0% found this document useful (0 votes)
78 views11 pages

The Definition of The Riemann Integral

The document defines the Riemann integral of a bounded function over an interval. It introduces partitions of the interval and uses them to define the lower and upper Riemann sums. It proves that as partitions are refined, the lower sums increase and upper sums decrease. The Riemann integral is defined as the supremum of lower sums or equivalently the infimum of upper sums, provided these values are equal.

Uploaded by

Zidii Rajpoot
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
78 views11 pages

The Definition of The Riemann Integral

The document defines the Riemann integral of a bounded function over an interval. It introduces partitions of the interval and uses them to define the lower and upper Riemann sums. It proves that as partitions are refined, the lower sums increase and upper sums decrease. The Riemann integral is defined as the supremum of lower sums or equivalently the infimum of upper sums, provided these values are equal.

Uploaded by

Zidii Rajpoot
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 11

The Riemann Integral

The definition of the Riemann integral

Let f : [a, b] → R be a bounded function. By a partition we mean a set of points

a = x0 < x1 < x2 < · · · < xN −1 < xN = b.

These points are called mesh points. We will denote this partition by P . A partition Q is called a refinement of
P if P ⊂ Q. The graph of the the function f lies in a vertical strip in the xy-plane: {(x, y) : a ≤ x ≤ b}. This
strip consist of N panels {(x, y) : xi−1 ≤ x ≤ xi }. Loosely speaking, on this panel the graph of f varies between a
minimum height mi and a maximum height Mi . More precisely

mi = inf{f (x) | xi−1 ≤ x ≤ xi }, Mi = sup{f (x) | xi−1 ≤ x ≤ xi }.

We can now define the lower Riemann sum L(P, f ) and the upper Riemann sum U (P, f ):


N ∑
N
L(P, f ) := mi ∆xi , U (P, f ) := Mi ∆xi .
i=1 i=1

Note that
m(b − a) ≤ L(P, f ) ≤ U (P, f ) ≤ M (b − a) (1)
Suppose we refine the partition P by adjoining one more mesh point z. Suppose j is the integer such that
xj−1 ≤ z ≤ xj and let P̃ denote the refined partition: P̃ = P ∪ {z}. The extra point z divides the j th panel of the
original partition into a left panel and a right panel. Let

i = inf{f (x) | xi−1 ≤ x ≤ z},


mL MiL = sup{f (x) | xi−1 ≤ x ≤ z},

i = inf{f (x) | z ≤ x ≤ xi },
mR MiR = sup{f (x) | z ≤ x ≤ xi }.
When we compute the upper and lower Riemann sums for the refined partition we see that the only difference is
that the j th term is replaced by two new terms corresponding to the left and right panels:

j (z − xj−1 ) + mj (xj − z),


mL R
mi ∆xj becomes

Mi ∆xj becomes MjL (z − xj−1 ) + MjR (xj − z).


Hence

L(P, f ) − L(P̃ , f ) = (mj − mL


j )(z − xj−1 ) + (mj − mj )(xj − z),
R
(2)
U (P, f ) − U (P̃ , f ) = (Mj − MjL )(z − xj−1 ) + (Mj − MjR )(xj − z). (3)

Note that mj ≤ mL
j ≤ Mj ≤ Mj and mj ≤ mj ≤ Mj ≤ Mj . This implies
L R R

L(P, f ) ≤ L(P̃ , f ) ≤ U (P̃ , f ) ≤ U (P, f ).

Since we can think of a refinement Q of a partition P as being obtained by successively adding one point at a time
we have:

1
Lemma 1. Let Q be a refinement of P then

L(P, f ) ≤ L(Q, f ) ≤ U (Q, f ) ≤ U (P, f ). (4)

Note that equation (1) is actually a special case of equation (4).

Lemma 2. Let P1 and P2 be any two partitions of [a, b] then L(P1 , f ) ≤ U (P2 , f ) and hence

sup L(P, f ) ≤ inf U (P, f ).


P P

Proof. Let Q := P1 ∪ P2 , then it refines both P1 and P2 . Therefore, by lemma 1:

L(P1 , f ) ≤ L(Q, f ) ≤ U (Q, f ) ≤ U (P2 , f ).

We define ∆xi := xi − xi−1 and


µ(P ) := max ∆xi .
i

The number µ(P ) is called the mesh size of the partition P .


For later use we will also need another estimate:

Lemma 3. Let Q be a refinement of P obtained by adjoining K points, then

|L(P, f ) − L(Q, f )| ≤ K(M − m)µ(P ), |U (P, f ) − U (Q, f )| ≤ K(M − m)µ(P ).

The proof is simply a consequence of the fact that neither suprema nor infima over any subinterval can differ by more
than M − m. If we add an extra mesh point between xi−1 and xi then the contribution from subinterval [xi−1 , xi ] is
not changed by more than (Mi − mi )∆xi ≤ (M − m)µ(P ). So if we adjoin K points the change in either the lower
sum or the upper sum is no more in magnitude than K(M − m)µ(P ).

∫b ∫b
We define the lower Riemann integral a
f (x) dx and the upper Riemann integral a
f (x) dx as follows
∫ b ∫ b
f (x) dx = sup L(P, f ), and f (x) dx = inf U (P, f ).
a P a P

By lemma 2 the lower Riemann integral is less than or equal to the upper Riemann integral. We say that the function
f is Riemann integrable on [a, b] if its lower and upper Riemann integrals have the same value. In that case we denote
∫b
that common value by a f dx, called the Riemann integral of f on [a, b]. We now summarize

Definition. Let f : [a, b] → R be a bounded function.

M := sup{f (x) | a ≤ x ≤ b}, m := inf{f (x) | a ≤ x ≤ b}.

Let P denote the partition a = x0 < x1 < x2 < · · · < xN = b and define ∆xi := xi − xi−1 . The mesh or norm of this
partition is defined as |P | := min{∆xi | 1 ≤ i ≤ N }. We define

Mi := sup{f (x) | xi−1 ≤ x ≤ xi }, mi := inf{f (x) | xi−1 ≤ x ≤ xi }.

We define the upper Riemann sum and the lower Riemann sum respectively by

N ∑
N
U (P, f ) := Mi ∆xi , L(P, f ) := mi ∆xi ,
i=1 i=1

and the upper Riemann integral and the lower Riemann integral respectively by
∫ b ∫ b
f (x) dx := inf U (P, f ), f (x) dx := sup U (P, f ).
a P a P

2
We say that f is Riemann integrable on [a, b] if the upper and lower Riemann integrals are equal. Their common
value is then called Riemann integral and is denoted by
∫ b
f (x) dx.
a

We have the following important result:

Riemann Lemma. f : [a, b] → R is Riemann integrable iff for any ϵ > 0 there exists a partition Q such that
U (Q, f ) − L(Q, f ) < ϵ.

Proof. Suppose that f is Riemann integrable, then by the definitions of lower and upper Riemann integrals (which
now have the same value) there exist partitions P1 and P2 such that
∫ b ∫ b
L(P1 , f ) > f (x) dx − ϵ/2, U (P1 , f ) < f (x) dx + ϵ/2.
a a

Let Q = P1 ∪ P2 , then
( ∫ ) (∫ )
b b
ϵ ϵ
U (Q, f ) − L(Q, f ) ≤ U (P2 , f ) − L(P1 , f ) ≤ U (P2 , f ) − f (x) dx + f (x) dx − L(P1 , f ) < + = ϵ.
a a 2 2

Conversely, if for each ϵ > 0 we can find a partition Q such that U (Q, f ) − L(Q, f ) < ϵ then
∫ b ∫ b
0≤ f (x) dx − f (x) dx ≤ U (Q, f ) − L(Q, f ) < ϵ.
a a

Since ϵ is an arbitrarily small positive number, the upper and lower integrals must have the same value.
Notation. We use C[a, b] to denote the set of all continuous functions from [a, b] to R and R[a, b] to denote the set
of all functions that are Riemann integrable on [a, b].

Theorem. C[a, b] ⊂ R[a, b].

Proof. Let f ∈ C[a, b]. Then f must, in fact, be uniformly continuous. Given any ϵ > 0 we can find a δ > 0 such
that if u, v ∈ [a, b] with |u − v| < δ then |f (u) − f (v)| < ϵ/(b − a). Let P be any partition with mesh size µ(P ) < δ.
This means that on each panel Mi − mi < ϵ/(b − a). Therefore
∑ ∑
U (P, f ) − L(P, f ) = (Mi − mi )∆xi < [ϵ/(b − a)]∆xi = ϵ.
i i

Integrability now follows from the Riemann lemma.

Telescoping sum. Note that



N
[γj − γj−1 ] = γN − γ0 .
j=1

Such a sum is called a telescoping sum.

Theorem. Let f : [a, b] → R be a monotone function. Then f ∈ R[a, b].

Proof. We prove the case where f is increasing; the case where f is decreasing is handled similarly. Given any ϵ
choose a positive integer N such that (f (b) − f (a))(b − a)/N < ϵ. Choose the partition P with xi := a + i(b − a)/N

3
so that ∆xi = ∆x := (b − a)/N . From the fact that f is increasing it follows that Mi = f (xi ) and mi = f (xi−1 ).
Hence

N
U (P, f ) − L(P, f ) = (f (xi ) − f (xi−1 ))∆x
i=1

which is a telescoping sum that is equal to (f (xN ) − f (x0 ))∆x = (f (b) − f (a))(b − a)/N < ϵ. Integrability, once
again, follows from the Riemann lemma.

Riemann sums
Let P := {x0 , x1 , · · · , xN } be a partition. A set of points {t1 , t2 , · · · , tN } is called a marking of the partition P if
for each i we have xi−1 ≤ ti ≤ xi . We denote by P T the partition P together with the marking T . We call P T a
marked partition. Sometimes we will simplify the notation and denote P T by Π and define µ(Π) := µ(P ). Given
such a marked partition we define the corresponding Riemann sum as


N
S(Π, f ) = S(P T , f ) := f (ti ) ∆xi .
i=1

Clearly, since mi ≤ f (ti ) ≤ Mi , we have

L(P, f ) ≤ S(P T , f ) ≤ U (P, f ).

Theorem. Let {Pk | k ∈ N} be a family of partitions for [a, b] such that

lim µ(Pk ) = 0.
k→∞

For each k let Tk be a marking for Pk . If f is Riemann integrable on [a, b] then


∫ b
lim S(PkTk , f ) = f (x) dt.
k→∞ a

Proof. Suppose that m ≤ f (x) ≤ M for all x ∈ [a, b]. We assume M > m; the case M = m would make the
proof trivial. By the Riemann lemma, given any ϵ > 0 we can find a partition Q such that U (Q, f ) − L(Q, f ) < ϵ/2.
Suppose that Q consists of K mesh points. Let k0 be an integer such that
ϵ
µ(Pk ) < ∀k ≥ k0 .
4K(M − m)

Let P̃k be the the partition of Pk obtained by adjoining all the mesh points of Q. By our earlier lemma
ϵ ϵ
|L(P̃k , f ) − L(Pk , f )| ≤ K(M − m)µ(Pk ) < , |U (P̃k , f ) − U (Pk , f )| ≤ K(M − m)µ(Pk ) < .
4 4
Moreover, since
∫ b
L(Pk , f ) ≤ f (x) dx ≤ U (Pk , f ), , L(Pk , f ) ≤ S(PkTk , f ) ≤ U (Pk , f )
a
we have for k ≥ k0 : ∫
b

f (x) dx − S(Pk , f ) ≤ U (Pk , f ) − L(Pk , f )
Tk
a

≤ U (P̃k , f ) − L(P̃k , f ) + 2K(M − m)µ(Pk ) ≤ U (Q, f ) − L(Q, f ) + ϵ/2 < ϵ.

We also have a converse to this theorem:

4
Theorem. Suppose f : [a, b] → R is bounded and has the property that for any family of marked partitions {Πk }
with µ(Πk ) → 0 as k → ∞ the sequence {S(Πk , f )}∞
k=1 converges. Then f ∈ R[a, b], i.e. f is Riemann integrable on
[a, b].

We omit the proof but the idea is simple. We can construct partitions Pk such that µ(Pk ) → 0 as k → ∞. Then
we can mark these partitions so that of S(Π2k−1 , f ) − L(Pk , f ) < 1/k and U (Pk , f ) − S(Π2k , f ) < 1/k. This implies
that U (Pk , f ) − L(Pk , f ) will converge to zero and therefore, by the Riemann lemma f ∈ R[a, b]

The algebra of integrable functions


Riemann sums are real handy to use to prove various algebraic properties for the Riemann integral. Below we list
the main properties as a theorem and indicate how the proofs go. First a couple of simple lemmas that are rather
easy to prove by just writing out what the claims are.

Lemma A. Let f and g be bounded functions on [a, b] and k a constant. Let Π be a marked partition on [a, b].
Then
1. S(Π, f + g) = S(Π, f ) + S(Π, g).
2. S(Π, kf ) = kS(Π, f ).
3. If f (x) ≤ g(x) ∀x ∈ [a, b] then S(Π, f ) ≤ S(Π, g).
4. |S(Π, f )| ≤ S(Π, |f |).

Lemma B. Let f be bounded function on [a, b] and let c ∈ (a, b). Let P1 be a partition on [a, c] and P2 a partition
on [c, b]. Let P = P1 ∪ P2 . Then P is a partition of [a, b]. Then U (P, f ) = U (P1 , f ) + U (P2 , f ), L(P, f ) =
L(P1 , f ) + L(P2 , f ), and hence U (P, f ) − L(P, f ) ≤ [U (P1 , f ) − L(P, f )] + [U (P2 , f ) − L(p2 , f )]

Theorem (Algebra of R[a, b]). Let f, g ∈ R[a, b] and let k ∈ R. Then


∫b ∫b ∫b
1. f + g ∈ R[a, b] and a (f (x) + g(x)) dx = a f (x) dx + a g(x) dx.
∫b ∫b
2. kf ∈ R[a, b] and a kf (x) dx = k a f (x) dx.
∫b ∫b
3. If f (x) ≤ g(x) ∀x ∈ [a, b] then a f (x) dx ≤ a g(x) dx.
4. If [p, q] ⊂ [a, b] then f ∈ R[p, q].
∫b ∫c ∫b
5. If a < c < b then a f (x) dx = a f (x) dx + c f (x) dx.
6. If f ([a, b]) ⊂ [c, d] and ϕ ∈ C[c, d] then ϕ ◦ f ∈ R[a, b].
∫ ∫
b b
7. |f | ∈ R[a, b] and a f (x) dx ≤ a |f (x)| dx.

8. f g ∈ R[a, b].

Proof. Let Πk := Π(Pk , Tk ), k = 1, 2, · · · be marked partitions such that µ(Πk ) → 0 as k → ∞. Then by Lemma A

lim S(Πk , f + g) = lim S(Πk , f ) + lim S(Πk , g).


k→∞ k→∞ k→∞

∫b ∫b
the right hand side converges to a f (x) dx + a f gx) dx as k → ∞. Therefore f + g is integrable and its integral
is the sum of the integrals of f and g. The proofs of (2) and (3) follow similarly. To prove (4) we use the Riemann
Lemma. Given ϵ > 0 we can find a partition P on [a, b] such that U (P, f ) − L(P, f ) < ϵ. We modify this partition

5
by adjoining p and q. This refines the partition P to a partition Q that can be written as a union Q1 ∪ Q2 ∪ Q3
of partitions on each of the intervals [a, p], p, q] and [q, b]. Let U∗ (Q2 , f ) and L∗ (Q2 , f ) be the sum of only those
terms in the upper and lower Riemann sums that correspond to panels that lie between p and q. Then by Lemma
B, 0 ≤ U∗ (Q2 , f ) − L∗ (Q2 , f ) ≤ U (Q, f ) − L(Q, f ) ≤ U (P, f ) − L(P, f ) < ϵ, and hence, by the Riemann lemma, f is
integrable on [p, q]. To prove (5) we arrange it so that c ∈ Pk for all k. That is c = xm(k) ∈ Pk . Then

S(Πk , f ) = SL (Πk , f ) + SR (Πk , f ),

where SL denotes the sum of terms 1, 2, ·, m(k) and SR denotes the sum of the remaining terms. Clearly
∫ c ∫ b ∫ b
SL (Πk , f ) → f (x) dx, SR (Πk , f ) → f (x) dx, S(Πk , f ) → f (x) dx.
a c a

The proof of (6) is somewhat lengthy and so we will do that last The first part of (7) follows immediately from (6)
with ϕ(s) := |s|. The second part follows from the triangle inequality applied to the Riemann sums Lemma B-4:

|R(Πk , f )| ≤ R(Πk , |f |).

Taking limits we have the inequality in (7). We can also use (6) with ϕ(s) := s2 to deduce that f 2 , g 2 and (f + g)2
are all in R[a, b]. Therefore
1[ ]
fg = (f + g)2 − f 2 − g 2 ∈ R[a, b].
2
To prove (6), let ϵ > 0 be given. Since ϕ is uniformly continuous and bounded, there exists a δ > 0 such that
|ϕ(u) − ϕ(v)| < ϵ/[2(b − a)] whenever u, v ∈ [c, d] and |u − v| < δ, and there exist numbers p and q such that
ϕ([c, d]) ⊂ [p, q]. Since f is integrable we can find a partition P := {x0 , x1 , · · · , xN } such that


N
U (P, f ) − L(P, f ) = (Mi − mi )∆xi < η := ϵδ/[2(q − p)],
i=1

where Mi (resp. mi ) is the supremum (resp. infimum) of f on [xi−1 , xi ]. Let

Jδ := {j ∈ N : 1 ≤ j ≤ N, Mj − mj < δ}, Kδ := {j ∈ N : 1 ≤ j ≤ N, Mj − mj ≥ δ}.

We note that

N ∑
η> (Mi − mi )∆xi ≥ δ∆xi .
i=1 i∈Kδ

This implies that ∑


∆xi < η/δ = ϵ/[2(q − p)].
i∈Kδ

Let Mi∗ (resp. m∗i ) be the supremum (resp. infimum) of ϕ ◦ f on [xi−1 , xi ], then


N ∑ ∑
U (P, ϕ ◦ f ) − L(P, ϕ ◦ f ) = (Mi∗ − m∗i )∆xi = (Mi∗ − m∗i )∆xi + (Mi∗ − m∗i )∆xi <
i=1 i∈Jδ i∈Kδ
∑ ∑
ϵ∆xi /[2(b − a)] + (q − p)∆xi ≤ ϵ/2 + ϵ/2 = ϵ.
i∈Jδ i∈Kδ

Warning. If |f | is Riemann integrable it does not follow that f is Riemann integrable. Counterexample : f (x) = 0
if x rational, f (x) = −1 if x is irrational.

Remark. In general composition of two Riemann integrable functions is not integrable as can be seen from the
following example. Let
f (x) = 0 ∀x > 0 and f (x) = 1 ∀x ≤ 0,

6
ψ(x) = 0 ∀x ∈
/ Q, ψ(p/q) = 1/q ∀p ∈ Z, ∀q ∈ N,
where we assume that p and q have no common factors in N except 1. The function ψ is continuous at all irrationals.
It is a fact (not proven here) that a bounded function that is continuous on [a, b] except at countably many points
is Riemann integrable, and hence ψ is Riemann integrable on any bounded interval. The function f is integrable
on [0, 1], but the function f ◦ ψ is the Dirichlet function (f (x) is zero on all rationals and 1 on all irrationals). The
Dirichlet function is not Riemann integrable on any interval [a, b] with a < b, since all upper sums have the value
b − a while all lower sums are zero. This provides an example of a composition of two Riemann integrable functions
that is not Riemann integrable. There are even examples where f is Riemann integrable, ψ is continuous, but f ◦ ψ
fails to be Riemann integrable.

Definition. Suppose that α < β the we define


∫ α ∫ β
f (x) dx := − f (x) dx.
β α

Theorem. Let f ∈ R[a, b] and p, q, r ∈ [a, b] then


∫ q ∫ r ∫ r
f (x) dx + f (x) dx = f (x) dx
p q p

irrespective of the relative sizes of p, q, and r.

The Fundamental Theorem of Calculus

Theorem. Suppose f : [a, b] → R is a differentiable function and suppose f ′ is Riemann integrable on [a, b]. Then
∫ b
f ′ (x) dx = f (b) − f (a).
a

(n) (n) (n)


Proof. Let Pn , n = 1, 2, · · · be partitions of [a, b] such that limn→∞ µ(Pn ) = 0. That is Pn = {x0 , x1 , x2 , · · ·}.
(n) (n) (n)
Let Tn := {t0 , t1 , t2 , · · ·} be a marking of the partition Pn in such a way that
f ′ (tj ) = [f (xj ) − f (xj−1 )]/[xj
(n) (n) (n) (n) (n)
− xj−1 ].
This is possible by the Mean Value Theorem. Let Πn be the partition Pn together with the marking Tn . Then
∑ ∑
R(Πn , f ′ ) = f ′ (ti )∆xi =
(n) (n) (n)
[f (xi ) − f (xi−1 )] = f (b) − f (a). (5)
i i

The last equality follows from the fact that the sum is a telescoping series. Now, letting n tend to infinity we obtain
the result we wanted.

Notation.
f (x)|ba := f (b) − f (a).
The formula for the Fundamental Theorem of Calculus then reads
∫ b
f ′ (x) dx = f (x)|ba .
a

7
Leibniz’s Rule

Lemma. Let g : [a, b] → R be Riemann integrable on [a, b] and suppose that g is continuous at c ∈ [a, b]. Let
∫ x
G(x) := g(s) ds.
a

Then G is differentiable at c and G′ (c) = g(c).

Proof. It suffices to show that


lim {[G(c + t) − G(c)]/t − g(c)} = 0.
t→0

But the left side can be written as


∫ t ∫ c+t
1 1
lim [g(c + s) − g(c)] ds = lim [g(s) − g(c)] ds.
t→0 t 0 t→0 t c

Given any ϵ > 0 we can find a δ > 0 such that |g(c + s) − g(c)| < ϵ whenever |s| < δ. Hence, if |t| < δ then
∫ t
1
[g(c + s) − g(c)] ds < ϵ.
t
0

A more general result is

Leibniz’s Rule Let g : [a, b] → R be Riemann integrable on [a, b] and suppose that g is continuous on [a, b]. Let
α : [c, d] → [a, b], and β : [c, d] → [a, b] be differentiable functions. Let
∫ β(x)
Φ(x) := g(s) ds.
α(x)

Then Φ is differentiable at x0 and

Φ′ (x0 ) = g(β(x0 ))β ′ (x0 ) − g(α(x0 ))α′ (x0 ).

Proof 1
Choose x∗ ̸= x0 and define ∫ s
G(s) := g(t) dt,
x∗

then we can write


Φ(x) = G(β(x)) − G(α(x)).
The result then follows from the chain rule together with the above lemma.

1 Actually Leibniz’s rule is a bit more general. Under the right hypotheses it is true that
∫ β(x) ∫ β(x)
d ∂g(x, s)
g(x, s) ds = ds + g(x, β(x))β ′ (x) − g(x, α(x))α′ (x).
dx α(x) α(x) ∂x

8
A few more theorems on integrals.

Mean Value Theorem for Integrals Let f ∈ C[a, b] and g ∈ R[a, b] with g(x) ≥ 0 ∀x ∈ [a, b]. Then there exists
a c ∈ [a, b] such that
∫ b ∫ b
f (x)g(x) dx = f (c) g(x) dx. (6)
a a

∫b
Proof. Let m := min(f ), and M := max(f ) and G := a g(x) dx. Then since g(x) ≥ 0 we have mg(x) ≤ f (x)g(x) ≤
M g(x) so that
∫ b ∫ b ∫ b
mG = m g(x) dx ≤ f (x)g(x) dx ≤ M g(x) dx = M G. (7)
a a a

If G = 0 then all integrals in (3) are zero and obviously (2) will be true for any choice of c. If G > 0 then m ≤
∫b ∫b
G−1 a f (x)g(x) dx ≤ M so that by the intermediate value theorem there exists a c such that f (c) = G−1 a f (x)g(x)
for some c ∈ [a, b].

Change of Variables Theorem. Let J be an interval, ϕ : [a, b] → J a differentiable function with ϕ′ ∈ R[a, b]. Let
f : J → R be continuous. Then
∫ b ∫ ϕ(b)
f (ϕ(t)) ϕ′ (t) dt = f (x) dx.
a ϕ(a)

Proof. Let ∫ ∫
x ϕ(x)
F (x) := f (ϕ(t)) ϕ′ (t) dt − f (s) ds.
a ϕ(a)

Clearly F (a) = 0 and using Leibniz’s rule we see that F ′ (x) = 0. This means F (x) = 0 ∀x ∈ [a, b]. In particular
F (b) = 0, which is precisely what the theorem asserts.

Integration by parts. Let f : [a, b] → R and g : [a, b] → R be differentiable functions with f, g ∈ R[a, b]. Then
∫ b ∫ b
f (x)g ′ (x) dx = − f ′ (x)g(x) dx + [f (b)g(b) − f (a)g(a)].
a a

The proof is an immediate consequence of applying the Fundamental Theorem of Calculus to the function f (x)g(x).

Taylor’s Theorem. Let J be an interval, a ∈ J and f ∈ C n+1 (J), i.e. f is n + 1 times continuously differentiable
on J. Then for all x ∈ J we have

n
f (k) (a)
f (x) = f (a) + (x − a)k + Rn+1 (x), (8)
k!
k=1

where ∫ x
1
Rn+1 (x) = f (n+1) (t)(x − t)n dt.
n! a
Moreover, for each x ∈ J there is a number c between x and a such that

f (n+1) (c)
Rn+1 (x) = (x − a)n+1 .
(n + 1)!

9
Note. For n = 0 equation (2) reduces to the Fundamental Theorem of Calculus:
∫ x
f (x) = f (a) + f ′ (t) dt.
a

The proof of the theorem for arbitrary n is effected by repeated integration by parts on the integral or, more simply,
by mathematical induction.

The logarithm, exponential and power functions.


Definition We define the natural logarithm function as
∫ x
1
ln(x) := dt.
1 t
Using this definition we can can derive all the properties of the natural logarithm

Theorem Let a > 0, b > 0.


1. ln : (0, ∞) → (−∞, ∞) is a bijection. It is strictly increasing: if a > b then ln(a) > ln(b).
2.
ln(ab) = ln(a) + ln(b).

3. If m ∈ Z and n ∈ N then
m
ln(am/n ) = ln(a).
n
4.
ln(1) = 0, ln(a/b) = ln(a) − ln(b).

5.
[ln(x)]′ = 1/x.

We can define the exponential function as the inverse function for the natural logarithm function:
−1
exp := [ln] : (−∞, ∞) → (0, ∞).

Now it is not difficult to prove all the standard properties of the exponential function.

Theorem Let y, z ∈ R, then


1.
exp : (−∞, ∞) → (0, ∞)
is a bijection. It is strictly increasing.
2.
exp(y + z) = exp(y) exp(z).

3. If m ∈ Z and n ∈ N then (m )
exp y = [exp(y)]m/n .
n

10
4.
exp(0) = 1.

5.
[exp(x)]′ = exp(x).

Definition We define e to be the number exp(1).

This definition implies ln(e) = 1.

Let a be a positive real number and let p ∈ R. Temporarily let us use the notation

E(a, p) := exp(p ln(a)).

If p = m/n where m ∈ Z and n ∈ N then


m m/n
E(a, p) = exp( ln(a)) = [exp(ln(a))] = am/n = ap .
n
This suggests that we define ap as E(a, p):

Definition. Let a > 0, and p ∈ R then we define

ap := exp(p ln(a)).

∫q
Exercise: Let a > 0. Find the derivative of f (x) := ax and p
ax dx.

Exercise: Let a > 1 and let λ(x) := ax /[K + ax ] (a logistic function) . Show that λ is a bijection between R and
(0, 1). Find its inverse function and find an antiderivative Λ, i.e. a function such that Λ′ = λ.

11

You might also like