
From (1.46),
( )
∥f (n) ∥[0,1]
q q
≤ [max(1, 2 )] , n ≥ 0.
n! n
where the last equality follows from the absolute convergence of the series in (1.47)
on (−1, 1).
Theorem: If
∞
∑
f (x) = an (x − x0 )n , |x − x0 | < R1 , (1.48)
n=0
∞
∑
g(x) = bn (x − x0 )n , |x − x0 | < R2 , (1.49)
n=0
∞
∑
αf (x) + βg(x) = (αan + βbn )(x − x0 )n , |x − x0 | < R,
n=0
where R ≥ min{R1 , R2 }.
∞
∑
f (x) = an (x − x0 )n , |x − x0 | < R1 ,
n=0
∞
∑
g(x) = bn (x − x0 )n , |x − x0 | < R2 ,
n=0
then
∞
∑
f (x)g(x) = cn (x − x0 )n , |x − x0 | < R, (1.50)
n=0
∑n ∑
n
cn = ar bn−r = an−r br
r=0 r=0
and R ≥ min{R1 , R2 }.
∞
∑
f (x) = an (x − x0 )n , |x − x0 | < R1 ,
n=0
1.10. The Taylor's Series 40
∞
∑
g(x) = bn (x − x0 )n , |x − x0 | < R2 ,
n=0
( )
∑
n ∑
n
ar (x − x0 ) bn−r (x − x0 )
r n−r
= ar bn−r (x − x0 )n = cn (x − x0 )n .
r=0 r=0
Example: If
∑ ∞
1
f (x) = = xn , |x| < 1,
1−x
n=0
∞
∑
g(x) = b n xn , |x| < R,
n=0
∑∞
g(x)
= sn xn , |x| < min{1, R},
1−x
n=0
where
∞ ( )
∑ p n
(1 + x)p = x , |x| < 1.
n
n=0
Also
∞ ( )
∑ q n
(1 + x) = q
x , |x| < 1.
n
n=0
Since
∞ (
∑ )
p q p+q p+q n
(1 + x) (1 + x) = (1 + x) = x ,
n
n=0
∑∞ n
while the Cauchy product is n=0 cn x , with
n ( )(
∑ )
p q
cn = .
r n−r
r=0
1.10. The Taylor's Series 41
( ) ∑n ( )( )
p+q p q
= ,
n r n−r
r=0
h(x)
f (x) = (1.51)
g(x)
∞
∑
h(x) = cn (x − x0 )n , |x − x0 | < R1 ,
n=0
∑∞
g(x) = bn (x − x0 )n , |x − x0 | < R2 ,
n=0
∞
∑
f (x) = an (x − x0 )n (1.52)
n=0
b0 = g(x0 ) ̸= 0.
This is surely plausible. Since g(x0 ) ̸= 0 and g is continuous near x0 , the denomina-
tor of (1.51) diers from zero on an interval about x0 . Therefore,f has derivatives
of all orders on this interval, because g and h do.
Since
f (x)g(x) = h(x),
The result about the product of Power series implies that
∑
n
ar bn−r = cn , n ≥ 0.
r=0
c0
a0 = ,
b0
( )
1 ∑
n−1
an = cn − bn−r ar , n ≥ 1.
b0
r=0
1.10. The Taylor's Series 42
Example: Suppose that we wish to nd the coecients in the Maclaurin series
tan x = a0 + a1 x + a2 x2 + · · · .
We rst observe that since tan x is an odd function, its derivatives of even order
vanish at x0 = 0, so a2m = 0, m ≥ 0. Therefore,
tan x = a1 x + a3 x3 + a5 x5 + · · · .
Since
sin x
tan x = ,
cos x
it follows from series of sin x and cos x that
x3 x5
x− + 120 + · · ·
a1 x + a3 x + a5 x + · · · =
3 5 6
2 4
1 − x2 + x24 + · · ·
so
( )
x2 x4 x3 x5
(a1 x + a3 x + a5 x + · · · ) 1 −
3 5
+ + ··· = x − + + ··· ,
2 24 6 120
or
( a1 ) 3 ( a3 a1 ) 5 x3 x5
a1 x + a3 − x + a5 − + x + ··· = x − + + ··· .
2 2 24 6 120
Comparing coecients of like powers of x on the two sides of this equation must be
equal; hence,
a1 = 1, a3 − a1
2 = − 16 , a5 − a3
2 + a1
24 = 1
120 ,
(1)
a1 = 1, a3 = − 16 + 21 (1) = 1
3, a5 = 1
120 + 1
2 3 − 1
24 (1) = 2
15 .
Therefore,
x3 2
tan x = x + + x5 + · · · .
3 15
∞
∑ xn
g(x) = 1 + ex = 2 + ,
n!
n=1
1.10. The Taylor's Series 43
2a0 = 1,
a0 + 2a1 = 0,
a0
+ a1 + 2a2 = 0,
2
a0 a1
+ + a2 + 2a3 = 0.
6 2
Solving these equations successively yields
1
a0 = ,
2
a0 1
a1 = − =− ,
2 4 ( )
1 ( a0 ) 1 1 1
a2 = − + a1 = − − = 0,
2 2 2 4 4
( )
1 ( a0 a1 ) 1 1 1 1
a3 = − + + a2 = − − +0 = ,
2 6 2 2 12 8 48
so
1 1 x x3
= − + + ··· .
1 + ex 2 4 48
∞
∑ xn
g(x) = ex = , (1.53)
n!
n=0
∞
∑
(ex )−1 = an x n ,
n=0
then ( ∞
)( ∞
) ∞
∑ ∑ xn ∑
n
1= an x = cn xn ,
n!
n=0 n=0 n=0
1.11. The Abel's Theorem 44
where
∑
n
ar
cn = .
(n − r)!
r=0
∑
n−1
ar
an = − , n ≥ 1. (1.54)
(n − r)!
r=0
1
a1 = − (1.32) = −1,
1!
[ ]
1 1 1
a2 = − (1) + (−1) = ,
2! 1! 2
[ ( )]
1 1 1 1 1
a3 = − (1) + (−1) + =−
3! 2! 1! 2 6
[ ( ) ( )]
1 1 1 1 1 1 1
a4 = − (1) + (−1) + + − = .
4! 3! 2! 2 1! 6 24
∑n−1 (−1)r
an = − 1
r=0 (n−r)! r!
∑n−1 ( )
= − n!
1
r=0 (−1)
r n
r
(−1)n
= n! .
∞
∑ xn
(ex )−1 = (−1)n .
n!
n=0
Since this is precisely the series that results if x is replaced by −x in (1.53), we have
veried a fundamental property of the exponential function: that
∞
∑
lim f (x) = an R n .
x→(x0 +R)−
n=0
∑∞
• If n=0 (−1)
na
nR
n converges, then
∞
∑
lim f (x) = (−1)n an Rn .
x→(x0 −R)+
n=0
Proof : Let
∞ ∞
∑ ∑
n
g(y) = bn y , bn = s (nite).
n=0 n=0
We will show that
lim g(y) = s. (1.55)
y→1−
We have
∞
∑
g(y) = (1 − y) sn y n , (1.56)
n=0
where
sn = b0 + b1 + · · · + bn .
Since
∑ ∞ ∞
∑
1
= yn and therefore 1 = (1 − y) yn, |y| < 1, (1.57)
1−y
n=0 n=0
we can multiply through by s and write
∞
∑
s = (1 − y) sy n , |y| < 1.
n=0
|sn − s| < ε if n ≥ N + 1.
Then, if 0 < y < 1,
∑
N ∞
∑
|g(y) − s| ≤ (1 − y) |sn − s|y + (1 − y)
n
|sn − s|y n
n=0 n=N +1
∑
N ∞
∑
< (1 − y) |sn − s|y + (1 − y)εy
n N +1
yn
n=0 n=0
∑N
< (1 − y) |sn − s| + ε,
n=0
1.11. The Abel's Theorem 46
∞
∑ xn+1
log(1 + x) = (−1)n , |x| < 1,
n+1
n=0
where the power series converges at x = 1. The Abel's theorem implies that
∞
∑ (−1)n+1
log 2 = .
n+1
n=0
∞ ( )
∑ q n
x
n
n=0
∞ ( )
∑ ∞
∑ ( )
q q
= 2q and (−1)n = 0, q ≥ 0.
n n
n=0 n=0
1.12. Pointwise and Uniform Bounded Functions 47
Suppose there exists a sequence {nk } such that {sin nk x} converges, for every
x ∈ [0, 2π]. Then we must have
Hence
lim (sin nk x − sin nk+1 x)2 = 0, x ∈ [0, 2π].
k→∞
By Lebesgue's theorem concerning integration of bounded convergent sequences,
we have
∫2π
lim lim (sin nk x − sin nk+1 x)2 = 0.
k→∞ k→∞
0
But we have
lim (sin nk x − sin nk+1 x)2 = 2π.
k→∞
which is a contradiction.
x2
Fn (x) = , S = [0, 1].
x2 + (1 − nx)2
1.13. Equicontinuous Functions on a Set 48
Then |Fn | ≤ 1, so that {Fn (x)} is uniformly bounded on [0, 1]. Also
But
1
Fn ( ) = 1,
n
so no subsequence can converge uniformly on [0, 1].
• Due to the boundedness of {Fn (xn )}, we can say that Fn,k (xn ) converges, as
k → ∞.
• The order in which the functions appear is the same in each sequence, i.e., if
one function precedes another in S1 , they are in the same relation in every
Sn , until one or the other is deleted. Hence, when going from one row in the
above array to the next below, functions may move to the left but never to
the right.
Proof : Since the sequence of functions {Fn } is uniformly convergent, for every
ε > 0, there is an integer N such that
We know that continuous functions on compact sets are uniformly continuous, there
is a δ>0 such that
Proof : Since {Fn } is equicontinuous then by denition for every ε > 0, we have
From Analysis I, we know that S is compact then there are nitely many points
p1 , p2 , ...pr in S such that to every x ∈ S corresponds at least one p1 such that
|x − p1 | < δ .
Since {Fn } is pointwise bounded, there exists Mi < ∞ such that
If we take
M = max{M1 , ..., Mr },
then |Fn (x)| < M + ε for every x ∈ S. This proves the rst part of the theorem.
for every i.
If i≥N and j ≥ N, it follows that
|gi (x) − gj (x)| ≤ |gi (x) − gi (xs )| + |gi (xs ) − gj (xs )| + |gj (xs ) − gj (x)|
Proof : Without any loss of generality, we may assume that [a, b] = [0, 1].
We may also assume that f (0) = f (1) = 0. As we can consider
∫1
Qn (x)dx = 1, n = 1, 2, ....
−1
∫1 ∫1
(1 − x ) dx = 2
2 n
(1 − x2 )n dx
−1 0
√
∫ n
1/
≥ 2 (1 − x2 )n dx
0
√
∫ n
1/
≥ 2 (1 − nx2 )dx
0
4
= √
3 n
1
= √
n
1.14. The Stone-Weierstrass Theorem 52
It follows from
∫1
Qn (x)dx = 1, n = 1, 2, ....
−1
√
that cn < n.
For any δ > 0, we have
√
Qn (x) ≤ n(1 − δ 2 )n , δ ≤ |x| ≤ 1.
∫
1−x ∫1
Pn (x) = f (x + t)Qn (t)dt = f (t)Qn (t − x)dt,
−x 0
x.
and the last integral is clearly a polynomial in
Thus {Pn } is a sequence of polynomials.
Given ε > 0, we chose δ > 0 such that |y − x| < δ implies
ε
|f (y) − f (x)| < .
3
Let M = sup |f (x)|, we see that for x ∈ [a, b], we have
∫1
|Pn (x) − f (x)| = | [f (x + t) − f (x)]Qn (t)dt|
−1
∫1
≤ |f (x + t) − f (x)|Qn (t)dt
−1
∫−δ ∫δ
ε
≤ 2M Qn (t)dt + Qn (t)dt
2
−1 −δ
∫1
+2M Qn (t)dt
δ
√ ε
≤ 4M n(1 − δ 2 )n +
2
1.15. Fourier Series 53
Here the functions ϕk (x) are suitable elementary functions, also called the base
set of functions, and the ck are called the coecients of the expansion.
For the Taylor series
∞
∑
f (x) = ck xk , |x| < R,
k=0
f (x + ω1 ) = f (x), f (x + ω2 ) = f (x).
Then so is ω1 ± ω2 .
f (x + (ω1 ± ω2 ))
1.15. Fourier Series 54
Example: sin(cx) and cos(cx) are periodic functions with period 2π/c.
The function
∞
∑
(an cos nx + bn sin nx),
n=1
∫ c+ω ∫ ω
f (x)dx = f (x)dx,
c 0
Consider
∫
c+ω ∫0 ∫ω ∫
c+ω
showing that the integral of a periodic function with period ω taken over an arbitrary
interval of length ω always has the same value.
{
f (x), a ≤ x < a + ω,
f˜( x) =
f (x − nω), a + nω ≤ x < a + (n + 1)ω,
where n is an integer.
n ( )
a0 ∑
sn (x) = + ak cos kx + bk sin kx , x ∈ R,
2
k=1
Recall the Stone and Weierstrass theorem stating that the trigonometric polynomials
are dense in C[a, b] for any closed interval [a, b], provided that b − a < 2π .
n ( )
a0 ∑
sn (x) = + ak cos kx + bk sin kx , x ∈ R,
2
k=1
The sequence {sn }, converges on a set E, then we may dene a function f :E→R
by
∞ ( )
a0 ∑
f (x) = lim sn (x) = + ak cos kx + bk sin kx , x ∈ E.
n→∞ 2
k=1
The series on the right is called a trigonometric series. The constants a0 , ak , bk
(k ∈ N) are called coecients of the trigonometric series.
We have taken the constant term in series as a0 /2 rather than a0 so that we can
make a0 /2 t in a general formula later.
We observe that if the series on the right converges for all real t[0, 2π], then the
sum f must satisfy
f (x) = f (x + 2π), x ∈ R.
2. u + v = v + u.
3. ( u + v ) + w = u + ( v + w)
8. (c + d)u = cu + du.
9. c(du) = (cd)u.
10. 1u = u.
Remark: Using only these axioms, one can show that the zero vector in Axiom 4 is
unique, and the vector −u, called the negative of u, in Axiom 5 is unique for each
u in V .
The Inner Product: Let u, v, and w be vectors in vector space V , and let c be
a scalar. Then an inner product is a function < ., . >: V × V → F such that
Theorem: The space E is a linear space, that is, a vector space. Moreover, E an
inner product space with respect to the inner product
∫π
1
< f, g >= f (x)g(x)dx.
π
−π
1
Φ = { √ , cos(nx), sin(nx) : n ∈ N}
2
1.16. The space E 58
∫π
1
< f, g >= f (x)g(x)dx.
π
−π
∞
∑ ∞
∑
< f, ϕk > ϕk (x) := ck ϕk (x),
k=1 k=1
is called the Fourier series of f (relative to Φ), and the coecients ck =< f, ϕk >
are called the kth Fourier coecient of f.
We introduce ∫ π
1
∥f ∥ =< f, f >=
2
|f (x)|2 dx.
π −π
Suppose that we are given a trigonometric series of the form
∞ ( )
a0 ∑
f (x) = + ak cos kx + bk sin kx , x ∈ E.
2
k=1
Clearly, since each term of the series has period 2π , if it converges to a function
f (x), then f (x) must be a periodic function with period 2π .
Thus, only 2π -periodic functions are expected to have trigonometric series of the
above form.
∞ ( )
a0 ∑
f (x) = + an cos nx + bn sin nx .
2
n=1
∞ ( )
a0 ∑
f (x) = + an cos nx + bn sin nx , (∗)
2
n=1
∞ ( )
|a0 | ∑
+ |an | + |bn |
2
n=1
∫ ∞ ∫ { }
1 π
a0 ∑ 1 π
f (x)dx = + (an cos nx + bn sin nx)
π −π −π 2 π
n=1
{ ∫ } ∑ ∞ { ∫
a0 1 π an π
= dx + cos nxdx
2 π −π π −π
n=1
∫ }
bn π
+ sin nx)
π −π
= a0
Recall:
∫ π ∫ π
1 1
cos nx cos kxdx = δnk = sin nx sin kxdx
π −π π −π
∫ π
cos nx sin kxdx = 0
−π
and
2 cos α cos β = cos(α + β) + cos(α − β)
2 sin α sin β = cos(α − β) − cos(α + β)
2 sin α cos β = sin(α + β) + sin(α − β).
∞ ( )
a0 ∑
f (x) = + an cos nx + bn sin nx , (∗)
2
n=1
Multiply by cos kx and the series forf (x) cos kx can be integrated term by term for
each xed k, we can determine ak and bn .
∫ π ∫
1 a0 1 π
f (x) cos kxdx = cos kxdx +
π −π 2 π −π
∞ (
∑ ∫ π
an cos kx cos nxdx
n=1 −π
∫ π )
+bn sin nx cos kxdx .
−π
∫ π
1
ak = f (x) cos kxdx.
π −π
Multiply by sin kx and the series forf (x) sin kx can be integrated term by term
for each xed k, we can determine bk and bn .
∫ π ∫
1 a0 1 π
f (x) sin kxdx = sin kxdx +
π −π 2 π −π
∞ (
∑ ∫ π
an sin kx cos nxdx
n=1 −π
∫ π )
+bn sin nx sin kxdx .
−π
1.16. The space E 60
∫ π
1
bk = f (x) sin kxdx.
π −π
Fourier Series: For any integrable function f on [−π, π], the numbers ak and bk
dened by
∫ π ∫ π
1 1
ak = f (x) cos kxdx, k ≥ 0, bk = f (x) sin kxdx, k ≥ 1.
π −π π −π
∞ ( )
a0 ∑
+ ak cos kx + bk sin kx ,
2
k=1
∞ ( )
∼ a0 ∑
f (x) = + ak cos kx + bk sin kx ,
2
k=1
to indicate that the Fourier series on the right may or may not converge to f at
some point t ∈ [−π, π].
∞ ( )
a0 ∑
+ ak cos kx + bk sin kx , (∗)
2
k=1
converges uniformly on [−π, π], then it is the Fourier series of its sum.
More precisely, if the trigonometric series (*) converges uniformly to f on [−π, π],
then the ak bk are given by
and
∫ π ∫
1 1 π
ak = f (x) cos kxdx, k ≥ 0, bk = f (x) sin kxdx, k ≥ 1.
π −π π −π
∞ ( )
a0 ∑
+ ak cos kx + bk sin kx , (∗)
2
k=1
However, since
|ak cos kx + bk sin kx| ≤ |ak | + |bk |,
Weierstrass M-test shows that the trigonometric series (*) converges absolutely and
uniformly on every closed interval [a, b] whenever
∞
∑
(|ak | + |bk |)
k=1
is convergent.
1.16. The space E 61
• The sum of two even (or odd) functions is an even (or odd) function.
Then the product function f (x) sin kx is odd, which means that bk = 0 for all k ≥ 1,
and hence we have the Fourier cosine series
∞ ∫
a0 ∑ 1 π
f (x) ∼
= + ak cos kx, ak = f (x) cos kxdx.
2 π −π
k=1
Then the product function f (x) cos kx is odd, which means that ak = 0 for all k ≥ 0,
and hence we have the Fourier cosine series
∞
∑ ∫ π
1
f (x) ∼
= ak sin kx, bk = f (x) sin kxdx.
π −π
k=1
2(1 − (−1)n )
an = − .
n2 π
We have
∞
π 4 ∑ cos(2k + 1)x
|x| = − .
2 π (2k + 1)2
k=0
1.16. The space E 62
∞
π 4 ∑ cos(2k + 1)x
|x| = − .
2 π (2k + 1)2
k=0
Remark: Note that the Fourier series here converges uniformly to |x| on [−π, π]
but not on the whole interval (−∞, ∞), and so outside the interval (−∞, ∞), f (x)
is determined by the periodicity condition f (x) = f (x + 2π).
we can make use of this series to nd the values of some numerical series. For instant
x=0 gives
∞
π2 ∑ 1
= .
8 (2k + 1)2
k=1
• For what values of x does the Fourier series of f converge? Does it converge
for all x in [−π, π]? If it converges on [−π, π] but not to f , what will be its
sum?
• There exists a continuous function whose Fourier series diverges for points in
some set S and converges on (−π, π) \ S .
1.16. The space E 63
The space E: Let us dene the space E be the set of all real valued piecewise
dened periodic function f on the interval [−π, π].
Dene
{
f (x + h) − f (x+)
E′ = f ∈E : lim exists x ∈ [−π, π)
h→0+ h
}
f (x + h) − f (x−)
lim exists x ∈ (−π, π]
h→0− h
Theorem: Let f ∈ E ′. Then for each x ∈ (−π, π), the Fourier series of f (x)
converges to the value
f (x−) + f (x+)
.
2
At the end points x = ±π , the series converges to
f (π−) + f (−π+)
.
2
At the point of discontinuity x, the Fourier series of f assumes the mean of the
one-sided limits of f.
Corollary: f : [−π, π] → R is continuous, and if f (−π) = f (π), f ′ (x)
If exists and
is piecewise continuous on [−π, π], then the Fourier series of f converges to f (x) at
every point x ∈ [−π, π].
Let f (x) be dened for other values of x by the periodicity condition f (x) = f (x +
2π). Then the Fourier series of f on [−π, π] converges to
Example:If f (x) = x on [−π, π) and f (π) = −π. Find the Fourier sine series
of f.
∞
∑ (−1)k−1
x∼
=2 sin kx.
k
k=1
Remarks: Note that the Fourier series does not necessarily agree with f (x) = x
at every point in [−π, π].
The Fourier series vanishes at both endpoints x = ±π , whereas the function
does not vanish at either endpoint.
However, the Dirichlet's theorem states that series converges to f (x) at every
interior point of (−π, π).
For example at x = π/2 ∼
= could be replaced by =
the symbol and so
( )
π 0 (−1) 0 1
=2 1− + − + + ... .
2 2 3 4 5
1.16. The space E 65
∞
∑ (−1)k−1
x∼
=2 sin kx.
k
k=1
Remarks: Finally, we remark that at the endpoints x = ±π, the series converges
to
f (π−) + f ((−π)−) π + (−π)
= = 0.
2 2
we could also consider f as follows: f (x) = x on (−π, π) and f (−π) = f (π) = 0.
∫ ∫ ∫
einx = cos nxdx + i sin nxdx.
According to this, the Fourier coecients are easy to derive quickly by writing
∫ π
1
an − ibn = e−inx ex dx
π −π
1 e (1−in)x
= |π
π 1 − in −π
( )
1 e(1−in)π − e−(1−in)π
=
π 1 − in
(−1) (e − e−π )
n π
=
π(1 − in)
We have
∞
x ∼ sinh π 2 sinh π ∑ (−1)n
e = + cos nx
π π (1 + n2 )
n=1
∞
2 sinh π ∑ (−1)n−1 n
+ sin nx.
π (1 + n2 )
n=1
∞
sinh π 2 sinh π ∑ (−1)n
1= + .
π π (1 + n2 )
n=1
∞
π csc π − 1 ∑ (−1)n
= .
2 (1 + n2 )
n=1
∞
eπ + e−π sinh π 2 sinh π ∑ 1
= + ,
2 π π (1 + n2 )
n=1
∞
∑ 1
π coth π = 1 + 2 .
(1 + n2 )
n=1
Which reduces to
∞
π coth π − 1 ∑ 1
= .
2 (1 + n2 )
n=1
In particular, f ((L/π)t) is 2π -periodic, and so the Fourier series expansion has the
following in terms of the variable t:
∞ ( )
L ∼ a0 ∑
f ( t) = + an cos nt + bn sin nt , t ∈ [−π, π],
π 2
n=1
where ∫ π ∫
1 L 1 L kπ
an = f ( t) cos ntdt = f (x) cos( x)dx.
π −π π L −L L
∞ ( )
L a0 ∑
f ( t) ∼
= + an cos nt + bn sin nt , t ∈ [−π, π],
π 2
n=1
1.17. Fourier Series for Arbitrary Periodic Function 67
where ∫ π ∫ L
1 L 1 kπ
an = f ( t) cos ntdt = f (x) cos( x)dx,
π −π π L −L L
and similarly,
∫ L
1 kπ
bn = f (x) sin( x)dx.
L −L L
We remark that the interval of integration in the last two formulas for the Fourier
coecients can be replaced with an arbitrary interval [c, c + 2L], of length 2L.
Changing the variable t, by setting t = (π/L)x.
Theorem: Let f be a periodic function with period 2L. Then the Fourier expansion
of f is given by
∞ ( )
a0 ∑ nπ nπ
f (x) ∼
= + an cos( x) + bn sin( x) , x ∈ [−L, L],
2 L L
n=1
where ∫ L
1 nπ
an = f (x) cos( x)dx,
L −L L
and ∫ L
1 nπ
bn = f (x) sin( x)dx.
L −L L
Remark: The interval of integration in the last formulas for the Fourier coecients
can be replaced with the interval [c, c + 2L], where c is any real number; we usually
let c = −L. Notice that
nπ nπ
cos( (x + 2L)) = cos( x)
L L
nπ nπ
sin( (x + 2L)) = sin( x).
L L
where ∫ c+2L
1 nπ
an = f (x) cos( x)dx.
L c L
and the Fourier series of an odd function f with period 2L is a Fourier sine series
∞
∑ nπ
f (x) ∼
= bn sin( x), x ∈ [c, c + 2L],
L
n=1
1.17. Fourier Series for Arbitrary Periodic Function 68
where ∫ c+2L
1 nπ
bn = f (x) sin( x)dx
L c L
where c is any real number.
Example: Consider the function f (x) = | sin x|. The function is dened for all x
and the function has period π.
Clearly, f represents a continuous, piecewise smooth, even function of period π,
and therefore it is everywhere equal to its Fourier series, consisting of cosine terms
only.
We have c = 0, and L = π/2, then we have
∫
2 π
ak = f (x) cos(2kx)dx
π 0
∫
2 π
= sin x cos(2kx)dx
π 0
∫ π[ ]
1
= sin(1 + 2k)x − sin(2k − 1)x dx
π 0
∫ [ ]
1 π
= sin(1 + 2k)x − sin(2k − 1)x dx
π 0
( )
1 cos(1 + 2k)x cos(2k − 1)x π
= − +
π 2k + 1 2k − 1 0
( )
1 (−1)2k+1 − 1 (−1)2k−1 − 1
= − −
π 2k + 1 2k − 1
4
ak = − .
π(4k 2 − 1)
1.18. Best Approximation Theorem 69
∞
2 4 ∑ cos 2kx
| sin x| = − , x ∈ [−π, π].
π π 4k 2 − 1
k=1
∑
n ∫π
1
|ck | ≤
2
|f (x)|2 dx.
π
k=1 −π
∑n
Proof : Setting Sn = k=1 ck ϕk (x). Then we have
∫π
1
∥f − Tn ∥2 = |f (x) − Tn (x)|2 dx
π
−π
∫π ∫π
1 1
= |f (x)| dx +
2
|Tn (x)|2 dx
π π
−π −π
∫π
1
−2 dk f (x)ϕk (x)dx
π
−π
∫π ∑
n ∑
n
1
= |f (x)|2 dx + |dk |2 − 2 ck dk
π
−π k=1 k=1
∫π ∑
n ∑
n
1
= |f (x)|2 dx + |dk |2 − 2 ck dk
π
−π k=1 k=1
∫π ∑
n ∑
n
1
= |f (x)|2 dx + |ck − dk |2 − 2 ck
π
−π k=1 k=1
∑
n
= ∥f − Sn ∥2 + |ck − dk |2 .
k=1
1.18. Best Approximation Theorem 70
Therefore,
∥f − Tn ∥2 ≥ ∥f − Sn ∥2 ,
with equality if and only if ck = dk for each k = 1, ..., n.
∥f − Tn ∥2 ≥ ∥f − Sn ∥2 ,
Note that f and ϕk are xed, while the dk are allowed to vary.
In particular, setting dk = ck , shows that the minimum value of ∥f − Tn ∥2 ≥
∥f − Sn ∥2 , is given by
∫π ∑
n ∑
n
1
min ∥f − Tn ∥ = 2
|f (x)| dx −
2
|ck | = ∥f ∥ −
2 2
|ck |2 ,
Tn π
−π k=1 k=1
∑
n ∫π
1
|ck | ≤
2
|f (x)|2 dx for all n.
π
k=1 −π
Chapter 2
is
X + Y = (x1 + y1 , x2 + y2 , . . . , xn + yn ). (2.1)
In R4 , let
( )
X = (1, −2, 6, 5) and Y = 3, −5, 4, 12 .
Then
( )
X + Y = 4, −7, 10, 11
2
and
6X = (6, −12, 36, 30).
• There is a unique vector 0, called the zero vector, such that X + 0 = X for all
in R .
X n
• a(bX) = (ab)X.
• (a + b)X = aX + bX.
• a(X + Y) = aX + aY.
• 1X = X.
2.2. Schwarz's Inequality 72
( )
X = (1, −2, 6, 5) and Y = 3, −5, 4, 12
are
√
|X| = (12 + (−2)2 + 62 + 52 )1/2 = 66
and √
201
|Y| = (3 + (−5) + 4 +
2 2 2
( 12 )2 )1/2 = .
2
The distance between X and Y is
√
149
|X − Y| = ((1 − 3) + (−2 + 5) + (6 − 4) + (5 −
2 2 2 1 2 1/2
2) ) = .
2
The inner product X · Y of X = (x1 , x2 , . . . , xn ) and Y = (y1 , y2 , . . . , yn ) is
X · Y = x1 y1 + x2 y2 + · · · + xn yn .
with equality if and only if one of the vectors is a scalar multiple of the other.
∑
n
0 ≤ (xi − tyi )2
i=1
∑n ∑n ∑n (2.4)
= x2i − 2t i=1 xi yi + t2 2
i=1 yi
i=1
= |X|2 − 2(X · Y)t + t2 |Y|2 .
2.2. Schwarz's Inequality 73
because if not, then p would have two distinct real zeros and therefore be negative
between them, contradicting the inequality (2.4).
Proof :
(X · Y)2 ≤ |X|2 |Y|2 , (2.6)
Conversely, if equality holds in (2.3), then p has the real zero t0 = (X · Y)/|Y|2 ,
and
∑
n
(xi − t0 yi )2 = 0
i=1
with equality if and only if one of the vectors is a nonnegative multiple of the other.
Proof : By denition,
∑n
|X + Y|2 = (xi + yi )2
∑i=1
n 2
∑n ∑n 2
= i=1 xi + 2 i=1 xi yi + i=1 yi
= (|X| + |Y|)2 .
Hence,
|X + Y|2 ≤ (|X| + |Y|)2 .
Taking square roots yields (2.7).
From the third line of (2.8), equality holds in (2.7) if and only if X · Y = |X||Y|,
which is true if and only if one of the vectors X and Y is a nonnegative scalar
multiple of the other.
2.2. Schwarz's Inequality 74
|X − Z| ≤ |X − Y| + |Y − Z|.
Proof : Write
X − Z = (X − Y) + (Y − Z),
and apply triangle inequality with X and Y replaced by X−Y and Y − Z.
|X − Y| ≥ ||X| − |Y|| .
Proof : Since
X = Y + (X − Y),
Triangle inequality implies that
Since |X − Y| = |Y − X|, the last two inequalities imply the stated conclusion.
• X · Y = Y · X.
• X · (Y + Z) = X · Y + X · Z.
x = x0 + u1 t, y = y0 + u2 t, z = z0 + u3 t, −∞ < t < ∞,
where u1 , u2 , and u3 are not all zero. We write this in vector form as
with U = (u1 , u2 , u3 ), and we say that the line is through X0 in the direction of U.
There are many ways to represent a given line parametrically.
For example,
X = X0 + sV, −∞ < s < ∞, (2.10)
represents the same line as (2.9) if and only if V = aU for some nonzero real number
a.
Then the line is traversed in the same direction as s and t vary from −∞ to ∞ if
a > 0, or in opposite directions if a < 0. To write the parametric equation of a line
through two points X0 and X1 in R3 .
Then the line through X0 in the direction of U is the set of all points in Rn of the
form
X = X0 + tU, −∞ < t < ∞.
A set of points of the form
X = X0 + tU, t 1 ≤ t ≤ t2 ,
is called a line segment . The line segment from X0 to X1 is the set of points of the
form
X = X0 + t(X1 − X0 ) = tX1 + (1 − t)X0 , 0 ≤ t ≤ 1.
Nε (X0 ) in R2
2.3. Neighbourhoods and Open Sets in Rn 76
We are going to dene neighborhood, interior point, interior of a set, open set,
closed set,limit point, boundary point, boundary of a set, closure of a set, isolated
point, exterior point, and exterior of a set.
(−1, 1) (1, 1)
x x
Example: X0 is
If a point in Rn and r is a positive number, the open n-ball of
radius r about X0 is the set
|X − X1 | < ε = r − |X − X0 |,
We can show that the closure of Br (X0 ) is the closed n-ball of radius r about X0 ,
dened by
S r (X0 ) = {X : |X − X0 | ≤ r}.
Remark: Open and closed n-balls are generalizations to Rn of open and closed
intervals.
Lemma: If X1 andX2 are in Sr (X0 ) for some r > 0, then so is every point on the
line segment from X1 to X2 .
Theorem: Let
lim xir = xi , 1 ≤ i ≤ n;
r→∞
that is, a sequence {Xr } of points in Rn converges to a limit X if and only if the
sequences of components of {Xr } converge to the respective components of X.
2.5. Principle of nested sets 78
|Xr − Xs | < ε if r, s ≥ K.
d(S) = sup{|X − Y| : X, Y ∈ S}
is the diameter of S.
S1 ⊃ S2 ⊃ · · · ⊃ Sr ⊃ · · · (2.11)
and
lim d(Sr ) = 0, (2.12)
r→∞
|X − Y| ≤ d(Sk ), k ≥ 1,
S ⊂ ∪{H : H ∈ H}.
T = {(x, y)|a1 ≤ x ≤ a1 + L, a2 ≤ x ≤ a2 + L}
T (1) T (2)
S (1) S (2)
S (4) S (3)
T (4) T (3)
Bisecting the sides of T leads to four closed squares, T (1) , T (2) , T (3) , and T (4) ,
with sides of length L/2. Let
S (i) = S ∩ T (i) , 1 ≤ i ≤ 4.
∪
4
S= S (i) .
i=1
2.6. Heine-Borel Theorem 80
Moreover, H covers each S (i) , but at least one S (i) cannot be covered by any nite
subcollection of H, since if all the S
(i) could be, then so could S. Let S1 be a set
with this property, chosen from S
(1) , S (2) , S (3) , and S (4) .
We are now back to the situation we started from: a compact set S1 covered by H,
but not by any nite subcollection of H. However, S1 is contained in a square T1
with sides of length L/2 L. Bisecting the sides of T1 and repeating the
instead of
argument, we obtain a subset S2 of S1 that has the same properties as S , except that
it is contained in a square with sides of length L/4. Continuing in this way produces
a sequence of nonempty closed sets S0 (= S), S1 , S2 , . . . , such that Sk ⊃ Sk+1 and
d(Sk ) ≤ L/2k−1/2 (k ≥ 0).
∩∞
From Principle of Nested Sets Theorem, there is a point X in k=1 Sk .
Since X ∈ S , there is an open set H in H that contains X, and this H must also
contain some ε-neighborhood of X. Since every X in Sk satises the inequality
|X − X| ≤ 2−k+1/2 L,
T = {(x1 , x2 , . . . , xn ) : ai ≤ xi ≤ ai + L, i = 1, 2, . . . , n}.
A ̸= ∅, B ̸= ∅, A ∩ B = ∅, and A ∩ B = ∅. (2.13)
Example: Rn is connected.
The space
If R
= A ∪ B with A ∩ B = ∅ and A ∩ B = ∅, then A ⊂ A and B ⊂ B .
n
That is, A and B are both closed and therefore are both open.
Since the only nonempty subset of R that is both open and closed is R
n n itself,
one of A and B is R and the other is empty.
n
Proof : For suciency, we will show that if S is disconnected, then S is not poly-
gonally connected.
A ̸= ∅, B ̸= ∅, A ∩ B = ∅, and A ∩ B = ∅.
2.8. Polygonally Connected Set 82
.
Suppose that X1 ∈ A and X2 ∈ B , and assume that there is a polygonal path
in S connecting X1 to X2 . Then some line segment L in this path must contain a
point Y1 in A and a point Y2 in B .
The line segment
X = tY2 + (1 − t)Y1 , 0 ≤ t ≤ 1,
X = tY1 + (1 − t)Y0 , 0 ≤ t ≤ 1,
which lies in Nε (Y0 ) and therefore in S. This implies that Y0 can be connected
to X0 by a polygonal path in S Nε (Y0 ) can also.
if and only if every member of
Thus, Nε (Y0 ) ⊂ A if Y0 ∈ A, and Nε (Y0 ) ∈ B if Y0 ∈ B . Therefore, A and B are
open. Since A ∩ B = ∅, this implies that A ∩ B = A ∩ B = ∅. Since A is nonempty
(X0 ∈ A), it now follows that B = ∅, since if B ̸= ∅, S would be disconnected.
Therefore, A = S , which completes the proof of necessity.
Remark: Any polygonally connected set, open or not, is connected. The converse
is false. A set (not open) may be connected but not polygonally connected.
Example: Intervals are the only regions in R. The n-ball Br (X0 ) is a region in Rn ,
S r (X0 ). The set
as is its closure S = {(x, y) : x2 + y 2 ≤ 1 or x2 + y 2 ≥ 4} is not
a region in R , since it is not connected.
2
to S is connected but is not a region, since points on the line segment are not limit
points of S10 . The set S2 obtained by adding to S1 the points in the rst quadrant
bounded by the circles x2 + y 2 = 1 and x2 + y 2 = 4 and the line segments L1 and
is a region.
2.9 Sequences in Rn
A sequence {Xr } of points in Rn converges to a limit X if and only if for every ε > 0
there is an integer K such that
|Xr − X| < ε if r ≥ K.
Since Rn is not ordered for n > 1, monotonicity, limits inferior and superior of
sequences in R , and divergence to ±∞ are undened for n > 1.
n
( )1/2
f (X) = 1 − x21 − x22 − · · · − x2n (2.14)
( )
2 −1
g(X) = 1 − x21 − x22 − · · · − xn (2.15)
lim f (X) = L,
X→X0
if X0 is a limit point of Df and, for every ε > 0, there is a δ>0 such that
|f (X) − L| < ε
+2|(y + y0 )(y − y0 )|
≤ |X − X0 |(|x + x0 | + 2|y + y0 )|),
2.11. Limit at a Point of a Function of n Variables 86
since
|x − x0 | ≤ |X − X0 | and |y − y0 | ≤ |X − X0 |.
If |X − X0 | < 1, then |x| < |x0 | + 1 and |y| < |y0 | + 1.
This and (2.17) imply that
where
K = (2|x0 | + 1) + 2(2|y0 | + 1).
Therefore, if ε>0 and
then
g(x, y) − (1 − x20 − 2y02 ) < ε.
x2 + 2y 2 = 1.
It is not dened at any point of the ellipse itself or on any deleted neighborhood of
y y
X0
X − X0 = δ
x x
x 2 + 2y 2 = 1 x 2 + 2y 2 = 1
(a) (b)
if
x20 + 2y02 = 1. (2.19)
Then
sin u(x, y)
h(x, y) = . (2.20)
u(x, y)
Recall that
sin r
lim = 1.
r→0r
Therefore, if ε > 0, there is a δ1 > 0 such that
sin u
u − 1 < ε if 0 < |u| < δ1 . (2.21)
lim (1 − x2 − 2y 2 ) = 0.
(x,y)→(x0 ,y0 )
if X = (x, y) is in the interior of the ellipse and |X − X0 | < δ ; that is, if X is in the
shaded region.
Therefore,
√
0<u= 1 − x2 − 2y 2 < δ1 (2.22)
if X is in the interior of the ellipse and |X − X0 | < δ ; that is, if X is in the shaded
region. This, (2.20), and (2.21) imply that
|h(x, y) − 1| < ε
If we try to answer this question by letting (x, y) approach (0, 0) along the line
y = x, we see the functional values
x2 1
f (x, x) = =
2x2 2
and conclude that the limit is 1/2.
2.12. Innite Limits and Limits at X → ∞ 88
However, if we let (x, y) approach (0, 0) along the line y = −x, we see the
functional values
x2 1
f (x, −x) = − 2
=−
2x 2
and conclude that the limit equals −1/2.
In fact, they are both incorrect. What we have shown is that
1 1
lim f (x, x) = and lim f (x, −x) = − .
x→0 2 x→0 2
Since limx→0 f (x, x) and limx→0 f (x, −x) must both equal lim(x,y)→(0,0) f (x, y).
Then
We say that
lim f (X) = −∞
X→X0
if
lim (−f )(X) = ∞.
X→X0
Example: If
f (X) = (1 − x21 − x22 − · · · − x2n )−1/2 ,
2.12. Innite Limits and Limits at X → ∞ 89
then
lim f (X) = ∞
X→X0
if |X0 | = 1, because
1
f (X) = ,
|X − X0 |
so
1
f (X) > M if 0 < |X − X0 | < δ = .
M
Example: If
1
f (x, y) = ,
x + 2y + 1
then lim(x,y)→(1,−1) f (x, y) does not exist (why not?).
But
lim |f (x, y)| = ∞.
(x,y)→(1,−1)
1 1
|f (x, y)| = ≥√ .
|x + 2y + 1| 5|X − X0 |
Therefore,
1
|f (x, y)| > M if 0 < |X − X0 | < √ .
M 5
1
xk = yk = zk = √ ( ) ,
3 k + 12 π
then ( )
1
f (Xk ) = k + π.
2
2.12. Innite Limits and Limits at X → ∞ 90
However, this does not imply that limX→0 f (X) = ∞. Since, for example, every
neighborhood of (0, 0, 0) also contains points
( )
1 1 1
Xk = √ ,√ ,√ .
3kπ 3kπ 3kπ
Example: If ( )
1
f (x, y, z) = cos ,
x + 2y 2 + z 2
2
then
lim f (X) = 1. (2.27)
|X|→∞
To see this, we recall that the continuity of cos u at u = 0 implies that for each
ε>0 there is a δ>0 such that
Since
1 1
≤ .
x2 + 2y 2 + z 2 |X|2
√
It follows that if |X| > 1/ δ , then
1
< δ.
x2 + 2y 2 + z 2
Therefore,
|f (X) − 1| < ε.
by
1
f (x, y) = .
x−y
We will show that
lim f (x, y) = 0. (2.28)
|X|→∞
and
|X|2 = x2 + y 2 ≤ x2 (1 + a2 ).
So
|X|
x≥ √ .
1 + a2
This and (2.29) imply that
1−a
x−y ≥ √ |X|, X ∈ D.
1 + a2
So √
1 + a2 1
|f (x, y)| ≤ , X ∈ D.
1 − a |X|
This and (2.29) imply that
1−a
x−y ≥ √ |X|, X ∈ D.
1 + a2
So √
1 + a2 1
|f (x, y)| ≤ , X ∈ D.
1 − a |X|
Therefore,
|f (x, y)| < ε
if X∈D and √
1 + a2 1
|X| > .
1−a ε
We will say that limX→X0 f (X) exists in the extended reals. A similar convention
applies to limits as |X| → ∞.
2.13. Continuity 92
2.13 Continuity
If X0 is in Df and is a limit point of Df , then we say that f is continuous at X0 if
whenever
|X − X0 | < δ and X ∈ Df .
Example: The function
f (x, y) = 1 − x2 − 2y 2
is continuous on R2 .
Solution: See lecture.
Example: Consider the function
√
√ 1−x −2y , x2 + 2y 2 < 1,
sin 2 2
then it follows from the example we have discussed that h is continuous on the
ellipse
x2 + 2y 2 = 1.
xy
f (x, y) = ,
x2 + y 2
to make it continuous at (0, 0).
The limit
lim f (x, y)
(x,y)→(0,0)
if
lim gi (U) = Li , 1 ≤ i ≤ n,
U→U0
lim G(U) = L
U→U0
Similarly, G is continuous at U0 if and only if for each ε>0 there is a δ>0 such
that
h=f ◦G
on T by
h(U) = f (G(U)), U ∈ T.
T = {U : U ∈ DG and G(U) ∈ Df },
2.14. Vector Valued Functions 94
R(G) = range of G
n
m
G
DG
Df
if
|X − G(U0 )| < ε1 and X ∈ Df . (2.31)
if
|U − U0 | < δ and U ∈ T.
Example: If
√
f (s) = s
and
g(x, y) = 1 − x2 − 2y 2 ,
then Df = [0, ∞], Dg = R2 , and
T = {(x, y) : x2 + 2y 2 ≤ 1}.
√
h(x, y) = f (g(x, y)) = 1 − x2 − 2y 2
is continuous on T.
Example: If
√
g(x, y) = 1 − x2 − 2y 2
and { sin s
s , s ̸= 0,
f (s) =
1, s = 0,
then Df = (−∞, ∞) and
Dg = T = {(x, y) : x2 + 2y 2 ≤ 1}.
is continuous on
Dg = T = {(x, y) : x2 + 2y 2 ≤ 1}.
Then
f (X1 ) = α and f (X2 ) = β
for some X1 and X2 in S.
and X ∈ S.
This means that X0 ̸∈ T . Therefore, R ∩T = ∅. Similarly, R ∩T = ∅. Therefore,
S is disconnected, which contradicts the assumption that S is a region. Hence, we
conclude that f (C) = u for some C in S.
whenever
|X − X′ | < δ
and X, X′ ∈ S .
Remark: We emphasize that δ must depend only on ε and S, and not on the
particular points X ′
and X .
h(t) = f (X + tΦ)
at t = 0, if h′ (0) exists.
Then
h(t) = 3(x + tϕ1 )(y + tϕ2 )(z + tϕ3 ) + 2(x + tϕ1 )2 + (z + tϕ3 )2
Then we have
Therefore,
∂f (X)
= h′ (0) = (3yz + 4x)ϕ1 + 3xzϕ2 + (3xy + 2z)ϕ3 .
∂Φ
Since X and X + tEi dier only in the ith coordinate, ∂f (X)/∂Ei is called the
partial derivative of f with respect to xi at X.
It is also denoted by ∂f (X)/∂xi or fxi (X); thus,
if the limits exist. If we write X = (x, y), then we denote the partial derivatives
accordingly; thus,
∂f (x, y) f (x + h, y) − f (x, y)
= fx (x, y) = lim
∂x h→0 h
∂f (x, y) f (x, y + h) − f (x, y)
= fy (x, y) = lim .
∂y h→0 h
It can be seen from these denitions that to compute fxi (X) we simply dierentiate
f with respect to xi according to the rules for ordinary dierentiation, while treating
the other variables as constants.
Example: Let
f (x, y, z) = 3xyz + 2x2 + z 2 .
Taking Φ = E1 (that is, setting ϕ1 = 1 and ϕ2 = ϕ3 = 0), we nd that
∂f (X) ∂f (X)
= = 3yz + 4x,
∂x ∂E1
which is the result obtained by regarding y and z as constants in and taking the
ordinary derivative with respect to x. Similarly,
∂f (X) ∂f (X)
= = 3xz
∂y ∂E2
∂f (X) ∂f (X)
= = 3xy + 2z.
∂z ∂E3
∂(f + g)(X)
= fxi (X) + gxi (X),
∂xi
∂(f g)(X)
= fxi (X)g(X) + f (X)gxi (X),
∂xi
and, if g(X) ̸= 0,
∂(f /g)(X) g(X)fxi (X) − f (X)gxi (X)
= .
∂xi [g(X)]2
If fxi (X) exists at every point of a set D, then it denes a function fxi on D.
If this function has a partial derivative with respect to xj on a subset of D, we
denote the partial derivative by
( )
∂ ∂f ∂2f
= = fxi xj .
∂xj ∂xi ∂xj ∂xi
Similarly,
( )
∂ ∂2f ∂3f
= = fxi xj xk .
∂xk ∂xj ∂xi ∂xk ∂xj ∂xi
2.16. Directional Derivative 99
∂rf
= fxi1 · · · xir−1 xir ;
∂xir ∂xir−1 · · · ∂xi1
f (x, y) = 3x2 y 3 + xy
Compute fxx (0, 0), fyy (0, 0), fxy (0, 0), and fyx (0, 0) if
{ 2
(x y+xy 2 ) sin(x−y)
x2 +y 2
, (x, y) ̸= (0, 0),
f (x, y) =
0, (x, y) = (0, 0).
If (x, y) ̸= (0, 0), the ordinary rules for dierentiation, applied separately to x
and y , yield
(2xy+y 2 ) sin(x−y)+(x2 y+xy 2 ) cos(x−y)
fx (x, y) = x2 +y 2
(2.32)
2x(x2 y+xy 2 ) sin(x−y)
− (x2 +y 2 )2
, (x, y) ̸= (0, 0),
and
(x2 +2xy) sin(x−y)−(x2 y+xy 2 ) cos(x−y)
fy (x, y) = x2 +y 2
(2.33)
2y(x2 y+xy 2 ) sin(x−y)
− (x2 +y 2 )2
, (x, y) ̸= (0, 0).
These formulas do not apply if (x, y) = (0, 0), so we nd fx (0, 0) and fy (0, 0) from
their denitions as dierence quotients:
so
fx (x, 0) − fx (0, 0) 0−0
fxx (0, 0) = lim = lim = 0.
x→0 x x→0 x
fy (x, 0) − fy (0, 0) sin x − 0
fyx (0, 0) = lim = lim = 1.
x→0 x x→0 x
Setting x=0 in (2.32) and (2.33) yields
so
Proof : Suppose that ε > 0. Choose δ>0 so that the open square
is in N.
|fxy (b
x, yb) − fxy (x0 , y0 )| < ε if x, yb) ∈ Sδ .
(b (2.35)
where
ϕ(x) = f (x, y0 + k) − f (x, y0 ).
Since
ϕ′ (x) = fx (x, y0 + k) − fx (x, y0 ), |x − x0 | < δ,
2.16. Directional Derivative 101
x, y0 + k) − fx (b
A(h, k) = [fx (b x, y0 )] h. (2.38)
where b
x is between x0 and x0 + h.
The mean value theorem, applied to fx (b
x, y) (where b
x is regarded as constant),
also implies that
x, y0 + k) − fx (b
fx (b x, yb)k,
x, y0 ) = fxy (b
where yb is between y0 and y0 + k .
From this and (2.38),
x, yb)hk.
A(h, k) = fxy (b
Now (2.35) implies that
A(h, k)
− f (x , y ) = |fxy (b
x, yb) − fxy (x0 , y0 )| < ε
hk xy 0 0
if 0 < |h|, |k| < δ.
fy (x0 + h, y0 ) − fy (x0 , y0 )
− f (x , y )≤ε if 0 < |h| < δ.
h
xy 0 0
If this number is rk , we denote the common value of the two sides of (2.39) by
∂ r f (X)
. (2.40)
∂xr11 ∂xr22 · · · ∂xrnn
It being understood that
0 ≤ rk ≤ r, 1 ≤ k ≤ n, (2.41)
r1 + r2 + · · · + rn = r, (2.42)
Then
= x2 − x20 + 2(xy − x0 y0 )
= (x − x0 )(x + x0 ) + 2(xy − x0 y)
+2(x0 y − x0 y0 )
= (x + x0 + 2y)(x − x0 ) + 2x0 (y − y0 )
+ (x − x0 )(x − x0 + 2y − 2y0 )
= m1 (x − x0 ) + m2 (y − y0 ) + (x − x0 )(x − x0 + 2y − 2y0 ),
where
Therefore,
|f (x, y) − f (x0 , y0 ) − m1 (x − x0 ) − m2 (y − y0 )|
|X − X0 |
|x − x0 ||(x − x0 ) + 2(y − y0 )|
=
|X − X0 |
√
≤ 5|X − X0 |,
f (x, y) − f (x0 , y0 ) − m1 (x − x0 ) − m2 (y − y0 )
lim = 0,
X→X0 |X − X0 |
so f is dierentiable at (x0 , y0 ).
∑
f (X) − f (X0 ) − ni=1 mi (xi − xi0 )
lim = 0,
X→X0 |X − X0 |
are given by
mi = fxi (X0 ), 1 ≤ i ≤ n; (2.46)
2.17. Dierentiability of Functions of Several Variables 104
that is, ∑
f (X) − f (X0 ) − ni=1 fxi (X0 )(xi − xi0 )
lim = 0.
X→X0 |X − X0 |
L(X) = m1 x1 + m2 x2 + · · · + mn xn , (2.47)
where
lim E(X) = 0. (2.49)
X→X0
|L(X − X0 )| ≤ M |X − X0 |,
where
M = (m21 + m22 + · · · + m2n )1/2 .
This and f (X) − f (X0 ) = L(X − X0 ) + E(X)(|X − X0 |), imply that
2.17.2 Dierential
The linear function
L(X) = fx1 (X0 )x1 + fx2 (X0 )x2 + · · · + fxn (X0 )xn .
(dX0 f )(X) = fx1 (X0 )x1 + fx2 (X0 )x2 + · · · + fxn (X0 )xn . (2.50)
dxi (X) = xi ;
that is, dxi is the function whose value at a point in Rn is the ith coordinate of the
point.
It is the dierential of the function gi (X) = xi . From (2.50),
dX0 f = fx1 (X0 ) dx1 + fx2 (X0 dx2 + · · · + fxn (X0 ) dxn . (2.51)
dx(X) = x, dy(X) = y, . . .
When it is not necessary to emphasize the specic point X0 , (2.51) can be written
more simply as
df = fx1 dx1 + fx2 dx2 + · · · + fxn dxn .
When dealing with a specic function at an arbitrary point of its domain, we may
use the hybrid notation
d X0 f = 6 dx + 2 dy
(dX0 f )(X − X0 ) = 6(x − 1) + 2(y − 2).
dx(t) = t.
For example, if
f (x) = 3x2 + 5x3 ,
then
df = (6x + 15x2 ) dx.
If x0 = −1, then
Proof : Let X0 = (x10 , x20 , . . . , xn0 ) and suppose that ε > 0. Our assumptions
imply that there is a δ > 0 such that fx1 , fx2 , . . . , fxn are dened in the n-ball
Sδ (X0 ) = {X : |X − X0 | < δ}
and
|fxj (X) − fxj (X0 )| < ε if |X − X0 | < δ, 1 ≤ j ≤ n. (2.52)
∑
n
f (X) − f (X0 ) = f (Xn ) − f (X0 ) = [f (Xj ) − f (Xj−1 )], (2.53)
j=1
where, in each case, all variables except t are temporarily regarded as constants.
Since
f (Xj ) − f (Xj−1 ) = gj (xj ) − gj (xj0 ),
2.17. Dierentiability of Functions of Several Variables 108
b j ),
gj′ (τj ) = fxj (X
where bj
X is on the line segment from Xj−1 to Xj . Therefore,
b j )(xj − xj0 ),
f (Xj ) − f (Xj−1 ) = fxj (X
∑
n
f (X) − f (X0 ) = b j )(xj − xj0 )
fxj (X
j=1
∑n ∑
n
= fxj (X0 )(xj − xj0 ) + b j ) − fx (X0 )](xj − xj0 ).
[fxj (X j
j=1 j=1
∑ ∑
n
n
f (X) − f (X0 ) − f (X )(x − x ) ≤ ε |xj − xj0 | ≤ nε|X − X0 |,
xj 0 j j0
j=1 j=1
Example: If
x2 + y 2
f (x, y) = ,
x−y
then
2x x2 + y 2
fx (x, y) = −
x − y (x − y)2
2y x2 + y 2
fy (x, y) = + .
x − y (x − y)2
Since fx and fy are continuous on
S = {(x, y) : x ̸= y},
2.17. Dierentiability of Functions of Several Variables 109
f is continuously dierentiable on S.
Example: let {
(x − y)2 sin x−y
1
, x ̸= y,
f (x, y) =
0, x = y.
Then
1 1
fx (x, y) = 2(x − y) sin − cos , x ̸= y,
x−y x−y
and
f (x + h, x) − f (x, x) h2 sin(1/h) − 0
fx (x, x) = lim = lim = 0,
h→0 h h→0 h
so fx exists for all (x, y), but is not continuous on the line y = x.
Example: Let {
(x − y)2 sin x−y
1
, x ̸= y,
f (x, y) =
0, x = y.
The same is true of fy , since
1 1
fy (x, y) = −2(x − y) sin + cos , x ̸= y,
x−y x−y
and
f (x, x + k) − f (x, x) k 2 sin(−1/k) − 0
fy (x, x) = lim = lim = 0.
k→0 k k→0 k
Now,
f (x, y) − f (0, 0) − fx (0, 0)x − fy (0, 0)y
√
x2 + y 2
{ (x−y)2
√ 1
sin x−y , x ̸= y,
2 +y 2
= x
0, x = y,
and Schwarz's inequality implies that
(x − y)2 2(x2 + y 2 ) √
1
√ sin ≤ √ = 2 x2 + y 2 , x ̸= y.
x2 + y 2 x − y x2 + y 2
Therefore,
so f is dierentiable at (0, 0), but fx and fy are not continuous at (0, 0).
2.17. Dierentiability of Functions of Several Variables 110
f (x) − T (x)
lim = 0.
x→x0 x − x0
Moreover, the tangent line is the limit of the secant line through the points
(x1 , f (x0 )) and (x0 , f (x0 )) as x1 approaches x0 . Dierentiability of a function of n
variables has an analogous geometric interpretation. We will illustrate it for n = 2.
If f is dened in a region D in R2 , then the set of points (x, y, z) such that
z = f (x, y)
y
D
intersects the surface z = f (x, y) at (x0 , y0 , f (x0 , y0 )) and approximates the surface
so well near (x0 , y0 ) that
f (x, y) − T (x, y)
lim √ = 0.
(x,y)→(x0 ,y0 ) (x − x0 )2 + (y − y0 )2
Ta
n ge
nt
pla
z = f (x,y) ne
y
(x 0 , y0 )
Show that the tangent plane to the surface z = f (x, y) is the limit of the
secant planes.
Let Xi = (xi , yi ) (i = 1, 2, 3). The equation of the secant plane through the
points (xi , yi , f (xi , yi )) (i = 1, 2, 3) on the surface z = f (x, y) is of the form
and
u1 v2 − u2 v1 ̸= 0.
If f is dierentiable at X0 , then
where
lim ε(X) = 0. (2.64)
X→X0
because of (2.64). Substituting (2.65) and (2.66) into (2.61) and (2.62) yields
where
v2 |U|E1 (t) − u2 |V|E2 (t)
∆1 (t) =
u 1 v2 − u 2 v1
and
u1 |V|E2 (t) − v1 |U|E1 (t)
∆2 (t) = ,
u1 v2 − u2 v1
so
lim ∆i (t) = 0, i = 1, 2, (2.69)
t→0
because of (2.67).
From (2.57) and (2.68), the equation of the secant plane is
Therefore, because of (2.69), the secant plane approaches the tangent plane (2.56)
as t approaches zero.
2.18. Maxima and Minima 113
f (X) − f (X0 )
f (X) ≤ f (X0 )
f (x, y) = x3 + y 3 .
and dene
1/2
∑ m (
n ∑ )
∂gi (U0 2
M = .
∂uj
i=1 j=1
|G(U) − G(U0 )|
<M +ε if 0 < |U − U0 | < δ.
|U − U0 |
where
lim Ei (U) = 0, 1 ≤ i ≤ n. (2.72)
U→U0
where
∑m ( )2 1/2
∂g (U )
Mi = .
i 0
∂uj
j=1
Therefore,
( n )1/2
|G(U) − G(U0 )| ∑
≤ (Mi + |Ei (U)|) 2
.
|U − U0 |
i=1
From (2.72),
( n )1/2 ( )1/2
∑ ∑
n
lim (Mi + |Ei (U)|) 2
= Mi2 = M,
U→U0
i=1 i=1
is dierentiable at U0 , and
dU0 h = fx1 (X0 )dU0 g1 + fx2 (X0 )dU0 g2 + · · · + fxn (X0 )dU0 gn . (2.74)
xi0 = gi (U0 ), 1 ≤ i ≤ n,
by assumption.
Since f is dierentiable at X0 , which implies that
∑
n
f (X) − f (X0 ) = fxi (X0 )(xi − xi0 ) + E(X)|X − X0 |, (2.75)
i=1
where
lim E(X) = 0.
X→X0
∑
n
h(U) − h(U0 ) = fxi (X0 )(gi (U) − gi (U0 ))
i=1
+E(G(U))|G(U) − G(U0 )|. (2.76)
∑n
h(U) − h(U0 ) = i=1 fxi (X0 )(dU0 gi )(U − U0 )
∑n
+( i=1 fxi (X0 )Ei (U)) |U − U0 |
+ E(G(U))|G(U) − G(U0 |.
Since
lim E(G(U)) = lim E(X) = 0.
U→U0 X→X0
2.20. The Chain Rule 116
Example: Let
f (x, y, z) = 2x2 + 4xy + 3yz,
g1 (u, v) = u2 + v 2 , g2 (u, v) = u2 − 2v 2 , g3 (u, v) = uv,
and
h(u, v) = f (g1 (u, v), g2 (u, v), g3 (u, v)).
Let U0 = (1, −1) and
Then
fx (X0 ) = 4, fy (X0 ) = 5, fz (X0 ) = −3,
Since
g1 (u, v) = u2 + v 2 , g2 (u, v) = u2 − 2v 2 , g3 (u, v) = uv,
dU0 h = fx1 (X0 )dU0 g1 + fx2 (X0 )dU0 g2 + · · · + fxn (X0 )dU0 gn .
= 21 du + 9 dv.
Since
dU0 h = hu (U0 ) du + hv (U0 ) dv
we conclude that
hu (U0 ) = 21 and hv (U0 ) = 9. (2.77)
2.20. The Chain Rule 117
h(u, v) = 2[g1 (u, v)]2 + 4g1 (u, v)g2 (u, v) + 3g2 (u, v)g3 (u, v)
Hence,
Proof :Substituting
∂gi (U0 ) ∂gi (U0 ) ∂gi (U0 )
dU0 gi = du1 + du2 + · · · + dum , 1 ≤ i ≤ n,
∂u1 ∂u2 ∂um
into (2.74) and collecting multipliers of du1 , du2 ,
dum yields
...,
∑
m ∑
n
∂f (X ) ∂g (U )
d U0 h = 0 j 0
dui .
∂xj ∂ui
i=1 j=1
∂h ∑ ∂f ∂gj
n
= , 1 ≤ i ≤ m, (2.79)
∂ui ∂xj ∂ui
j=1
with the understanding that in calculating ∂h(U0 )/∂ui , ∂gj /∂ui is evaluated at U0
and ∂f /∂xj at X0 = G(U0 ).
∂h ∑ ∂f ∂gj
n
= , 1 ≤ i ≤ m, (2.80)
∂ui ∂xj ∂ui
j=1
2.21. Higher derivatives of composite functions 118
with the understanding that in calculating ∂h(U0 )/∂ui , ∂gj /∂ui is evaluated at U0
and ∂f /∂xj at X0 = G(U0 ). By replacing the symbol G with X = X(U); then we
write
h(U) = f (X(U))
and
∂h(U0 ) ∑ ∂f (X0 ) ∂xj (U0 )
n
= ,
∂ui ∂xj ∂ui
j=1
∂h ∑ ∂f ∂xj n
or simply = . (2.81)
∂ui ∂xj ∂ui
j=1
∑n ( )
∂2h ∂ ∂f ∂xj
∂uk ∂ui = j=1 ∂uk ∂xj ∂ui
∑n ∑n ( ) (2.82)
∂f ∂ 2 xj ∂xj ∂ ∂f
= j=1 ∂xj ∂uk ∂ui + j=1 ∂ui ∂uk ∂xj .
( )
∂ ∂f (X(U))
. (2.83)
∂uk ∂xj
∂f (X)
g(X) = ;
∂xj
Since
∂g ∂2f
= ,
∂xs ∂xs ∂xj
this yields
( ) ∑
n
∂ ∂f ∂ 2 f ∂xs
= .
∂uk ∂xk ∂xs ∂xj ∂uk
s=1
2.21. Higher derivatives of composite functions 119
∂2h ∑
n
∂f ∂ 2 xj ∑
n
∂xj ∑ ∂ 2 f ∂xs
n
= + ∂uk . (2.84)
∂uk ∂ui ∂xj ∂uk ∂ui ∂ui ∂xs ∂xj
j=1 j=1 s=1
To compute hui uk (U0 ) from this formula, we evaluate the partial derivatives of x1 ,
x2 , ..., xn U0 and those of f at X0 = X(U0 ). The formula is valid if x1 , x2 ,
at
..., xn and their rst partial derivatives are dierentiable at U0 and f , fxi , fx2 ,
..., fxn and their rst partial derivatives are dierentiable at X0 .
x = r cos θ, y = r sin θ.
Suppose that f = f (x, y) is dierentiable on a set S, and let
∂h ∂f ∂x ∂f ∂y ∂f ∂f
= + = cos θ + sin θ (2.85)
∂r ∂x ∂r ∂y ∂r ∂x ∂y
∂h ∂f ∂x ∂f ∂y ∂f ∂f
= + = −r sin θ + r cos θ ,
∂θ ∂x ∂θ ∂y ∂θ ∂x ∂y
where fx and fy are evaluated at (x, y) = (r cos θ, r sin θ).
Example: Suppose that fx and fy just calculated are dierentiable on an open set
in R . Dierentiating (2.85) with respect to r yields
S 2
( ) ( )
∂2h ∂ ∂f ∂ ∂f
∂r 2
= cos θ ∂r ∂x + sin θ ∂r ∂y
(2.86)
( ) ( )
∂ 2 f ∂x ∂ 2 f ∂y ∂ 2 f ∂x ∂ 2 f ∂y
= cos θ ∂x2 ∂r
+ ∂y ∂x ∂r + sin θ ∂x ∂y ∂r + ∂y 2 ∂r
.
if (x, y) ∈ S . Since
∂x ∂y ∂2f ∂2f
= cos θ, = sin θ, and =
∂r ∂r ∂x ∂y ∂y ∂x
if (x, y) ∈ S . The equation (2.86) yields
∂2h 2
2 ∂ f ∂2f 2
2 ∂ f
= cos θ + 2 sin θ cos θ + sin θ .
∂r2 ∂x2 ∂x ∂y ∂y 2
Dierentiating (2.85) with respect to θ yields
( ) ( )
∂2h ∂f ∂f ∂ ∂f ∂ ∂f
= − sin θ + cos θ + cos θ + sin θ
∂θ ∂r ∂x ∂y ∂θ ∂x ∂θ ∂y
( 2 2
)
∂f ∂f ∂ f ∂x ∂ f ∂y
= − sin θ + cos θ + cos θ +
∂x ∂y ∂x2 ∂θ ∂y ∂x ∂θ
( 2 2
)
∂ f ∂x ∂ f ∂y
+ sin θ + 2 .
∂x ∂y ∂θ ∂y ∂θ
2.21. Higher derivatives of composite functions 120
Since
∂x ∂y
= −r sin θ and = r cos θ,
∂θ ∂θ
it follows that
( )
∂2h ∂f ∂f ∂2f ∂2f
= − sin θ + cos θ − r sin θ cos θ −
∂θ ∂r ∂x ∂y ∂x2 ∂y 2
∂2f
+ r(cos2 θ − sin2 θ) .
∂x∂y
∑
n
h′ (t0 ) = fxj (X(t0 ))x′j (t0 ). (2.87)
j=1
∑
n
f (X2 ) − f (X1 ) = fxi (X0 )(xi2 − xi1 ) = (dX0 f )(X2 − X1 ) (2.88)
i=1
Proof : An equation of L is
h(t) = f (X(t))
t0 ∈ (0, 1).
for some Since h(1) = f (X2 ) and h(0) = f (X1 ), this implies (2.88) with
X0 = X(t0 ), i.e.,
∑
n
f (X2 ) − f (X1 ) = fxi (X0 )(xi2 − xi1 ) = (dX0 f )(X2 − X1 ).
i=1
Theorem: If fx1 , fx2 , ..., f xn are identically zero in an open region S of Rn , then
f is constant in S.
X0 , X1 , . . . , Xn = X
such that the line segment Li from Xi−1 to Xi is in S , 1 ≤ i ≤ n. From mean value
theorem
∑
n
f (Xi ) − f (Xi−1 ) = e i f )(Xi − Xi−1 ),
(dX
i=1
where e
X is on Li and therefore in S.
Therefore,
e i ) = fx (X
fxi (X e i ) = · · · = fxn (X
e i ) = 0,
2
so the function
h(t) = f (X(t))
is dened for 0 ≤ t ≤ 1.
We know that
∑
n
h′ (t) = fxi (X(t)(xi − xi0 ).
i=1
2.22. rth Dierential 122
( n )
∑
n
∂ ∑ ∂f (X(t))
h′′ (t) = (xi − xi0 ) (xj − xj0 )
∂xj ∂xi
j=1 i=1
∑n
∂ 2 f (X(t))
= (xi − xi0 )(xj − xj0 )
∂xj ∂xi
i,j=1
If fx1 , fx2 , ..., fxn are dierentiable in Bρ (X0 ). Continuing in this way, we see
that
∑
n
∂ r f (X(t))
h(r) (t) = (xi − xi1 ,0 )(xi2 − xi2 ,0 )
∂xir ∂xir−1 · · · ∂xi1 1
i1 ,i2 ,...,ir =1
· · · (xir − xir ,0 )
(r)
∑
n
∂ r f (X0 )
d X0 f = dxi1 dxi2 · · · dxir , (2.89)
∂xir ∂xir−1 · · · ∂xi1
i1 ,i2 ,...,ir =1
where dx1 , dx2 , . . . , dxn are the dierentials, that is, dxi is the function whose value
point in R is the ith coordinate of the point. For convenience, we dene
at a
n
(0)
(dX0 f ) = f (X0 ).
(1)
Notice that d X0 f = d X0 f .
Remark: Suppose that r≥1 and all partial derivatives of f of order ≤ r−1 are
dierentiable in a neighborhood of X0 , the value of
∂ r f (X0 )
∂xir ∂xir−1 · · · ∂xi1
depends only on the number of times f is dierentiated with respect to each variable,
and not on the order in which the dierentiations are performed.
(r)
∑ r! ∂ r f (X0 )
d X0 f = (dx1 )r1 (dx2 )r2 · · · (dxn )rn , (2.90)
r
r1 !r2 ! · · · rn ! ∂x1 ∂xr22 · · · ∂xrnn
r1
2.22. rth Dierential 123
∑
where r indicates summation over all ordered n-tuples (r1 , r2 , . . . , rn ) of nonneg-
ative integers such that
r1 + r2 + · · · + rn = r
and ∂xri i is omitted from the denominators of all terms in (2.90) for which ri = 0.
In particular, if n = 2,
r ( ) r
∑
(r) r ∂ f (x0 , y0 )
d X0 f = (dx)j (dy)r−j .
j ∂xj ∂y r−j
j=0
Example: Let
1
f (x, y) = ,
1 + ax + by
where a and b are constants.
Then
∂ r f (x, y) r aj br−j
= (−1) r! ,
∂xj ∂y r−j (1 + ax + by)r+1
so
r ( )
∑
(r) (−1)r r! r j r−j
d X0 f = r+1
a b (dx)j (dy)r−j
(1 + ax0 + by0 ) j
j=0
(−1)r r!
= (a dx + b dy)r
(1 + ax0 + by0 )r+1
if 1 + ax0 + by0 ̸= 0.
Let
∑
n
f (X) = exp − aj xj ,
j=1
Therefore,
( )
(r) r! ∑
(dX0 f )(Φ) = (−1) r
ar11 ar22 · · · arnn (dx1 )r1 (dx2 )r2 · · · (dxn )rn
r
r !r
1 2 ! · · · rn !
∑
n
× exp − aj xj0
j=1
∑
n
= (−1)r (a1 dx1 + a2 dx2 + · · · + an dxn )r exp − aj xj0
j=1
2.23. Taylor's Theorem for Functions of n Variables 124
∑
k
1 (r) 1 (k+1)
f (X) = (d f )(X − X) + (d e f )(X − X0 ) (2.91)
r! X0 (k + 1)! X
r=0
for some e
X on L distinct from X0 and X.
Proof : Dene
h(t) = f (X0 + t(X − X0 )). (2.92)
∑
k
h(r) (0) h(k+1) (τ )
h(1) = + , (2.93)
r! (k + 1)!
r=0
We have Φ = X − X0 ,
(r)
h(r) (0) = (dX0 f )(X − X0 ), 1 ≤ r ≤ k, (2.95)
( )
h(k+1) (τ ) = dk+1
e f (X − X0 ) (2.96)
X
where
e = X0 + τ (X − X0 )
X
is on L and distinct from X0 and X.
Substituting (2.94), (2.95), and (2.96) into (2.93) yields (2.91).
Let
1
f (x, y) = ,
1 + ax + by
where a and b are constants.
Then
∂ r f (x, y) r aj br−j
= (−1) r! ,
∂xj ∂y r−j (1 + ax + by)r+1
so
r ( )
∑
(r) (−1)r r! r j r−j
d X0 f = r+1
a b (dx)j (dy)r−j
(1 + ax0 + by0 ) j
j=0
(−1)r r!
= (a dx + b dy)r
(1 + ax0 + by0 )r+1
2.23. Taylor's Theorem for Functions of n Variables 125
if 1 + ax0 + by0 ̸= 0.
Example: The Taylor series with X0 = (0, 0) and Φ = (x, y) imply that if 1 + ax +
by > 0, then
1 ∑ k
(ax + by)k+1
= (−1)r (ax + by)r + (−1)k+1
1 + ax + by (1 + aτ x + bτ y)k+2
r=0
for some τ ∈ (0, 1). (Note that τ depends on k as well as (x, y).)
Remark: By analogy with the situation for functions of one variable, we dene the
k th Taylor polynomial of f about X0 by
∑
k
1 (r)
Tk (X) = (d f )(X − X0 ).
r! X0
r=0
1 (k+1)
f (X) = Tk (X) + (d e f )(X − X0 ).
(k + 1)! X
Theorem: Suppose that f and its partial derivatives of order ≤ k − 1 are dieren-
tiable in a neighborhood N of a point X0 in Rn and all k th-order partial derivatives
of f are continuous at X0 . Then
f (X) − Tk (X)
lim = 0. (2.97)
X→X0 |X − X0 |k
Proof : If ε > 0, there is a δ > 0 such that Bδ (X0 ) ⊂ N and all k th-order partial
derivatives of f satisfy the inequality
e
∂ k f (X) ∂ k f (X0 )
e ∈ Bδ (X0 ).
− < ε, X (2.98)
∂xik ∂xik−1 · · · ∂xi1 ∂xik ∂xik−1 · · · ∂xi1
Now suppose that X ∈ Bδ (X0 ). From Taylor series expansion, with k replaced by
k − 1,
1 (k)
f (X) = Tk−1 (X) + (d e f )(X − X0 ), (2.99)
k! X
where e is some point on the line segment from X0
X to X and is therefore in Bδ (X0 ).
We can rewrite (2.99) as
1 [ (k) (k)
]
f (X) = Tk (X) + (d e f )(X − X0 ) − (dX0 f )(X − X0 ) . (2.100)
k! X
|f (X) − Tk (X)| nk ε
< , X ∈ Bδ (X0 ),
|X − X0 |k k!
where the coecients {ar1 r2 ...rn } are constants and the summation is over all n-tuples
of nonnegative integers (r1 , r2 , . . . , rn ) such that
r1 + r2 + · · · + rn = r,
p(x, y, z) = x2 + y 2 + z 2 + xy + xz + yz
1[ ]
p(x, y, z) = (x + y)2 + (y + z)2 + (z + x)2 .
2
so p is nonnegative, and p(x, y, z) = 0 if and only if
x + y = y + z = z + x = 0,
p1 (x, y, z) = x2 + y 2 + z 2 + 2xy
p1 (x, y, z) = (x + y)2 + z 2 ,
(r) (k)
dX0 f ≡ 0 (1 ≤ r ≤ k − 1), dX0 f ̸≡ 0. (2.103)
Then
(k)
• X0 is not a local extreme point of f unless d X0 f is semi-denite as a polynomial
in X − X0 .
(k)
• X0 is a local minimum point of f if d X0 f is positive denite, or a local maxi-
(k)
mum point if dX f is negative denite.
0
(k)
• Ifd X0 f is semidenite, then X0 may be a local extreme point of f, but it need
not be.
(2)
p(u, v) = (dX0 f )(u, v) = Au2 + 2Buv + Cv 2 ,
where A = fxx (x0 , y0 ), B = fxy (x0 , y0 ), and C = fyy (x0 , y0 ), so
D = AC − B 2 .
If D > 0, then A ̸= 0, and we can write
( ) ( )
2B B2 2 B2
2
p(u, v) = A u + uv + 2 v + C − v2
A A A
( )
B 2 D 2
= A u+ v + v .
A A
(2)
This cannot vanish unless u = v = 0. Hence, dX0 f is positive denite if A>0 or
negative denite if A < 0, and Theorem implies the rst part of the corollary.
If D < 0, there are three possibilities:
( )
1. A ̸= 0; then p(1, 0) = A and p −B
A,1 =
D
A.
( )
2. C ̸= 0; then p(0, 1) = C and p 1, − B
C =
D
C.
Example: If
2 +by 2
f (x, y) = eax .
We have
fx (x, y) = 2axf (x, y), fy (x, y) = 2byf (x, y),
so
fx (0, 0) = fy (0, 0) = 0,
and (0, 0) is a critical point of f.
To apply Corollary, we calculate
Integral Calculus
1
g(x) =
x(x − 1)
is locally integrable on (−∞, 0), (0, 1), and (1, ∞).
The function
√
h(x) = x
is locally integrable on [0, ∞).
If f is locally integrable on [a, b), we dene
∫ b ∫ c
f (x) dx = lim f (x) dx (3.1)
a c→b− a
if the limit exists (nite). To include the case where b = ∞, we adopt the convention
that ∞− = ∞.
Remarks:
• The limit in (3.1) always exists if [a, b) is nite and f is locally integrable and
bounded on [a, b).
• In this case, the denition of Riemann integral and locally integrable function
∫b
assign the same value to
a f (x) dx no matter how f (b) is dened. However,
the limit may also exist in cases where b=∞ or b < ∞ and f is unbounded
as x approaches b from the left.
3.1. Locally Integrable Functions 130
Remarks:
∫b
• We also say in this case that f is integrable on [a, b) and that a f (x) dx exists .
If the limit in (3.1) does not exist (nite), we say that the improper integral
∫b
a f (x) dx diverges , and f is nonintegrable on [a, b).
∫c ∫b
• In particular, if limc→b− a f (x) dx = ±∞, we say that
a f (x) dx diverges to
±∞, and we write
∫ b ∫ b
f (x) dx = ∞ or f (x) dx = −∞,
a a
∫ b ∫ b
f (x) dx = lim f (x) dx
a c→a+ c
∫ b ∫ α ∫ b
f (x) dx = f (x) dx + f (x) dx,
a a α
where a < α < b, provided that both improper integrals on the right exist (nite).
∫b
Remarks: The existence and value of
a f (x) dx according to the above denition
do not depend on the particular choice of α in (a, b).
1
F (x) = x2 sin
x
on [−2/π, 0).
3.1. Locally Integrable Functions 131
Hence,
∫
c
1 c 1 4 2
f (x) dx = x sin = c2 sin + 2
−2/π x −2/π c π
∫ 0 ( )
2 1 4 4
f (x) dx = lim c sin + 2 = 2 .
−2/π c→0− c π π
However, this is not an improper integral, even though f (0) is not dened and
cannot be dened so as to make f continuous at 0. If we dene f (0) arbitrarily
(say f (0) = 10), then f ∫is bounded on the closed interval [−2/π, 0] and continuous
0
except at 0. Therefore, −2/π f (x) dx exists and equals 4/π 2 as a proper integral, in
the sense of denition of improper integral.
Hence,
∫ {
1
−p (1 − p)−1 , p < 1,
(1 − x) dx =
0 ∞, p ≥ 1.
Hence,
∫ ∞ {
(p − 1)−1 , p > 1,
x−p dx =
1 ∞, p ≤ 1.
∫ ∫ c
c
1 1 c
1 1
2 1
log dx = − log x dx = − (log x) = − (log c)2 .
1 x x 1 x 2 1 2
Hence, ∫ c
1 1
lim log dx = −∞,
c→∞ 1 x x
so ∫ ∞
1 1
log dx = −∞.
1 x x
The function f (x) = log x is locally integrable on (0, 1], but unbounded as
x → 0+. Since
∫ 1
1
lim log x dx = lim (x log x − x) = −1 − lim (c log c − c) = −1,
c→0+ c c→0+ c c→0+
Denition ?? yields ∫ 1
log x dx = −1.
0
The function f (x) = cos x is locally integrable on [0, ∞) and
∫ c
lim cos x dx = lim sin c
c→∞ 0 c→∞
∫∞
does not exist; thus,
0 cos x dx diverges, but not to ±∞.
In connection with Denition
∫α ∫b ??, it is important to recognize ∫that the improper
integrals
a f (x) dx and
α f (x) dx must converge separately for
b
a f (x) dx to con-
verge. For example, the existence of the symmetric limit
∫ R
lim f (x) dx,
R→∞ −R
∫∞ ∫∞
which is called the principal value of −∞ f (x) dx, does not imply that −∞ f (x) dx
converges; thus,
∫ R
lim x dx = lim 0 = 0,
R→∞ −R R→∞
∫∞ ∫0 ∫∞
but
0 x dx and
−∞ x dx diverge and therefore so does −∞ x dx.
Furthermore,
∫ b ∫ b
(c1 f1 + c2 f2 + · · · + cn fn )(x) dx = c1 f1 (x) dx
a a
∫ b
+c2 f2 (x) dx
a
∫ b
+ · · · + cn fn (x) dx.
a
∫b
Theorem: If f is nonnegative and locally integrable on [a, b), then
a f (x) dx con-
verges if the function ∫ x
F (x) = f (t) dt
a
∫b
is bounded on [a, b), and
a f (x) dx = ∞ if it is not.
These are the only possibilities, and
∫ b
f (t) dt = sup F (x)
a a≤x<b
in either case.
∫ b
f (x) dx > −∞
a
then
∫b ∫b
1.
a f (x) dx < ∞ if
a g(x) dx < ∞
∫b ∫b
2.
a g(x) dx = ∞ if
a f (x) dx = ∞.
Proof : Since
0 ≤ f (x) ≤ g(x), a ≤ x < b,
we have ∫ x ∫ x
f (t) dt ≤ g(t) dt, a ≤ x < b.
a a
So ∫ x ∫ x
sup f (t) dt ≤ sup g(t) dt.
a≤x<b a a≤x≤b a
∫b
If
a g(x) dx < ∞, the right side of this inequality is nite by the previous Theorem,
so the left side is also.
∫b
This implies that f (x) dx < ∞.
a ∫b
The proof is by contradiction. g(x) dx < ∞, then (1)
If implies that
∫b ∫b a
∫ 1
2 + sin πx
I= dx.
0 (1 − x)p
3.1. Locally Integrable Functions 135
Solution: We are going to show that the improper integral converges if p < 1.
Since
2 + sin πx 3
0< ≤ , 0 ≤ x < 1.
(1 − x) p (1 − x)p
We have ∫ 1
3 dx
< ∞, p < 1.
0 (1 − x)p
∫ 1
2 + sin πx
I= dx.
0 (1 − x)p
1 2 + sin πx
0< ≤ , 0 ≤ x < 1,
(1 − x)p (1 − x)p
and ∫ 1
dx
= ∞, p ≥ 1.
0 (1 − x)p
∫ c ∫ a1 ∫ c
f (x) dx = f (x) dx + f (x) dx.
a a a1
∫ a1
Since f (x) dx is a proper integral, on letting c → b− we conclude that if
a ∫b ∫b
either of the improper integrals
a f (x) dx and
a1 f (x) dx converges then so does
the other, and in this case
∫ b ∫ a1 ∫ b
f (x) dx = f (x) dx + f (x) dx.
a a a1
From this, you can see that if f (x) ≥ 0 on some subinterval [a1 , b) of [a, b), but
not necessarily for all x in [a, b), we can still use the convention introduced earlier
∫b
for positive functions; that is, we can write f (x) dx < ∞ if the improper integral
∫b a
converges or
a f (x) dx = ∞ if it diverges.
Theorem: Suppose that f and g are locally integrable on [a, b), g(x) > 0 and
f (x) ≥ 0 on some subinterval [a1 , b) of [a, b), and
f (x)
lim = M. (3.3)
x→b− g(x)
∫b ∫b
• If 0 < M < ∞, then
a f (x) dx and
a g(x) dx converge or diverge together.
∫b ∫b
• If M =∞ and
a g(x) dx = ∞, then
a f (x) dx = ∞.
∫b ∫b
• If M =0 and
a g(x) dx < ∞, then
a f (x) dx < ∞.
M f (x) 3M
0< < < , a2 ≤ x < b,
2 g(x) 2
and therefore
M 3M
g(x) < f (x) < g(x), a2 ≤ x < b. (3.4)
2 2
The rst inequality in (3.4) imply that
∫ b ∫ b
g(x) dx < ∞ if f (x) dx < ∞.
a2 a2
∫ b ∫ b
f (x) dx < ∞ if g(x) dx < ∞.
a2 a2
∫b ∫b
Therefore,
a2 f (x) dx and a2 g(x) dx converge or diverge together, and in the latter
case they must diverge to ∞, since their integrands are nonnegative. If M = ∞,
there is a point a2 in [a1 , b) such that
f (x) ≥ g(x), a2 ≤ x ≤ b,
∫b
We have
a f (x) dx = ∞. If M = 0, there is a point a2 in [a1 , b) such that
f (x) ≤ g(x), a2 ≤ x ≤ b,
∫b
so we have
a f (x) dx < ∞.
3.2. Absolute integrability 137
Example: Since
sin x 1
xp ≤ xp
∫∞
and
1 x−p dx < ∞ if p > 1.
The comparison theorem implies that
∫ ∞
| sin x|
dx < ∞, p > 1.
1 xp
The function
sin x
f (x) =
xp
is absolutely integrable on [1, ∞) if p > 1.
∑k−1 ∫ (j+1)π
> 1
j=1 (j+1)π jπ | sin x| dx.
But
∫ (j+1)π ∫ π
| sin x| dx = sin x dx = 2,
jπ 0
so (3.5) implies that
∫
2∑ 1
k−1
kπ
| sin x|
dx > . (3.6)
1 x π j+1
j=1
However,
∫ j+2
1 dx
≥ , j = 1, 2, . . . ,
j+1 j+1 x
so (3.6) implies that
∫ k−1 ∫
kπ
| sin x| 2 ∑ j+2 dx
>
1 x π x
j=1 j+1
∫
2 k+1 dx 2 k+1
= = log .
π 2 x π 2
3.3. Nonoscillatory and Oscillatory Functions 138
∫b ∫b
Theorem: If f is locally integrable on [a, b) and |f (x)| dx < ∞,
a then
a f (x) dx
converges; that is, an absolutely convergent integral is convergent.
Proof : If
g(x) = |f (x)| − f (x).
Then
0 ≤ g(x) ≤ 2|f (x)|
∫b
and
a g(x) dx < ∞, because of comparison theorem and the absolute integrability
of f. Since
f = |f | − g,
∫b
Due to comparison test, we can conclude that
a f (x) dx converges.
The comparison test implies that the integral on the right converges absolutely
∫b
as c → b−, since
a |g ′ (x)| dx < ∞ by assumption, and
where M is an upper bound for |F | on [a, b). Moreover, (3.9) and the boundedness
of F imply that limc→b− F (c)g(c) = 0.
Remark: Dirichlet's test can also be used to show that certain integrals diverge.
3.6 Rectangles in Rn
The
S1 × S2 × · · · × Sn
of subsets S1 , S2 , . . . , Sn of R is the set of points (x1 , x2 , . . . , xn ) in Rn such that
x1 ∈ S1 , x2 ∈ S2 , . . . , xn ∈ Sn . For example, the Cartesian product of the two closed
intervals
[a1 , b1 ] × [a2 , b2 ] = {(x, y) : a1 ≤ x ≤ b1 , a2 ≤ y ≤ b2 }
is a rectangle in R2 with sides parallel to the x- and y -axes.
The Cartesian product of three closed intervals
b2
a2
x
a1 b1
The content of R is
and
Pr : ar = ar0 < ar1 < · · · < armr = br
is a partition of [ar , br ], 1 ≤ r ≤ n, then the set of all rectangles in Rn that can be
written as
P = P1 × P2 × · · · × Pn . (3.11)
3.7. Riemann Sum in Rn 142
Put another way, ∥P∥ is the largest of the edge lengths of all the subrectangles
in P. Geometrically, a rectangle in R2 is partitioned by drawing horizontal and ver-
tical lines through it; in R , by drawing planes through it parallel to the coordinate
3
b2
a2
x
a1 b1
• IfP = P1 ×P2 ×· · ·×Pn and P′ = P1′ ×P2′ ×· · ·×Pn′ are partitions of the same
rectangle, then P is a renement of P if Pi is a renement of Pi , 1 ≤ i ≤ n.
′ ′
|σ − L| < ε.
∫
Remarks: The integral
R f (X)dX is also written as
∫ ∫
f (x, y) d(x, y) (n = 2), f (x, y, z) d(x, y, z) (n = 3),
R R
or ∫
f (x1 , x2 , . . . , xn ) d(x1 , x2 , . . . , xn ) (n arbitrary).
R
Here dX does not stand for the dierential of X.
It merely identies x1 , x2 , ..., xn , the components of X, as the variables of
∫
integration.
∫ To avoid this minor inconsistency, some authors write simply
Rf
rather than
R f (X) dX.
As in the case where n = 1, we will say simply integrable or integral when
we mean Riemann integrable or Riemann integral. If n ≥ 2, we call the integral
of above denition a multiple integral ; for n = 2 and n = 3 we also call them double
and triple integrals , respectively. When we wish to distinguish between multiple
integrals and the integral we studied in Chapter (n = 1), we will call the latter an
ordinary integral.
∫
Example: Find
R f (x, y) d(x, y), where
R = [a, b] × [c, d]
and
f (x, y) = x + y.
and
P2 : c = y0 < y1 < · · · < ys = d.
A typical Riemann sum of f over P = P1 × P2 is given by
∑
r ∑
s
σ= (ξij + ηij )(xi − xi−1 )(yj − yj−1 ), (3.12)
i=1 j=1
xi + xi−1 yj + yj−1
xi = and yj = , (3.14)
2 2
and (3.13) implies that
∑r ∑s
σ = i=1 j=1 (xi + y j )(xi − xi−1 )(yj − yj−1 )
∑r ∑s [ ]
+ i=1 j=1 (ξij − xi ) + (ηij − y j ) (3.17)
∑
r ∑
s
(xi − xi−1 ) = b − a, (yj − yj−1 ) = d − c (3.18)
i=1 j=1
and
∑
r ∑
s
(x2i − x2i−1 ) = b2 − a2 , (yj2 − yj−1
2
) = d2 − c2 . (3.19)
i=1 j=1
Because of (3.15) and (3.16) the absolute value of the second sum in (3.17) does not
exceed
[ r ]
∑
r ∑
s ∑
∥P∥ (xi − xi−1 )(yj − yj−1 ) = ∥P∥ (xi − xi−1 )
j=1 j=1 i=1
∑s
(yj − yj−1 )
j=1
= ∥P∥(b − a)(d − c)
3.8. Riemann Integral in Rn 145
∑r ∑
s
σ − (x + y )(x − x )(y − y ) ≤ ∥P∥(b − a)(d − c). (3.20)
i j i i−1 j j−1
i=1 j=1
Similarly,
∑
r ∑
s
b−a 2
y j (xi − xi−1 )(yj − yj−1 ) = (d − c2 ).
2
i=1 j=1
d − c b − a
σ − (b − a ) −
2 2
(d − c ) ≤ ∥P∥(b − a)(d − c).
2 2
2 2
Since the right side can be made as small as we wish by choosing ∥P∥ suciently
small, ∫
1[ ]
(x + y) d(x, y) = (d − c)(b2 − a2 ) + (b − a)(d2 − c2 ) .
R 2
∑
k
σ= f (Xj )V (Rj )
j=1
M
|f (X) − f (Xi )| ≥ (3.22)
V (Ri )
for some X in Ri , because if this were not so, we would have
M
|f (X) − f (Xj )| < , X ∈ Rj , 1 ≤ j ≤ k.
V (Rj )
M
|f (X)| ≤ max |f (Xj )| + 1 ≤ j ≤ k, X ∈ R,
V (Rj )
∑
n
σ′ = f (X′j )V (Rj )
j=1
{
Xj , j ̸= i,
X′j =
X, j = i.
Since
|σ − σ ′ | = |f (X) − f (Xi )|V (Ri ),
(3.22) implies (3.21).
∑
k
S(P) = Mj V (Rj ).
j=1
3.9. Upper and Lower Integrals 147
Then
1. The upper sum S(P) of f over P is the supremum of the set of all Riemann
sums of f over P.
2. The lower sum s(P) of f over P is the inmum of the set of all Riemann sums
of f over P.
Remarks: If
m ≤ f (X) ≤ M for X in R,
then
mV (R) ≤ s(P) ≤ S(P) ≤ M V (R);
∫ ∫
therefore,
R f (X) dX and R f (X) dX exist, are unique, and satisfy the inequalities
∫
mV (R) ≤ f (X) dX ≤ M V (R)
R
and ∫
mV (R) ≤ f (X) dX ≤ M V (R).
R
∫ ∫
f (x, y) d(x, y) and f (x, y) d(x, y) (n = 2),
R R
3.9. Upper and Lower Integrals 148
∫ ∫
f (x, y, z) d(x, y, z) and f (x, y, z) d(x, y, z) (n = 3),
R R
or ∫
f (x1 , x2 , . . . , xn ) d(x1 , x2 , . . . , xn )
R
and ∫
f (x1 , x2 , . . . , xn ) d(x1 , x2 , . . . , xn ) (n arbitrary).
R
∫ ∫
Example: Find
R f (x, y) d(x, y) and R f (x, y) d(x, y), with R = [a, b] × [c, d] and
f (x, y) = x + y.
r ∑
∑ s
S(P) = (xi + yj )(xi − xi−1 )(yj − yj−1 ) (3.23)
i=1 j=1
∑r ∑ s
s(P) = (xi−1 + yj−1 )(xi − xi−1 )(yj − yj−1 ). (3.24)
i=1 j=1
By substituting
1
xi + yj = [(xi + xi−1 ) + (yj + yj−1 ) + (xi − xi−1 ) + (yj − yj−1 )]
2
into (3.23). We nd that
1
S(P) = (Σ1 + Σ2 + Σ3 + Σ4 ), (3.25)
2
where
∑r ∑s
i=1 (xi − xi−1 ) − yj−1 ) = (b2 − a2 )(d − c),
Σ1 = 2 2
j=1 (yj
∑r ∑s
i=1 (xi − xi−1 ) − yj−1 = (b − a)(d2 − c2 ),
Σ2 = 2 2 )
j=1 (yj
∑r ∑
Σ3 = − xi−1 )2 sj=1 (yj − yj−1 ) ≤ ∥P∥(b − a)(d − c),
i=1 (xi
∑r ∑s
i=1 (xi − xi−1 ) j=1 (yj − yj−1 )
Σ4 = 2 ≤ ∥P∥(b − a)(d − c).
3.9. Upper and Lower Integrals 149
where
(d − c)(b2 − a2 ) + (b − a)(d2 − c2 )
I= .
2
From this, we see that
∫
(x + y) d(x, y) = I.
R
After substituting
1
xi−1 + yj−1 = [(xi + xi−1 ) + (yj + yj−1 ) − (xi − xi−1 ) − (yj − yj−1 )]
2
into (3.24), a similar argument shows that
So ∫
(x + y) d(x, y) = I.
R
and
∑
n
rj
s(P) ≤ s(P′ ) ≤ s(P) + 2M V (R) ∥P∥. (3.27)
bj − aj
j=1
∫ ∫
f (X) dX ≤ f (X) dX.
R R
3.9. Upper and Lower Integrals 150
∫ ∫ ∫
f (X) dX = f (X) dX = f (X) dX.
R R R
∫ ∫
f (X) dX ≤ S(P) < f (X) dX + ε
R R
and ∫ ∫
f (X) dX ≥ s(P) > f (X) dX − ε
R R
if ∥P∥ < δ.
∫ ∫
f (X) dX = f (X) dX.
R R
∫ ∫
f (X) dX = f (X) dX = L,
R R
∫
f (X) dX = L.
R
and
∑
m
V (Tj ) < ε. (3.29)
j=1
Example: Since the empty set is contained in every rectangle, the empty set has
zero content.
If E consists of nitely many points X1 , X2 , ..., Xm , then Xj can be enclosed
in a rectangle Tj such that
ε
V (Tj ) < , 1 ≤ j ≤ m.
m
∪m ∑m
Then E⊂ j=1 Tj and j=1 V (Tj ) < ε hold, so E has zero content.
Example: Any bounded set E with only nitely many limit points has zero content.
To see this, we rst observe that if E has no limit points, then it must be nite,
by the BolzanoWeierstrass theorem, and therefore must have zero content.
Now suppose that the limit points of E are X1 , X2 , ..., Xm . Let R1 , R2 , ...,
Rm be rectangles such that Xi ∈ Ri0 and
ε
V (Ri ) < , 1 ≤ i ≤ m. (3.30)
2m
The set of points of E that are not in ∪m
j=1 Rj has no limit points (why?) and, being
bounded, must be nite (again by the BolzanoWeierstrass theorem).
If this set contains p points, then it can be covered by rectangles R1′ , R2′ , ...,
Rp′ with
ε
V (Rj′ ) < , 1 ≤ j ≤ p. (3.31)
2p
Now,
(m )
∪ ∪ ∪
p
E⊂ Ri Rj′
i=1 j=1
∑
m ∑
p
V (Ri ) + V (Rj′ ) < ε.
i=1 j=1
3.11. Integral Over Bounded Set 152
(that is, the set {(x, y) : y = f (x), a ≤ x ≤ b}), has zero content in R2 .
Lemma: The union of nitely many sets with zero content has zero content.
y = x, 0≤x≤1
Since the line segment has zero content, f is integrable on R.
Area
∫ and volume as integrals: If S is a bounded subset of Rn and the integral
Thus, ∫
V (S) = dX.
S
2 cos θ cos ϕ
X = G(θ, ϕ) = 2 sin θ cos ϕ .
2 sin ϕ
where
bi − ai = L, 1 ≤ i ≤ m.
Suppose that we partition C into Nm smaller cubes by partitioning each of the
intervals [ai , bi ] into N equal subintervals. Let R1 , R2 , ...,Rk be the smaller cubes
so produced that contain points of D, and select points X1 , X2 , . . . , Xk such that
Xi ∈ D ∩ Ri , 1 ≤ i ≤ k . If Y ∈ D ∩ Ri , then (3.35) implies that
Since Xi and Y are both in the cube Ri with edge length L/N ,
√
L m
|Xi − Y| ≤ .
N
This and (3.36) imply that
√
ML m
|G(Xi ) − G(Y)| ≤ ,
N
which in turn implies that G(Y) lies in a cube ei
R in Rn centered at G(Xi ), with
√
sides of length 2M L m/N . Now
∑
k ( √ )n ( √ )n
2M L m m 2M L m √
ei ) = k
V (R ≤N = (2M L m)n N m−n .
N N
i=1
Since n > m, we can make the sum on the left arbitrarily small by taking N
suciently large. Therefore, S has zero content.
Example: Let
S = {(x, y) : x2 + y 2 = 1, x ≥ 0}.
The set S is bounded by a semicircle and a line segment, both dierentiable curves
in R .
2
Let {
(1 − x2 − y 2 )1/2 , (x, y) ∈ S, y ≥ 0,
f (x, y) =
−(1 − x2 − y 2 )1/2 , (x, y) ∈ S, y < 0.
Then f is continuous on S except on the line segment
y = 0, 0 ≤ x < 1,
Example: Let
S1 = {(x, y) : 0 ≤ x ≤ 1, 0 ≤ y ≤ 1 + x}
S2 = {(x, y) : −1 ≤ x ≤ 0, 0 ≤ y ≤ 1 − x}
3.13. Iterated Integrals 156
y=1−x y=1+x
x
−1 1
Then
S1 ∩ S2 = {(0, y) : 0 ≤ y ≤ 1}
has zero content.
Hence, by using corollary implies that if f is integrable on S1 and S2 , then f is
also integrable over
S = S1 ∪ S2 = {(x, y) : −1 ≤ x ≤ 1, 0 ≤ y ≤ 1 + |x|}
and ∫ ∫ ∫
f (X) dX = f (X) dX + f (X) dX.
S1 ∪S2 S1 S2
∫ d ∫ b
I1 = dy f (x, y) dx.
c a
3.13. Iterated Integrals 157
∫ d
G(x) = f (x, y) dy, a ≤ x ≤ b,
c
Dening
∫ b ∫ b (∫ d )
I2 = G(x) dx = f (x, y) dy dx,
a a c
which we usually write as
∫ b ∫ d
I2 = dx f (x, y) dy.
a c
Example: Let
f (x, y) = x + y
and R = [0, 1] × [1, 2]. Then
∫ 1 ∫ 1
1
F (y) = (x + y) dx = + y
f (x, y) dx =
0 0 2
∫ 2 ∫ 2( ) ( )
1 y y 2 2
I1 = F (y) dy = + y dy = + = 2.
1 1 2 2 2 1
Also,
∫ ( )
2
y 2 2 3
G(x) = (x + y) dy =
xy + =x+ ,
1 2 y=1 2
∫ 1 ∫ 1( ) ( 2 )
3 x 3x 1
I2 = G(x) dx = x+ dx = + = 2.
0 0 2 2 2 0
∫ b
F (y) = f (x, y) dx
a
∫ d ∫
F (y) dy = f (x, y) d(x, y); (3.38)
c R
that is,
∫ d ∫ b ∫
dy f (x, y) dx = f (x, y) d(x, y). (3.39)
c a R
3.13. Iterated Integrals 158
∫ b ∫ d ∫ d ∫ b
dx f (x, y) dy = dy f (x, y) dx,
a c c a
∫d ∫b
provided that f (x, y) dy exists for a ≤ x ≤ b and a f (x, y) dx exists for c ≤ y ≤ d.
c
In particular, these hypotheses hold if f is continuous on [a, b] × [c, d].
Since f also satises the hypotheses of Fubini's Theorem with x and y interchanged,
we can calculate the double integral from the iterated integral in which the integra-
tions are performed in the opposite order.
Thus,
∫ ∫ 1∫ 2
(x + y) d(x, y) = dx (x + y) dy
∫ 1 [( ) ]
R 0 1
y 2 2
= xy + dx
0 2 y=1
∫ 1( ) ( 2 )
3 x 3x 1
= x+ dx = + = 2.
0 2 2 2 0
∫d ∫b ∫
Remark: If
c dy a f (x, y) dx exists then so does
R f (x, y) d(x, y). However, this
need not to be true.
then ∫ 1
f (x, y) dx = y, 0 ≤ y ≤ 1,
0
3.13. Iterated Integrals 159
and ∫ 1 ∫ 1 ∫ 1
1
dy f (x, y) dx = y dy = .
0 0 0 2
However, f is not integrable on R.
∫
Fp (xp+1 , xp+2 , . . . , xn ) = f (x1 , x2 , . . . , xn ) d(x1 , x2 , . . . , xp )
I1 ×I2 ×···×Ip
Then
∫
Fp (xp+1 , xp+2 , . . . , xn ) d(xp+1 , xp+2 , . . . , xn )
Ip+1 ×Ip+2 ×···×In
∫
exists and equals
R f (X) dX.
Theorem: Ij = [aj , bj ], 1 ≤ j ≤ n,
Let and suppose that f is integrable on
R = I1 × I2 × · · · × In .
Suppose also that the integrals
∫
Fp (xp+1 , . . . , xn ) = f (X) d(x1 , x2 , . . . , xp ), 1 ≤ p ≤ n − 1,
I1 ×I2 ···×Ip
∫ bn ∫ bn−1 ∫ b2 ∫ b1
dxn dxn−1 · · · dx2 f (X) dx1
an an−1 a2 a1
∫
exists and equals
R f (X) dX.
3.13. Iterated Integrals 160