Calculus II: For Biology and Medicine
Calculus II: For Biology and Medicine
Calculus II: For Biology and Medicine
Yi Li
Department of Mathematics, Johns Hopkins University, 3400 North
Charles Street, Baltimore, MD 21218
E-mail address: yli@math.jhu.edu; yilicms@gmail.com
Abstract. This is the lecture note on Calculus for biology and medicine. The
content contains additional topics beyond the textbook and lots of interesting
problems.
• Please inform me typos and other error (e.g., grammar) via
yilicms@gmail.com!
• I will keep to update material.
• All figures come from Google Image and Wiki.
Contents
3
CHAPTER 1
1http : //546834391620409041.weebly.com/
The above three types of intervals are clearly bounded; unbounded intervals are
sets of the forms {x ∈ R : x > a}. Here are the possible cases:
[a, ∞) := {x ∈ R : x ≥ a},
(−∞, a] := {x ∈ R : x ≤ a},
(a, ∞) := {x ∈ R : x > a},
(−∞, a) := {x ∈ R : x < a}.
The symbols “∞” and “−∞” “plus infinity” and “minus infinity”, respectively. In
particular,
R = (−∞, ∞).
Looking at Figure 1.3, we see the x = 0 is the “singular point” of the blue
line; a geometric interpretation says that the function y = x has no tangent line at
x = 0. The second picture in Figure 1.3 locally looks like the first one, at points
x = b, c, d.
1.1.2. Lines in the plane. The linear equation in the plane is given by
(1.1.2) Ax + By + C = 0,
1.1. REAL AND COMPLEX NUMBERS 7
1.1.3. Equation of the circle. A circle is the set of all points at a given
distance, called the radius, from a given point, called the center.
If r is the distance from (x0 , y0 ) to (x, y), then we have
(1.1.6) (x − x0 )2 + (y − y0 )2 = r2 .
If r = 1 and (x0 , y0 ) the circle is called the unit circle, denoted S1 (that is, S1 is
the sphere of dimension 1).
Furthermore,
(i) Symmetry:
(ii) Shifts:
π π
sin θ + = cos θ, cos θ + = − sin θ,
2 2
sin(θ + π) = − sin θ, cos(θ + π) = − cos θ,
sin(θ + 2π) = sin θ, cos(θ + 2π) = cos θ.
sin(α ± β) = sin α · cos β ± cos α · sin β, cos(α ± β) = cos α · cos β ± sin α · sin β.
ar as = ar+s , (ab)r = ar br ,
ar a r ar
= ar−s , = ,
as b br
1
a−r = , (ar )s = ars .
ar
10 1. NUMBERS AND FUNCTIONS
√
A complex number z = a + b −1 corresponds to a point (a, b) in the plane,
conversely,
√ a point (a, b) in the plane corresponds to a complex number z = a +
b −1. Consequently,
√ √
a + b −1 = c + d −1 ⇐⇒ a = c and b = d.
2This number is an irrational number and can be defined as (see later) lim 1 n
n→∞ (1 + n ) .
1.2. ELEMENTARY FUNCTIONS 11
√ √
Given two complex numbers a + b −1 and c + d −1, we have
√ √ √
(a + b −1) + (c + d −1) = (a + c) + (b + d) −1,
√ √ √ √ √
(a + b −1)(c + d −1) = ac + ad −1 + bc −1 + bd( −1)2
√
= (ac − bd) + (ad + bc) −1.
√
Definition 1.1.3. If z = a + b −1 is a complex number, its conjugate, denoted
by z̄, is defined as
√
(1.1.11) z̄ := a − b −1.
√
If z = a + b −1, then
p
|z| = r, r := a2 + b2 .
Thus z lies in the circle at the origin with radius r.
Remark 1.2.7. (Fourier series of periodic functions) suppose that f (x) is a periodic
function of period 2π. Then we may assume that f (x) is defined on [−π, π]. If f (x)
is integrable (for definition, see Chapter 6), we define
1 π 1 π
Z Z
(1.2.1) an := f (x) cos(nx) dx, bn := f (x) sin(nx) dx, n ∈ N≥0 .
π −π π −π
which are called the Fourier coefficients of f . The formal infinite sum
a0 X
(1.2.2) (Sf )(x) := + an cos(nx) + bn sin(nx)
2
n∈N
14 1. NUMBERS AND FUNCTIONS
0. Functions that are not algebraic are called transcendental. All the trigonometric,
exponential, and logarithmic functions are transcendental functions.
Remark 1.2.11. 3The name “transcendental” comes from Leibniz in his 1682
paper where he proved sin x is not an algebraic function of x, and Euler was probably
the first person to define transcendental numbers in the modern sense.
(1) Joseph Liouville first proved the existence of transcendental numbers in
1844, and in 1851 gave the first decimal examples such as the Liouville
constant X 1
.
10k!
k∈N
(2) Johann Heinrich Lambert conjectured that e and π were both transcen-
dental numbers in his 1761 paper proving the number π is irrational.
(3) Charles Hermit in 1873 proved that e is transcendental.
(4) In 1874, Georg Cantor proved that the algebraic numbers are countable
and the real numbers are uncountable. In 1878, Cantor showed that there
are as many transcendental numbers as there are real numbers.
(5) In 1882, Ferdinand von Lindemann published a proof that the number π
is transcendental. He first showed
√
that e to any nonzero
√ algebraic power
is transcendental, and since e −1π = −1 is algebraic, −1π and therefore
π must be transcendental.
(6) In 1900, David Hilbert posed the Hilbert’s seventh problem: if a is an
algebraic number which is not zero or one, and b is an irrational algebraic
number, is ab necessarily transcendental? The affirmative answer was
provided in 1934 by the Helfond-Schneider theorem.
2.1. Sequences
A sequence if the function
f : N −→ R, n 7−→ f (n)
as a list of numbers a0 , a1 , a2 , · · · where an = f (n). We will write {an : n ∈ N≥0 }
or {an }n∈N≥0 if we mean the entire sequence.
2.1.1. kth order recursion. From Example 2.1.1 (2), we see that
√
an+1 = (n + 1)2 = n2 + 2n + 1 = an + 2 an + 1;
thus an+1 depends only on an . In general, we have the following
Definition 2.1.3. The sequence {an }n∈N or {an }n∈N≥0 has limit a < ∞, written
as limn∈∞ an = a, if, for every > 0, there exists an integer N (depending on )
such that
|an − a| <
whenever n > N . If the limit exists, the sequence is called convergent and we say
that an converges to a as n tends to infinity. If the sequence has no limit, it is
called divergent.
Proposition 2.1.4. If the sequence {an }n∈N or {an }n∈N≥0 is convergent, then the
limit of this sequence is unique.
Proof. Suppose we have two limits a and a0 . By the definition, for every
> 0, there exist integers N and N 0 such that
Since the left-hand side does not depend on n and , letting → 0, we obtain
a = a0 .
1
Exercise 2.1.5. Show that limn→∞ n = 0.
Exercise 2.1.6. Show that the sequence {an }n∈N , where an = n, is divergent.
For example,
n+1 1 1
lim = lim 1 + = 1 + lim =1+0=1
n→∞ n n→∞ n n→∞ n
Remark 2.1.8. The assumption in Theorem 2.1.7 that limn→∞ an and limn→∞ bn
exists is necessary. For example, consider an = n and bn = −n; then (2.1.1) does
not hold.
In Example 2.1.9, we can give another short proof. Suppose the limit limn→∞ an =
a exists, by taking the limit of n, we have
a = lim an+1 = 4 − 2 lim an = 4 − 2a
n→∞ n→∞
which gives us a = 4/3. Thus approach shows that if we know the limit exists,
then we can take the limits on both-sides of an+1 = g(an ) to get the desired limit.
However, in order to applying this argument, we should know the existence at first.
This can be seen from the following fundamental theorem.
Theorem 2.1.10. Any bounded increasing or decreasing sequence has finite limit.
We say that a sequence {an }n∈N≥0 is bounded if |an | ≤ M for all n, where M
is some positive constant. A sequence {an }n∈N≥0 is increasing (or decreasing)
if an ≤ an+1 (or an ≥ an+1 ).
√
Example 2.1.11. Consider an+1 = 3an with a0 = 2. If limn→∞ an = a exists,
then √
a = 3a =⇒ a = 0 or a = 3.
We now prove that the limit of an exists. Since
√ √ √
a1 = 6 < 3, a2 = 3a1 < 3 × 3 = 3,
we claim that an ∈ (0, 3). If this claim holds for n, then
√ √
an+1 = 3an < 3 × 3 = 3, an+1 > 0,
by inductive hypothesis. Hence an ∈ (0, 3) for all n. From
r
an+1 3
= > 1,
an an
it follows that the sequence {an }n∈N≥0 is bounded and increasing, and therefore
the limit limn→∞ an exists.
Observe that if a is a fixed point of the sequence an+1 = g(an ), then a is a fixed
point of the function g : R → R. Conversely, if a is a fixed point of the function
f : R → R, then we can construct a sequence {an }n∈N≥0 of which a is a fixed
point.
√
For example f (x) = x2 + 5/2, x ∈ R, is a Lipschitz continuous function with
L ≤ 1/2.
22 2. SEQUENCES AND DIFFERENCE EQUATIONS
The equation comes from the following problem posed in 1202 of his Liber Abaci
(Book of Calculation) by Leonardo Pisano Bigollo (also known as Leonardo of Pisa
or most commonly Fibonacci):
How many pairs of rabbits are produced if each pair reproduces
one pair of rabbits at age one month and another pair of rabbits
at age two months and initially there is one pair of newborn
rabbits?
The Fibonacci sequence (2.2.9) has a closed-form solution
ϕt − ψ t
(2.2.10) Nt = ,
ϕ−ψ
2.2. POPULATION MODELS 25
where
√
1+ 5
(2.2.11) ϕ := ≈ 1.6180339887 · · ·
2
is the golden ratio or golden mean, and
√
1− 5 1
(2.2.12) ψ := = 1 − ϕ = − ≈ −0.6180339887 · · · .
2 ϕ
3.1. Limits
The statement
where c, L are finite, means that, for every > 0, there exists a number δ > 0 such
that
|f (x) − L| <
whenever 0 < |x − c| < δ. In this case, we say that f (x) converges to L. If this
limit does not exist, we say that f (x) diverges as x tends to c.
Remark 3.1.1. In the above definition, we exclude the value x = c from the
statement. (This is done in the inequality 0 < |x − c|.) A reason for that is the
function f (x) may not be defined at x = c. For example,
x, x 6= 0,
f (x) =
1, x = 0.
In this case, we have limx→0 f (x) = 0 6= f (0).
(2) limx→c− f (x) = L means that for every > 0 there exists a number δ > 0
such that
Observe that
27
28 3. LIMITS AND CONTINUITY
The statement
(3.1.2) lim f (x) = ∞ (or − ∞)
x→c
means that, for every M > 0, there exists a δ > 0 such that
f (x) > M (or < −M )
whenever 0 < |x − c| < δ. Similarly, we can define
lim f (x) = ∞, lim f (x) = −∞, lim f (x) = ∞, lim f (x) = −∞.
x→c+ x→c+ x→c− x→c−
1 1
Example 3.1.3. limx→0+ x = ∞ and limx→0− x = −∞.
The statement
(3.1.3) lim f (x) = L
x→∞
means that, for every > 0, there exists an x0 > 0 such that
|f (x) − L| <
whenever x > x0 . Similarly, we can define limx→−∞ f (x) = L.
3.1. LIMITS 29
Theorem 3.1.4. If limx→c f (x) (c can be infinity) and limx→c g(x) exist and a is
a constant, then
(3.1.4) lim [af (x)] = a lim f (x),
x→c x→c
(3.1.5) lim [f (x) + g(x)] = lim f (x) + lim g(x),
x→c x→c x→c
(3.1.6) lim [f (x)g(x)] = lim f (x) · lim g(x),
x→c x→c x→c
f (x) limx→c f (x)
(3.1.7) lim = , provided lim g(x) 6= 0.
x→c g(x) limx→c g(x) x→c
Here p(x) and q(x) are two polynomials of degrees deg(p) and deg(q), respectively,
and L is the ratio of the coefficients of the leading terms in the numerator and
denominator.
Theorem 3.1.7. If f (x) ≤ g(x) ≤ h(x) for all x in an open interval that contains
c (except possibly at c) and
lim f (x) = lim h(x) = L,
x→c x→c
then
lim g(x) = L.
x→c
sin x sin x
Figure 3.3. x and the inequality cos x ≤ x ≤1
Remark 3.1.8. If f (x) ≤ g(x) ≤ h(x) for all x > x0 , where x0 is a positive fixed
number, and
lim f (x) = lim h(x) = L,
x→∞ x→∞
then
lim g(x) = L.
x→∞
Proposition 3.1.10. If f (x) is bounded (i.e., |f (x)| ≤ M for some positive con-
stant M ) and limx→c g(x) = 0, then
lim f (x)g(x) = 0.
x→c
3.2. Continuity
A function f is said to be continuous at x = c if
(3.2.1) lim f (x) = f (c).
x→c
To check whether a function is continuous at x = c, we need to check the following
three conditions:
(1) f (x) is defined at x = c,
(2) limx→c f (x) exists,
(3) limx→c f (x) = f (c).
If any of these three conditions fails, the function is discontinuous at x = c.
Theorem 3.2.3. Suppose that a is a constant and the functions f and g are con-
tinuous at x = c. Then the following functions are continuous at x = c:
(1) a · f ,
(2) f + g,
(3) f · g,
(4) f /g provided that g(c) 6= 0.
Proof. Note that (f ◦ g)(c) = f [g(c)] exists. Given an > 0. There exists a
positive number δ such that
|f (y) − f (L)| <
whenever |y − L| < δ. For such δ, we can find another positive number δ 0 such that
Theorem 3.2.7. Suppose that f is continuous on the closed interval [a, b]. If M
is any real number with f (a) < M < f (b) or f (b) < M < f (a), then there exists at
least one number c on the open interval (a, b) such that f (c) = M .
34 3. LIMITS AND CONTINUITY
Remark 3.2.8. Consider the function f (x) = x. Then for each M ∈ (a, b), we
have a unique M ∈ (a, b) with f (M ) = M . Hence we can not weaken the open
interval (a, b) in Theorem 3.2.7 to the closed interval [a, b].
Remark 3.2.9. Suppose that f is continuous function defined on [a, b]. If f (a) <
0 < f (b), then Theorem 3.2.7 shows that f (x) = 0 has at least one solution in (a, b).
CHAPTER 4
Differentiation
4.1. Derivatives
The derivative of a function f at x, denoted f 0 (x), is
f (x + h) − f (x)
(4.1.1) f 0 (x) := lim
h→0 h
provided that the limit exists. If the limit exists, then we say that f is differen-
tiable at x. (see Figure 4.1) The quotient
f (x + h) − f (x)
h
is called the difference quotient, and we denote it by ∆f ∆x . We use the Leibniz
notation
dy df d
(4.1.2) y0 = = f 0 (x) = = f (x).
dx dx dx
For particular value c, w can write
df
(4.1.3) = f 0 (c).
dx x=c
Proof. From
f (x) − f (c)
f (x) − f (c) = · (x − c),
x−c
we have
f (x) − f (c)
lim [f (x) − f (c)] = lim · (x − c) = f 0 (c) · 0 = 0
x→c x→c x−c
by Proposition 3.1.10.
Theorem 4.2.1. Suppose a is a constant and f (x) and g(x) are differentiable at
x. Then
d d
(4.2.2) [af (x)] = a f (x),
dx dx
d d d
(4.2.3) [f (x) + g(x)] = f (x) + g(x).
dx dx dx
Theorem 4.2.3. If f (x) and g(x) are differentiable at x, and g(x) 6= 0, then
f 0 (x)g(x) − f (x)g 0 (x)
d f (x)
(4.2.5) = .
dx g(x) [g(x)]2
d −n
(4.2.6) (x ) = −nx−n−1 .
dx
In general,
d r
(4.2.7) (x ) = rxr−1 .
dx
4.2.3. Chain rule. To find the derivative of composite functions, we need the
chain rule.
For example,
d d 1
(4.2.17) arcsin x = sin−1 x = √ ,
dx dx 1 − x2
d d 1
(4.2.18) arctan x = tan−1 x = .
dx dx 1 + x2
4.2.4. Linear approximation and error propagation. Assume that y =
f (x) is differentiable at x = a; then
(4.2.19) L(x) := f (a) + f 0 (a)(x − a)
is the tangent line approximation or the linearization of f at x = a. If |x − a|
is sufficiently small, we have
(4.2.20) f (x) ≈ L(x) = f (a) + f 0 (a)(x − a).
CHAPTER 5
Applications of differentiation
The proof of Theorem 5.1.1 is beyond the scope of this course and will be
omitted.
Remark 5.1.2. (1) Theorem 5.1.1 only tells the existence of global (or absolute)
extrema (global maximum or global minimum). We can find functions that have
more than one global extrema; for example,
1, 0 ≤ x ≤ 1,
f (x) =
2 − x, 1 ≤ x ≤ 2.
(2) If f is continuous on an open interval (a, b), then f may not have a global
extrema. For example, f (x) = 1/x with x ∈ (0, 1).
5.1.2. Local extrema. A function f defined on a set D has a local (or rel-
ative) maximum at a point c if there exists a δ > 0 such that
f (c) ≥ f (x)
for all x ∈ (c − δ, c + δ) ∩ D. A function f defined on a set D has a local (or
relative) minimum at a point c if there exists a δ > 0 such that
f (c) ≤ f (x)
for all x ∈ (c − δ, c + δ) ∩ D. Local maxima and local minima are collectively called
local (or relative) extrema.
Proof. We only prove the case that f has a local minimum at c. Then there
exists a δ > 0 such that
f (c) ≤ f (x)
for all x ∈ (c − δ, c + δ) ∩ D. For x ∈ (c, c + δ) ∩ D, we have
f (c) − f (x) f (x) − f (c)
≤0 or ≥ 0;
x−c x−c
thus
f (x) − f (c)
lim ≥ 0.
x→c+ x−c
Similarly, we can show that
f (x) − f (c)
lim ≤ 0.
x→c− x−c
5.1. EXTREMA AND THE MEAN-VALUE THEOREM 41
However, the converse of Theorem 5.1.3 is not true. For example, consider thee
function f (x) = x, x ∈ [−1, 1]; we see that f 0 (0) = 0 but f has no local extremums
in (−1, 1).
Proof. Define
f (b) − f (a)
F (x) := f (x) − (x − a), x ∈ [a, b].
b−a
Then F is continuous on [a, b] and differentiable on (a, b); furthermore F (a) =
f (a) = F (b). By Theorem 5.1.4, F 0 (c) = 0 for some c ∈ (a, b). Since
f (b) − f (a)
F 0 (x) = f 0 (x) − ,
b−a
f (b)−f (a)
it follows that 0 = F 0 (c) = f 0 (c) − b−a .
Proof. We only prove part (a). Fixed x1 < x2 in (a, b). By Theorem 5.1.5,
we have
f (x2 ) − f (x1 )
= f 0 (c)
x2 − x1
for some c ∈ (x1 , x2 ). Since f 0 (x) > 0 for all x ∈ (a, b), we get f (x2 ) > f (x1 ).
5.2.2. Concavity. A differentiable function f (x) is concave up on an inter-
val I if the first derivative f 0 (x) is an increasing function on I. f (x) is concave
down on an interval I if the first derivative f 0 (x) os a decreasing function on I.
5.2.4. Inflection points. Inflection points are points where the concav-
ity of a function changes–that is, where the function changes from concave up to
concave down or from concave down to concave up.
Theorem 5.3.1. (L’Hôospital’s rule) Suppose that f and g are differentiable func-
tions and that
lim f (x) = lim g(x) = 0
x→a x→a
or
lim f (x) = lim g(x) = ∞.
x→a x→a
If
f 0 (x)
lim =L
x→a g 0 (x)
then
f (x)
lim = L.
x→a g(x)
46 5. APPLICATIONS OF DIFFERENTIATION
Then f (x) and g(x) are continuous at a. Fixed x > a. From Theorem 5.1.5, there
exist x1 , x2 ∈ (a, x) such that
f (x) − f (a) g(x) − g(a)
= f 0 (x1 ), = g 0 (x2 ).
x−a x−a
Observe that x1 , x2 → a+ as x → a+. Hence
f (x) f 0 (x1 )(x − a) f 0 (x1 )
lim = lim 0 = lim 0 = L.
x→a+ g(x) x→a+ g (x2 )(x − a) x→a+ g (x2 )
For (5),
lim xx = lim ex ln x = lim e0 = 1
x→0+ x→0+ x→0+
by (2).
5.4. ANTIDERIVATIVE 47
5.4. Antiderivative
A function F is called an antiderivative of f on an interval I if F 0 (x) = f (x)
for all x ∈ I. (see Figure 5.7)
Corollary 5.4.1. If F (x) and G(x) are antiderivatives of the continuous function
f (x) on an interval I, then there exists a constant C such that
G(x) = F (x) + C
for all x ∈ I.
Integrations
The value of the sum depends on the choice of the partition P and the choice of
the points ck ∈ [xk−1 , xk ] and is called a Riemann sum for f on [a, b].
The phase “if the limit exists” means, in particular, that the value of lim||P||→0 SP
does nor depend on how we choose the partitions and the points ck ∈ [xk−1 , xk ] as
we rake the limit.
Theorem 6.1.2. All continuous functions are Riemann integrals; that is, if f (x)
is continuous on [a, b] then
Z b
f (x) dx
a
exists.
Remark 6.1.3. (1) We can find a function that is Riemann integral but not con-
tinuous. For example,
|x|, x ∈ [−1, 0) ∪ (0, 1],
f (x) =
1, x = 0.
(2) Consider the function
x, x ∈ Q ∩ [0, 1],
f (x) =
0, x ∈ (R/Q) ∩ [0, 1].
It can be showed that the function f (x) is not Riemann integrable.
(3) Another important concept is Lebesgue integrals, which can be think of
the Riemann integral in terms of the partition of y-axis. A Lebesgue integrable
function is Riemann integrable, but the converse is not true. For example, the
Lebesgue integral of f (x) defined in (2) equals 0.
Proposition 6.1.5. Assume that f and g are integrable over [a, b].
Z b Z b
(6.1.4) kf (x) dx = k f (x) dx, k is constant,
a a
Z b Z b Z b
(6.1.5) [f (x) + g(x)] dx = f (x) dx + g(x) dx,
a a a
Z b Z c Z b
(6.1.6) f (x) dx = f (x) dx + f (x) dx,
a a c
where c is any interior point of [a, b]. Moreover
(i) If f (x) ≥ 0 on [a, b], then
Z b
(6.1.7) f (x) dx ≥ 0.
a
(ii) If f (x) ≤ g(x) on [a, b], then
Z b Z b
(6.1.8) f (x) dx ≤ g(x) dx.
a a
(iii) If m ≤ f (x) ≤ M on [a, b], then
Z b
(6.1.9) m(b − a) ≤ f (x) dx ≤ M (b − a).
a
Proposition 6.2.2. If g(x) and h(x) are differentiable functions and f (u) is con-
tinuous for u between g(x) and h(x), then
Z h(x)
d
(6.2.3) f (u) du = f [h(x)]h0 (x) − f [g(x)]g 0 (x).
dx g(x)
0 0
Then F (x) = f (x) = G (x) by Theorem 6.2.1. According to Corollary 5.4, we have
F (x) = G(x) + C
The above discussion tells us that any antiderivative can be obtained from a
special one by adding a constant. We denote by
Z
f (x) dx
Then G0 (x) = f (x) and G(x) is an antiderivative of f (x). If F (x) is any antideriv-
ative of f (x), then
G(x) = F (x) + C, x ∈ [a, b],
by Theorem 6.2.1. In particular,
Z b
f (u) du = G(b) = F (b) + C = F (b) + [G(a) − F (a)] = F (b) − F (a).
a
represents the area of the region bounded by the graph of f (x) between a and b,
the vertical lines x = a and x = b, and the x-axis between a and b. (see Figure 6.3)
If f and g are continuous on [a, b], with f (x) ≥ g(x) for all x ∈ [a, b], then the
area of the region between the curves y = f (x) and y = g(x) from a to b is equal to
Z b
(6.3.2) Area = [f (x) − g(x)] dx.
a
Suppose that a region is bounded by x = f (y) and x = g(y), with g(y) ≤ f (y)
for x ≤ y ≤ d; that is, f (y) is to the right of g(y) for all y ∈ [c, d]. Then the area
of the shaded region is given by
Z d
(6.3.3) Area = [f (y) − g(y)] dy.
c
6.3.2. Average values. If f (x) is a continuous function on [a, b]. The aver-
age value of f on [a, b] is
Z b
1
(6.3.4) favg := f (x) dx.
b−a a
By Theorem 5.1.5,
f (xk ) − f (xk−1 ) ∆yk
f 0 (ck ) = =
xk − xk−1 ∆xk
for some number ck ∈ (xk−1 , xk ). Consequently,
n p
X
LP = 1 + [f 0 (ck )]2 ∆xk
i=1
If f (x) is differentiable on (a, b) and f 0 (x) is continuous on [a, b], then the length
of the curve y = f (x) from a to b is given by
Z bp
(6.3.6) L= 1 + [f 0 (x)]2 dx.
a
6.3. APPLICATIONS OF INTEGRATION 57
Since
dy
= f 0 (x),
dx
it follows from (6.3.6) that
s 2
Z b Z bp
dy
(6.3.7) L= 1+ dx = (dx)2 + (dy)2 .
a dx a
p
We call the expression (dx)2 + (dy)2 the arc length differential and denote it
by ds.
Theorem 6.3.4. (Hölder’s inequality) Suppose that p and q are two real numbers,
1 ≤ p, q ≤ ∞ with p1 + 1q = 1, and f and g are two Riemann integrable functions
on [a, b]. Then the functions |f (x)|p and |g(x)|q are also Riemann integrable and
Z !1/p Z !1/q
b Z b b
p q
(6.3.11) f (x)g(x) dx ≤ |f (x)| dx |g(x)| dx .
a a a
Proof. We may assume that b ≤ ap−1 . Consider the function f (x) := xp−1
1
with x ∈ [0, a]. Then the inverse function g(x) of f (x) is equal to g(x) = x p−1 . By
Figure 6.6, we see that the area of the red region is given by
Z a Z a
ap
f (x) dx = xp−1 dx = ,
0 0 p
while the area of the yellow region is given by
Z b Z b Z b
1 p − 1 p−1
p bq
f −1 (y) dy = g(x) dx = x p−1 dx = b = .
0 0 0 p q
Since we assume that b ≤ ap−1 , from Figure 6.6 we have
ap bq
ab ≤ area of the red region + area of the yellow region = + .
p q
Similarly, we can deal with the case that b ≥ ap−1 .
Integration techniques
are the same. The above rule is the original stage of the following
several methods.
Hence Z Z Z
3
u du = 2x dx = y dy.
Proof. The proof is essentially similar as that of Proposition 7.1.1 Let F (x)
be an anti-derivative of f (x). Then F [g(x)] is an anti-derivative of f [g(x)]g 0 (x) and
hence, by the fundamental theorem of calculus, we have
Z b Z g(b)
f [g(x)]g 0 (x) dx = F [g(b)] − F [g(a)] = f (u) du,
a g(a)
d ax
Proof. The basic idea to deal with Ik and Ik (a) is an observation that dx e =
aeax . In other words,
deax = aeax dx.
Since Ik = Ik (1), we suffice to consider the second integral Ik (a). Letting u = xk /a
and v = eax in (7.2.3), we have, for k ≥ 1,
Z Z Z
1 1 k ax
Ik (a) = xk eax dx = xk aeax dx = x de
a a
Z Z
1 k ax 1 1 k ax 1
= x e − eax dxk = x e − kxk−1 eax dx
a a a a
1 k ax k
= x e − Ik−1 (a).
a a
Note that Z
1
I0 (a) = eax dx = eax + C.
a
for example,
1 ax 1 x ax 1
I1 (a) = xe − I0 (a) = e − 2 eax + C,
a a a a
x2
1 2 ax 2 2x 2
I2 (a) = x e − I1 (a) = − 2 + 3 eax + C.
a a a a a
It is a good excise for you to find the general formula for Ik (a).
Since Z Z
I1 = ln x dx = x ln x − dx = x ln x − x + C,
we can calculate all Ik ’s by the above recursion formula and the initial value.
7.2.2. Gamma function and Beta function. The gamma function Γ(x)
is defined by
Z ∞
(7.2.4) Γ(x) := tx−1 e−t dt, x > 0.
0
This is an improper integral which is well-defined (see later). By (7.2.3), we have
1 ∞ −t x tx
Z Z Z
x−1 −t 1
t e dt = e dt = t + tx e−t dt.
x 0 xe x
Since
tx tx
lim t
= 0 = lim t ,
t→∞ xe t→0 xe
it follows that
(7.2.5) Γ(1 + x) = xΓ(x).
Using the recursion formula (7.2.5), we have
(7.2.6) Γ(n + 1) = n!, n ∈ N.
There are some fundamental properties of Γ(x):
(1) Euler’s reflection formula: for any 0 < x < 1, we have
π
(7.2.7) Γ(x)Γ(1 − x) = .
sin(πx)
In particular,
√
r
1 π
(7.2.8) Γ = = π.
2 sin(π/2)
(2) Duplication formula: for any x > 0, we have
√
1
(7.2.9) Γ(x)Γ x + = 21−2x πΓ(2x).
2
Another important and useful integral is the Beta function defined by
Z 1
(7.2.10) B(x, y) := tx−1 (1 − t)y−1 dt, x, y > 0.
0
This is also an improper integral. The beta function was studied by Euler and
Legendre and was given its name by Jscques Binet.
(3) B(x, y) = B(y, x).
(4) For any x, y > 0 we have
Γ(x)Γ(y)
(7.2.11) B(x, y) = .
Γ(x + y)
7.3. RATIONAL FUNCTIONS AND PARTIAL FRACTIONS 65
If the degree of P (x) in (7.3.1) is greater than or equal to the degree of Q(x),
then the first step in the partial-fraction decomposition is to use long division to
write f (x) as a sum of a polynomial and a rational function, where the rational
function is such that the degree of the polynomial in the numerator is less than the
degree of the polynomial in the denominator. (Such rational functions are called
proper.)
7.3.1. Partial-fraction decomposition I. We discuss the cases of distinct
and repeated linear factors.
For example,
Z
−1
Z
1 1
dx = + dx = ln |x − 1| − ln |x| + C,
x(x − 1) x x−1
and
Z
−1
Z
x 1 1
dx = + dx = ln |x + 1| + + C.
(x + 1)2 x + 1 (x + 1)2 x+1
7.3.2. Irreducible quadratic factors. We discuss the cases of distinct and
repeated irreducible quadratic factors.
For example,
2x3 − x2 + 2x − 2
Z
−1
Z
2x
dx = + dx = ln(x2 + 2) − tan−1 x + C,
(x2 + 2)(x2 + 1) x2 + 2 x2 + 1
and
x2 + x + 1
Z Z
1 x 1
dx = + dx = tan−1 x − + C.
(x2 + 1)2 x2 + 1 (x2 + 1)2 2(x2 + 1)
Proof. Since
x4 (1 − x)4 4
2
= x6 − 4x5 + 5x4 − 4x2 + 4 − ,
1+x 1 + x2
it follows that
1
x4 (1 − x)4
Z
22
dx = − π.
0 1 + x2 7
2
Since 1 ≤ 1 + x ≤ 2 for any x ∈ [0, 1], we have
Z 1 Z 1 4 Z 1
1 4 x (1 − x)4
x (1 − x)4 dx ≤ 2
dx ≤ x4 (1 − x)4 dx.
0 2 0 1 + x 0
From the definition above, we know that for each z, the integrals
Z z Z a
f (x) dx, f (x) dx
a z
are finite, however, after taking the limits, the improper integrals may be infinity.
From Example 7.4.3 and Definition 7.4.4, we see that the integral
Z ∞
1
p
dx
1 x
and then Z ∞
1 π
2
dx = lim tan−1 z = .
0 1 + x z→∞ 2
For example
Z 0
1 π
dx = lim (− tan−1 z) = .
−∞ 1 + x2 z→−∞ 2
Remark 7.4.6. (1) More precisely, if for any given a ∈ R, the integrals
Z a Z ∞
f (x) dx, f (x) dx
−∞ a
R∞
are convergent, then we define the integral −∞ f (x) dx via (7.4.5). In this case, we
can show that the definition (7.4.5) does not depend on the particular choice of a.
Indeed, for any given two real numbers a and b, say a < b, we have
Z ∞ Z z "Z #
z Z b
f (x) dx = lim f (x) dx = lim f (x) dx + f (x) dx
a z→∞ a z→∞ b a
Z ∞ Z b
= f (x) dx + f (x) dx.
b a
Similarly, we can show
Z b Z b Z a
f (x) dx = f (x) dx + f (x) dx.
−∞ a −∞
7.4. IMPROPER INTEGRALS 71
Let
f (x) := ex − x, x ∈ [0, ∞).
Since f 0 (x) = ex − 1 ≥ e0 − 1 = 0, it follows from Proposition 5.2.2 that f (x) is
increasing on [0, b] for each fixed b and then on [0, ∞). Hence
(7.4.7) ex − x = f (x) ≥ f (0) = e0 − 0 = 1 =⇒ ex ≥ 1 + x, x ∈ [0, ∞).
We now use the inequality (7.4.8) to show that the improper integral (7.2.4) is
convergent at least for x ≥ 1.
provided that this limit exits. Note that the limit in (7.4.12) is the limit of multi-
variable which will be treat with in later.
For example
Z 1 1
dx 1/3
2/3
= 3(x − 1) = 3,
0 (x − 1) 0
Z 1 1
ln x dx = (x ln x − x) = −1 − lim x ln x = −1.
0 x→0+
0
For example,
Z 1 Z 1 Z −
1 dx dx
p.v. dx = lim + = 0.
−1 x3 →0+ x3 −1 x3
Proposition 7.4.11. Suppose that fR(x) and g(x) are continuous onR[a, ∞).
∞ ∞
(1) If f (x) ≤ g(x) on [a, ∞ and a g(x) dx is convergent, then a f (x) dx is
also convergent and
Z ∞ Z ∞
(7.4.15) f (x) dx ≤ g(x) dx.
a a
R∞ R∞
(2) If f (x) ≥ g(x) on [a, ∞) and a g(x) dx is divergent, then a f (x) dx is
also divergent.
Proof. (1) In this case, for any fixed number z > a, we have
Z z Z z
f (x) dx ≤ g(x) dx;
a a
0 00 (n)
Pn fi (0), f (0), · · · , f (0) exist.
provided all derivatives
For example, i=0 x /i! in (7.4.8) is the Taylor polynomial of degree n about
x = 0 for ex .
More general, the Taylor polynomial of degree n about x = a for the function
f (x) is given by
Exercise 7.5.1. Compute the Taylor polynomial of degree 3 about x = 0 for sin x
and cos x.
That is
Z x
f (x) = f (a) + f 0 (a)(x − a) + (x − t)f 00 (t) dt.
a
In general
Proof. Since f (n+1) (t) is continuous, it follows that m ≤ f (n+1) (t) ≤ M for
all t ∈ I. In particular,
m x M x
Z Z
(x − t)n dt ≤ Rn+1 (x) ≤ (x − t)n dt,
n! a n! a
m M
which gives us (n+1)! (x − a)n+1 ≤ Rn+1 (x) ≤ (n+1)! (x − a)n+1 . Thus m ≤
(n+1)!Rn+1 (x) n+1
(x−a)m+1 ≤ M. By Theorem 3.2.7, (n + 1)!Rn+1 (x)/(x − a) = f (n+1) (c)
for some c between a and x.
For example,
n
x
X xi
(7.5.6) e = + Rn+1 (x), x ∈ (−∞, ∞),
i=0
i!
n
X x2i+1
(7.5.7) sin x = (−1)i + Rn+1 (x), x ∈ (−∞, ∞),
i=0
(2i + 1)!
n
X x2i
(7.5.8) cos x = (−1)i + Rn+1 (x), x ∈ (−∞, ∞),
i=0
(2i)!
n
X xi
(7.5.9) ln(1 + x) = (−1)i+1 + Rn+1 (x), x ∈ (−1, 1],
i=1
i
n
1 X
(7.5.10) = xi + Rn (x), x ∈ (−1, 1).
1−x i=0
∞
X xi
(7.5.11) ex = , x ∈ (−∞, ∞),
i=0
i!
∞
X x2i+1
(7.5.12) sin x = (−1)i , x ∈ (−∞, ∞),
i=0
(2i + 1)!
∞
X x2i
(7.5.13) cos x = (−1)i , x ∈ (−∞, ∞),
i=0
(2i)!
∞
X xi
(7.5.14) ln(1 + x) = (−1)i+1 , x ∈ (−1, 1),
i=1
i
∞
1 X
(7.5.15) = xi , x ∈ (−1, 1).
1−x i=0
7.5. TAYLOR APPROXIMATION 77
A function f (x) defined on an open interval I is called smooth if all its higher
order derivatives are continuous. For example ex , sin x, cos x are smooth functions
on (−∞, ∞). The set of all smooth functions defined on I is denoted by C ∞ (I).
We say f ∈ C k (I) if all higher order derivative up to kth order are continuous.
Observe that
dn x
Since dxn e = ex , it follows that ex is a smooth function. In general, we have
Proposition 7.5.5. For any open interval I, we have C ω (I) ⊂ C ∞ (I) and C ω (I) 6=
C ∞ (I). If f ∈ C ω (I), then, for any a ∈ I,
∞
X f (n) (a)
(7.5.17) f (x) = (x − a)n
n=0
n!
for any x near a.
Calculate
1 −1/x
f 0 (x) = e ,
x2
1 −1/x
00 −2x −1/x 1 2
f (x) = e + 4 = − 3 e−1/x ,
x4 x x4 x
3 2
000 −4x 6x 1 2 −1/x 1 6 6
f (x) = + 6 + 6− 5 e = − 5 + 4 e−1/x .
x8 x x x x6 x x
By induction on n, we can show that
1
(n)
f (x) = P2n e−1/x
x
where P2n (t) denotes the polynomial (without constant term) of degree 2n in terms
of t. Hence
P2n (y)
lim f (n) (x) = lim ;
x→0+ y→∞ ey
1
by (7.4.8), we have ey ≥ (2n+1)! y 2n+1 and then
(2n + 1)!P2n (y)
lim f (n) (x) ≤ lim =0
x→0+ y→∞ y 2n+1
by (3.1.8). Thus f (x) is smooth.
We now show that f (x) is not analytic. Otherwise,
∞
X
f (x) = cn xn
n=0
around 0. By (7.5.17), we must have f (x) ≡ 0 near 0. But, for any > 0, by the
definition, we get f (x) > 0, a contradiction. Therefore, f (x) ∈ C ∞ ((−∞, ∞)) but
/ C ω ((−∞, ∞)).
f (x) ∈
α α(α − 1) · · · (α − n + 1)
(7.5.19) = .
n n!
When α is a positive integer greater than n, the definition (7.5.19) coincides with
the usual binomial coefficients.
We have
∞
X (2n)!
(7.5.21) sin−1 x = x2n+1 , |x| ≤ 1,
n=0
4n (n!)2 (2n + 1)
π
(7.5.22) cos−1 x = − sin−1 x, |x| ≤ 1,
2
∞
X (−1)n 2n+1
(7.5.23) tan−1 x = x , |x| ≤ 1.
n=0
2n + 1
Exercise 7.5.9. Evaluate the first two terms of their Taylor polynomials:
1
tanh x = x − x3 + · · · ,
3
1 1 1
coth x − = x − x3 + · · · .
x 3 45