Sie sind auf Seite 1von 101

Chapter 1.

Limits and Continuity


1.1. Rates of Change and Limits

Definition. The average rate of change of y = f (x) with respect to x


over the interval [x1, x2] is
∆y f (x2) − f (x1 ) f (x1 + h) − f (x1 )
= =
∆x x2 − x1 h
where h = x2 − x1.

Figure 1.1.1, page 87

1
Example. Page 95 number 4.

Definition. Informal Definition of Limit.


Let f (x) be defined on an open interval about x0, except possibly at
x0 itself. If f (x) gets arbitrarily close to L for all x sufficiently close to
x0, we say that f approaches the limit L as x approaches x0, and we
write
lim f (x) = L.
x→x0

Note. The above definition is informal (that is, it is not mathematically


rigorous) since the terms “arbitrarily close” and “sufficiently close” are not
defined.

Example. Page 96 number 10.

Example. Page 97 number 20.

2
Definition. Formal Definition of Limit
Let f (x) be defined on an open interval about x0, except possibly at x0
itself. We say that f (x) approaches the limit L as x approaches x0 and
write
lim f (x) = L,
x→x0

if, for every number  > 0, there exists a corresponding number δ > 0
such that for all x,

0 < |x − x0| < δ ⇒ |f (x) − L| < .

Figure 1.1.11, page 93


3
Example. Prove for f (x) = mx + b, m = 0, that x→a
lim f (x) = f (a).

Example. Page 97 number 30.

Example. Page 97 number 34.

Example. Page 98 number 42.

4
Chapter 1. Limits and Continuity
1.2. Finding Limits and One-Sided Limits

Theorem 1. Limit Rules.


If L, M , c, and k are real numbers and

lim f (x) = L
x→c
and lim g(x) = M,
x→c
then

lim (f (x) + g(x)) = L + M.


1. Sum Rule: x→c

lim (f (x) − g(x)) = L − M.


2. Difference Rule: x→c

lim (f (x) · g(x)) = L · M.


3. Product Rule: x→c

lim (k · f (x)) = k · L.
4. Constant Multiple Rule: x→c
f (x) L
5. Quotient Rule: x→c
lim = , M = 0.
g(x) M
6. Power Rule: If r and s are integers, s = 0, then

r/s r/s
lim
x→c
(f (x)) = L

provided that Lr/s is a real number AND L > 0.

1

Note. We must have L > 0 in part 6 of Theorem 1 since lim x does
x→0
not exist. NOTICE THAT THERE IS AN ERROR IN THE
TEXT!!!

lim (f (x) + g(x)) =


Proof of Theorem 1, part 1. We wish to prove x→c
L+M under the assumptions x→c
lim f (x) = L and x→c
lim g(x) = M . Let  > 0
be given. Then /2 > 0 and there exists δ1 > 0 such that for all x with
0 < |x − c| < δ1 we have |f (x) − L| < /2. Similarly, there exists δ2 > 0
such that for all x with 0 < |x − c| < δ2 we have |g(x) − M | < /2.
Therefore we choose δ = min{δ1, δ2}. Then for 0 < |x − c| < δ we have

|(f (x) + g(x)) − (L + M )| ≤ |(f (x) − L) + (g(x) − M )|

≤ |f (x) − L| + |g(x) − M |
 
< +
2 2
= .

This proves the result. Q.E.D.

Note. For proofs of parts 2 through 5 of Theorem 1, see pages 1147–1148.


Notice that the text does not provide a proof of part 6 (since as the text
states it, it is false!).
2
Example. Page 109 number 8.

Theorem 2. Limits of Polynomials Can Be Found by Substi-


tution.
If P (x) = anxn + an−1xn−1 + · · · + a0 then
n n−1
lim
x→c
P (x) = P (c) = a n c + a n−1 c + · · · + a0 .

Theorem 3. Limits of Rational Functions Can Be Found by


Substituting IF the Limit of the Denominator Is Not Zero.
If P (x) and Q(x) are polynomials and Q(c) = 0, then
P (x) P (c)
lim
x→c
= .
Q(x) Q(c)

Example. Page 109 number 12a.

Theorem. Dr. Bob’s Theorem. (NOT IN 10TH EDITION!)


If f (x) = g(x) for all x in an open interval containing c, except possibly
c itself, then
lim f (x) = x→c
x→c
lim g(x)

provided these limits exist.


3
Note. We have to be careful in our dealings with functions! Notice that
x(x − 1)
f (x) = and g(x) = x are NOT the same functions! They
x−1
do not even have the same domains. Therefore we cannot in general say
x(x − 1)
= x. However, this equality holds if x lies in the domains of the
x−1
functions. We can say:
x(x − 1)
= x IF x = 1.
x−1
We can also say f (x) = g(x) IF x = 1. If we are concerned with limits
as x approaches 1, then from the definition, x IS NOT EQUAL TO
1 (but near 1). Therefore we can say lim f (x) = lim g(x). We have not
x→1 x→1
said that the functions are equal, but that their limits are.

Example. Page 109 number 14b.

Example. Page 109 number 14a.

4
Theorem 4. Sandwich Theorem.
Suppose that g(x) ≤ f (x) ≤ h(x) for all x in some open interval contain-
ing c, except possibly at x = c itself. Suppose also that

lim g(x) = x→c


x→c
lim h(x) = L.

lim f (x) = L.
Then x→c

Figure 1.2.17, page 102

Example. Page 109 number 16.

5
Definition. Informal Definition of Right-Hand and Left-Hand
Limits.
Let f (x) be defined on an interval (a, b), where a < b. If f (x) approaches
arbitrarily close to L as x approaches a from within that interval, then we
say that f has right-hand limit L at a, and write

lim f (x) = L.
x→a+

Let f (x) be defined on an interval (c, a), where c < a. If f (x) approaches
arbitrarily close to M as x approaches a from within the interval (c, a),
then we say that f has left-hand limit M at a, and we write

lim f (x) = M.
x→a−

Definition. Formal Definitions of One Sided Limits. (NOT IN


10TH EDITION!)
We say that f (x) has right-hand limit L at x0, and write

lim f (x) = L
x→x+
0

if for every number  > 0 there exists a corresponding number δ > 0 such
that for all x

x 0 < x < x0 + δ ⇒ |f (x) − L| < .


6
We say that f (x) has left-hand limit L at x0, and write

lim f (x) = L
x→x−
0

if for every number  > 0 there exists a corresponding number δ > 0 such
that for all x

x 0 − δ < x < x0 ⇒ |f (x) − L| < .

Example. Consider limits as x approaches −1 and +1 for f (x) =



1 − x2 .

Theorem 5. Relation Between One-Sided and Two-Sided


Limits
A function f (x) has a limit as x approaches c if and only if it has left-hand
and right-hand limits there and these one-sided limits are equal:

lim f (x) = L
x→c
⇔ lim f (x) = L and lim f (x) = L.
x→c− x→c+

Example. Page 110 number 26.

Example. Page 110 number 42.


7
Theorem 6.
For θ in radians,
sin θ
lim = 1.
θ→0 θ

Proof. Suppose first that θ is positive and less than π/2. Consider the
picture:

Figure 1.2.25, page 106

Notice that

Area ∆OAP < area sector OAP < area ∆OAT .


8
We can express these areas in terms of θ as follows:

1
Area ∆OAP = 2
base × height = 12 (1)(sin θ) = 12 sin θ

Area sector OAP = 12 r2θ = 12 (1)2θ = θ


2

1
Area ∆OAT = 2 base × height = 12 (1)(tan θ) = 12 tan θ.

Thus,
1 1 1
sin θ < θ < tan θ.
2 2 2
This last inequality goes the same way if we divide all three terms by the
positive number (1/2) sin θ:
θ 1
1< < .
sin θ cos θ
Taking reciprocals reverses the inequalities:
sin θ
cos θ < < 1.
θ
Since lim+ cos θ = 1, the Sandwich Theorem gives
θ→0

sin θ
lim+ = 1.
θ→0 θ
sin θ
Since sin θ and θ are both odd functions, f (θ) = is an even function
θ
sin(−θ) sin θ
and hence = . Therefore
−θ θ
sin θ sin θ
lim− = 1 = lim+ ,
θ→0 θ θ→0 θ
9
sin θ
so lim = 1 by Theorem 4. QED
θ→0 θ

cos h − 1
Example. Page 107 example 10a: Show that lim = 0.
h→0 h

Solution. We use the trig identity cos h = 1 − 2 sin2(h/2) (a half-angle


identity). First, we have
cos h − 1 2 sin2(h/2)
lim = lim − .
h→0 h h→0 h
Now, replacing h/2 with θ we get
cos h − 1 2 sin2(h/2) sin θ
lim = lim − = − lim sin θ = −(1)(0) = 0.
h→0 h h→0 h θ→0 θ

QED

10
Chapter 1. Limits and Continuity
1.3. Limits Involving Infinity

Definition. Formal Definition of Limits at Infinity. (NOT IN


10TH EDITION!)

1. We say that f (x) has the limit L as x approaches infinity and we


write
lim f (x) = L
x→+∞

if, for every number  > 0, there exists a corresponding number M


such that for all x

x>M ⇒ |f (x) − L| < .

2. We say that f (x) has the limit L as x approaches negative infinity


and we write
lim f (x) = L
x→−∞

if, for every number  > 0, there exists a corresponding number N


such that for all x

x<N ⇒ |f (x) − L| < .

1
Definition. Informal Definition of Limits Involving Infinity.

1. We say that f (x) has the limit L as x approaches infinity and write

lim f (x) = L
x→+∞

if, as x moves increasingly far from the origin in the positive direction,
f (x) gets arbitrarily close to L.

2. We say that f (x) has the limit L as x approaches negative infinity


and write
lim f (x) = L
x→−∞

if, as x moves increasingly far from the origin in the negative direction,
f (x) gets arbitrarily close to L.

1
Example. Example 1 page 112. Show that x→∞
lim = 0.
x

Solution. Let  > 0 be given. We must find a number M such that for
all
   
1   1
   
x>M ⇒ − 0 =   < .



x  x
The implication will hold if M = 1/ or any larger positive number (see
1
the figure below). This proves x→∞
lim = 0. We can similarly prove that
x
1
lim = 0. QED
x→−∞ x

2
Figure 1.3.34, page 221 of 9th edition.

Theorem 7. Rules for Limits as x → ±∞.


If L, M , and k are real numbers and

lim f (x) = L and lim = M, then


x→±∞ x→±∞

1. Sum Rule: lim (f (x) + g(x)) = L + M


x→±∞

2. Difference Rule: lim (f (x) − g(x)) = L − M


x→±∞

3. Product Rule: lim (f (x) · g(x)) = L · M


x→±∞

3
4. Constant Multiple Rule: lim (k · f (x)) = k · L
x→±∞

f (x) L
5. Quotient Rule: lim = , M = 0
x→±∞ g(x) M
6. Power Rule: If r and s are integers, s = 0, then

lim (f (x))r/s = Lr/s


x→±∞

provided that Lr/s is a real number AND L > 0.

Note. As in section 1.1, there is an error in the text with part 6 of


1
Theorem 7, as can be seen by considering lim √ which clearly does
x→−∞ x
not exist (but would be 0 by the use of Theorem 7, part 6 as stated in the
text, with f (x) = 1/x, r = 1 and s = 2).

Example. Page 122 number 10.

4
Definition. Infinite Limits

1. We say that f (x) approaches infinity as x approaches x0, and we


write
lim f (x) = ∞,
x→x0

if for every positive real number B there exists a corresponding δ > 0


such that for all x

0 < |x − x0| < δ ⇒ f (x) > B.

2. We say that f (x) approaches negative infinity as x approaches x0,


and we write
lim f (x) = −∞,
x→x0

if for every negative real number −B there exists a corresponding


δ > 0 such that for all x

0 < |x − x0| < δ ⇒ f (x) < −B.

5
Figure 1.3.39 and 1.3.40, page 119.

lim f (x) = ∞ if f (x) can be made arbitrarily large by


Note. Informally, x→x
0

making x sufficiently close to x0 (and similarly for f approaching negative


infinity). We can also define one-sided infinite limits in an analogous
manner.

6
Definition. Horizontal and Vertical Asymptotes.
A line y = b is a horizontal asymptote of the graph of a function y = f (x)
if either
lim f (x) = b
x→∞
or lim f (x) = b.
x→−∞

A line x = a is a vertical asymptote of the graph if either

lim f (x) = ±∞ or lim f (x) = ±∞.


x→a+ x→a−

Example. Page 122 number 32.

Definition. End Behavior Model


The function g is

(a) a right end behavior model for f if and only if


f (x)
lim
x→∞
=1
g(x)
(b) a left end behavior model for f if and only if
f (x)
lim = 1.
x→−∞ g(x)

Definition. If g(x) = mx + b where m = 0 is an end behavior model for


f , then f is said to have an oblique (or slant) asymptote of y = mx + b.

Example. Page 122 number 38.


7
Chapter 1. Limits and Continuity
1.4. Continuity

Definition. Continuity at a Point.


Interior Point: A function y = f (x) is continuous at an interior
point c of its domain if
lim f (x) = f (c).
x→c

Endpoint: A function y = f (x) is continuous at a left endpoint a or


is continuous at a right endpoint b of its domain if

lim f (x) = f (a) or lim f (x) = f (b), respectively.


x→a+ x→b−

Figure 1.4.46, page 125.

1
Note. If a function is continuous at all interior points of its domain and
the domain is an interval, then the function can be “drawn without picking
up your pencil.”

Example. Page 132 number 4.

Continuity Test.
A function f (x) is continuous at an interior point of the domain of f ,
x = c if and only if it meets the following three conditions:

1. f (c) exists,

lim f (x) exists, and


2. x→c

lim f (x) = f (c).


3. x→c

Note. Polynomials, rational functions, and the six trigonometric func-


tions are continuous at every point of their domains.

2
Example. Consider the piecewise defined function









x if x ∈ (−∞, 0)



f (x) = 

0 if x = 0 .







 x2 if x ∈ (0, ∞)

Is f continuous at x = 0?

Definition. A function f has a removable discontinuity at x = a if


f (a) can be redefined in such a way that f is continuous at a. f has a
jump discontinuity at x = a if lim− f (x) and lim+ f (x) exist (as finite
x→a x→a
numbers) and are different.

|x|
Example. Discuss the discontinuities of f (x) = and g(x) = int x.
x

3
Theorem 8. Properties of Continuous Functions
If the functions f and g are continuous at x = c, then the following
combinations are continuous at x = c.

1. Sums: f + g

2. Differences: f − g

3. Products: f · g

4. Constant Multiples: k · f , for any number k

5. Quotients: f /g, provided g(c) = 0.

Theorem 9. Composite of Continuous Functions


If f is continuous at c and g is continuous at f (c), then the composite
g ◦ f is continuous at c.

Figure 1.4.53, page 129.


4
Example. Page 132 number 22.

Theorem. The Intermediate Value Theorem for Continuous


Functions
A function y = f (x) that is continuous on a closed interval [a, b] takes
on every value between f (a) and f (b). In other words, if y0 is any value
between f (a) and f (b), then y0 = f (c) for some c in [a, b].

Figure 1.4.44, page 93 of 9th edition.

Example. Page 133 number 28.

5
Chapter 1. Limits and Continuity
1.5. Tangent Lines

Definition. Slope and Tangent Line.


The slope of the curve y = f (x) at the point P (x0, f (x0)) is the number
f (x0 + h) − f (x0)
m = lim ,
h→0 h
provided the limit exists. The tangent line to the curve at P is the line
through P with this slope.

Figure 1.5.62, page 136

Example. Page 139 number 6.


1
Definition. If f (t) is the position of a particle at time t, then the
instantaneous rate of change of position with respect to time (i.e. the
instantaneous velocity) is
f (t + h) − f (t)
lim ,
h→0 h
provided the limit exists.

Example. Page 139 number 24 (this one is dumb!!!).

Example. Page 139 number 16.

2
Chapter 2. Derivatives
2.1. The Derivative as a Function

Definition. Derivative Function.


The derivative of the function f (x) with respect to the variable x is the
function f  whose value at x is
f (x + h) − f (x)
f  (x) = lim ,
h→0 h
provided the limit exists.

Note. Motivated by Section 1.5, we see that f  (x) is the slope of the line
tangent to y = f (x) as a function of x.

Note. There are a number of ways to denote the derivative of y = f (x):


df dy d
f  (x) = y  = = = [f ].
dx dx dx

Example. Page 157 number 3.

1
Note. We can also calculate higher order derivatives:
d   d d d
y  = [y ], y = [y  ], y (4) = [y  ], . . . , y (n) = [y (n−1) ].
dx dx dx dx

Example. Page 157 number 10.

Rule 1. Derivative of a Constant Function.


If f has the constant value f (x) = c, then
df d
= [c] = 0.
dx dx

Proof. From the definition:


f (x + h) − f (x) c−c
f  (x) = lim = lim = lim 0 = 0.
h→0 h h→o h h→0

QED

Rule 2. Power Rule for Positive Integers


If n is a positive integer, then
d n
[x ] = nxn−1.
dx

Note. Before we present the proof of the Power Rule, we introduce the
Binomial Theorem.
2
Theorem. Binomial Theorem
Let a and b be real numbers and let n be a positive integer. Then
n(n − 1) n−2 2
(a + b)n = an + nan−1b + a b + . . . + nabn−1 + bn
  2
n 
  n 

 
= 


 an−i bi
i=0  i 

 
 


n 
 n!
where 


 = and i! = (i)(i − 1)(i − 2) · · · (3)(2)(1).

i  (n − i)!i!

Note. We can prove the Binomial Theorem using Mathematical Induc-


tion.

Proof of the Power Rule. By definition,


f (x + h) − f (x)
f  (x) = lim
h→0 h
(x + h) − xn
n
= lim
h 
h→0
 
n 

n 

i=0 


 xn−i hi − xn
 
i
= lim
h→0
 h 
 
n 

n 

n
x + i=1 


 xn−i hi − xn
 
i
= lim
h→0 h
3
 
 
n 

n 

i=1 


 xn−i hi
 
i
= lim
h→0
 h 
 
n 

n 

h i=1 


 xn−i hi−1
 
i
= lim
h→0
  h
n 
  n 

 
= lim   xn−i hi−1
h→0 i=1 



i
 

n 
  n 

n−1  
= lim nx + 


 xn−i hi−1
h→0 i=2  i 

= nxn−1

QED

Note. See page 150 for a proof of the Power Rule that doesn’t (explicitly)
use the Binomial Theorem. Can you find the error in the computation
(it’s subtle and conceptual, but does not affect the conclusion)?

4
Rule 3. Constant Multiple Rule
If u is a differentiable function of x, and c is a constant, then
d du
[cu] = c .
dx dx

Rule 4. Derivative Sum Rule


If u and v are differentiable functions of x, then their sum u + v is dif-
ferentiable at every point where u and v are both differentiable. At such
points,
d du dv
[u + v] = + .
dx dx dx

Note. The proofs of Rules 3 and 4 follow from the corresponding rules
for limits (namely, the Constant Multiple Rule and the Sum Rule, respec-
tively).

Corollary. If P (x) = an xn + an−1xn−1 + · · · + a2x2 + a1x + a0, then


P (x) = nanxn−1 + (n − 1)an−1xn−1 + · · · + 2a2x + a1.

Example. Page 157 number 12.

5
Theorem 1. Differentiability Implies Continuity
If f has a derivative at x = c, then f is continuous at x = c.

lim f (x) = f (c), or equiva-


Proof. By definition, we need to show that x→c
lently that lim f (c + h) = f (c). Then
h→0
 
f (c + h) − f (c) 
lim f (c + h) = lim f (c) + · h
h→0 h→0 h
f (c + h) − f (c)
= lim f (c) + lim · lim h
h→0 h→0 h h→0

= f (c) + f  (c) · 0

= f (c).

Therefore f is continuous at x = c. QED

Theorem 2. Intermediate Value Property of Derivatives


If a and b are any two points in an interval on which f is differentiable,
then f  takes on every value between f  (a) and f  (b).

Example. Page 159 number 34.

6
Chapter 2. Derivatives
2.2. The Derivative as a Rate of Change

Definition. Instantaneous Rate of Change


The instantaneous rate of change of f with respect to x at x0 is the
derivative
f (x0 + h) − f (x0)
f  (x0) = lim ,
h→0 h
provided the limit exists.

Definition. (Instantaneous) Velocity


Velocity (instantaneous velocity) is the derivative of position with respect
to time. If a body’s position at time t is s = f (t), then the body’s velocity
at time t is
ds f (t + ∆t) − f (t)
v(t) = = lim .
dt ∆t→0 ∆t

Definition. Speed
Speed is the absolute value of velocity.
 

ds 

Speed = |v(t)| = 



dt 

1
Definition. Acceleration, Jerk
Acceleration is the derivative of velocity with respect to time. If a body’s
position at time t is s = f (t), then the body’s acceleration at time t is
dv d2s
a(t) = = 2.
dt dt
Jerk is the derivative of acceleration with respect to time:
da d3s
j(t) = = 3.
dt dt

Example. Page 169 number 10.

Note. At the surface of the Earth, if an object is fired directly upward


with an initial (upward) velocity v0 from an initial height s0 , then the
height of the object at time t is

s(t) = −16t2 + v0t + s0

if time is measured in seconds and distances are measured in feet, or

s(t) = −4.9t2 + v0t + s0

if time is measured in seconds and distances are measured in meters. No-


tice what this implies that the accelerations are.

2
Example. Page 171 number 18.

Note. In economics, the term “marginal” is used when referring to


derivatives. If a company produces and sells a number x of objects, and
the cost of producing those objects is c(x) and the revenue that results
from selling them is r(x), then the resulting profit is p(x) = r(x) − c(x).
The functions p(x), r (x), and c (x) are the marginal profit, revenue, and
cost funtions, respectively.

Example. Page 171 number 20.

3
Chapter 2. Derivatives
2.3. Derivative of Products, Quotients and Negative Powers

Rule 5. Derivative Product Rule


If u and v are differentiable at x, then so is their product uv, and
d du dv
[uv] = v + u = [u ]v + u[v  ].
dx dx dx

Proof. By definition we have:


d u(x + h)v(x + h) − u(x)v(x)
[uv] = lim
dx h→0 h
u(x + h)v(x + h) − u(x + h)v(x) + u(x + h)v(x) − u(x)v(x)
= lim
h→0
 h 
v(x + h) − v(x) u(x + h) − u(x) 
= lim 
u(x + h) + v(x) 
h→0 h h
v(x + h) − v(x) u(x + h) − u(x)
= lim u(x + h) · lim + v(x) · lim
h→0 h→0 h h→0 h
= u(x)[v (x)] + [u(x)]v(x)

where lim u(x + h) = u(x) since u is continuous at x by Theorem 1 of


h→0
section 2.1. QED

Example. Differentiate f (x) = (4x3 − 5x2 + 4)(7x2 − x).

1
Rule 6. Derivative Quotient Rule
If u and v are differentiable at x and if v(x) = 0, then the quotient u/v
is differentiable at x, and
d u du
dx
v − u dx
dv
[u]v − u[v ]
= = .
dx v v2 v2

Proof. By definition we have:

d u
u(x+h)
v(x+h)
− u(x)
v(x)
= lim
dx v h→0 h
v(x)u(x + h) − u(x)v(x + h)
= lim
h→0 hv(x + h)v(x)
v(x)u(x + h) − v(x)u(x) + v(x)u(x) − u(x)v(x + h)
= lim
h→0 hv(x + h)v(x)
v(x) u(x+h)−u(x)
h − u(x) v(x+h)−v(x)
h
= lim
x→0 v(x + h)v(x)
limh→0 v(x) u(x+h)−u(x)
h − limh→0 u(x) v(x+h)−v(x)
h
=
limh→0 v(x + h)v(x)
v(x) limh→0 u(x+h)−u(x)
h − u(x) limh→0 v(x+h)−v(x)
h
=
v(x) limh→0 v(x + h)
v(x)u (x) − u(x)v (x)

=
v 2(x)
QED

Example. Page 178 number 10.

2
Rule 7. Power Rule for Negative Integers
If n is a negative integer and x = 0, then
d n
[x ] = nxn−1.
dx

Proof. If n is a negative integer, then −n is a positive integer and we


can use the Power Rule for Nonnegative Integers to differentiate x−n . So
 
d n d  1 
[x ] =
dx dx x−n
−n
d
dx [1] (x ) − (1) d
dx [x−n ]
=
(x−n )2
[0](x−n) − (1)[(−n)x−n−1]
=
x−2n
= nxn−1

QED

d n
Note. We have now established that [x ] = nxn−1 for all integers n.
dx
We will eventually see that this is the way xn is differentiated for all real
numbers n, but we have not even defined what it means to raise a real
number to an irrational number! We will take care of this when we define
the natural logarithm and exponential functions.

Example. Page 179 number 30.


3
Example. Page 178 number 16.

Note. We will follow my “square brackets” notation as described in the


handout.

4
Chapter 2. Derivatives
2.4. Derivative of Trigonometric Functions

Recall. For all real numbers a and b,

sin(a + b) = sin a cos b + cos a sin b.

Theorem. Derivative of the Sine Function

d
[sin x] = cos x
dx

Proof. Let y = sin x. By definition we have


dy sin(x + h) − sin x
= lim
dx h→0 h
(sin x cos h + cos x sin h) − sin x
= lim
h→0 h
sin x(cos h − 1) + cos x sin h
= lim
h→0

h   
cos h − 1 sin h
= lim sin x ·  + lim cos x · 
h→0 h h→0 h
cos h − 1 sin h
= sin x · lim + cos x · lim
h→0 h h→0 h

= sin x · 0 + cos x · 1

= cos x.
1
cos h − 1 sin h
We have lim = 0 and lim = 1 by the results in section 1.2.
h→0 h h→0 h
QED

Example. Page 184 number 2.

Recall. For all real numbers a and b we have

cos(a + b) = cos a cos b − sin a sin b.

Theorem. Derivative of the Cosine Function

d
[cos x] = − sin x
dx

Proof. By definition we have


d cos(x + h) − cos x
[cos x] = lim
dx h→0 h
(cos x cos h − sin x sin h) − cos x
= lim
h→0 h
cos x(cos h − 1) − sin x sin h
= lim
h→0 h
cos h − 1 sin h
= lim cos x · − lim sin x ·
h→0 h h→0 h
cos h − 1 sin h
= cos x · lim − sin x · lim
h→0 h h→0 h

2
= cos x · 0 − sin x · 1

= − sin x.

QED

Examples. Page 184 number 22, page 183 Example 5, and page 185
number 44.

Note. In summary, we have the following derivatives of the six trigono-


metric functions:

f f

sin x cos x
cos x − sin x
tan x sec2 x

cot x −csc2 x
sec x sec x tan x
csc x − csc x cot x

Example. Page 185 number 32.

3
Chapter 2. Derivatives
2.5. The Chain Rule and Parametric Equations

Theorem 3. The Chain Rule.


If f (u) is differentiable at the point u = g(x) and g(x) is differentiable at
x, then the composite function (f ◦ g)(x) = f (g(x)) is differentiable at x,
and
(f ◦ g) (x) = f  (g(x))[g  (x)].

In Leibniz’s notation, if y = f (u) and u = g(x), then


dy dy du
= · ,
dx du dx
where dy/du is evaluated at u = g(x).

Note. The proof of the Chain Rule is rather complicated — see Appendix
3.

Note. If f (u) = un where n is an integer, then


d d
[f (g(x))] = [(g(x))n] = ng(x)n−1[g  (x)].
dx dx

1
Examples. Page 195 numbers 6, 24, 44.

Definition. A curve is given parametrically if its graph is determined


by graphing (x(t), y(t)) for t (the parameter) ranging over some interval.

Example. x(t) = cos t, y(t) = sin t for t ∈ [0, 2π) parametrically


determine the unit circle x2 + y 2 = 1.

Note. Parametric Formula for dy/dx and d2y/dx2 .


If x = x(t) and y = y(t) where x(t) and y(t) are differentiable with
x(t) = 0 then
dy dy/dt d2y dy  /dt
= and 2 = .
dx dx/dt dx dx/dt

Examples. Pages 196-97 numbers 54 and 64.

2
Chapter 2. Derivatives
2.6. Implicit Differentiation

Definition. The function f (x) is implicit to the equation F (x, y) = 0


if the substitution y = f (x) into the equation yields an identity.

√ √
Example. The functions f (x) = 1 − x2 and g(x) = − 1 − x2 are
implicit to the equation x2 + y 2 = 1. Can you find other functions implicit
to this equation?

Note. If y is a function implicit to F (x, y) = 0, then we can generate an


equation containing dy/dx by differentiating “implicitly.” This follows by
applying the Chain Rule.

1
Example. Suppose y = f (x) is implicit to x2 + y 2 = 1. Then differenti-
ating implicity:
d 2 d
[x + y 2] = [1]
dx  
dx
dy
2x + 2y   = 0
dx
dy x
= − .
dx y
Notice that dy/dx involves both x and y. This is because we cannot find
the slope of a line tangent to the graph of F (x, y) = 0 without knowing
the x and y coordinates of the point of tangency.

Example. Find the slope of the line tangent to x2 + y 2 = 1 at (x, y) =


√ √ √ √
( 2/2, 2/2). Do the same for the point (x, y) = ( 2/2, − 2/2).

Definition. A line is normal to a curve at a point if it is perpendicular


to the curve’s tangent line. The line is called the normal to the curve at
the point.

Examples. Page 204 number 22, page 205 number 44, and page 206
number 56.

2
Theorem 4. Power Rule for Rational Powers.
If n is a rational number, then xn is differentiable at every interior point
of the domain of xn−1 , and
d n
[x ] = nxn−1.
dx

q
Proof. Let p and q be integers with q > 0 and suppose that y = xp =
xp/q . Then y q = xp. By the Power Rule for Integer Exponents,
d q d p
[y ] = [x ]
dx  dx
dy
qy q−1   = pxp−1.
dx
dy
If y = 0, then we can solve for :
dx
dy pxp−1 p xp−1 p (p/q)−1
= q−1 = = x .
dx qy q (xp/q )q−1 q
Qed

3
Chapter 2. Derivatives
2.7. Related Rates

Example. Page 213 number 13.

Note. Related Rates Problem Strategy


We will follow this protocol when solving related rates problems:

Step 1. Draw a picture and name the variables and constants. Use
t for time. Assume that all variables are differentiable functions of
time.

Step 2. Write down the numerical information (in terms of the sym-
bols you have chosen) and write down what you are asked to find
(usually a rate, expressed as a derivative).

Step 3. Write an equation that relates the variables. You may have
to combine two or more equations to get a single equation that relates
the variable whose rate you want to the variables whose rates you
know.

1
Step 4. Differentiate with respect to t. Then express the rate you want
in terms of the rate and variables whose values you know.

Step 5. Evaluate. Use known values to find the unknown rate.

Example. Page 213 number 13 (again).

Examples. Page 213 number 18, page 215 number 30, and page 216
number 38.

2
Chapter 3. Applications of Derivatives
3.1.Extreme Values of Functions

Definition. Let f be a function with domain D. Then f (c) is the

(a) absolute maximum value on D if and only if f (x) ≤ f (c) for all x
in D

(b) absolute minimum value on D if and only if f (x) ≥ f (c) for all x
in D.

Theorem 1.The Extreme-Value Theorem for Continuous Func-


tions
If f is continuous at every point of a closed interval I, then f assumes both
an absolute maximum value M and an absolute minimum value m some-
where in I. That is, there are numbers x1 and x2 in I with f (x1) = m,
f (x2) = M , and m ≤ f (x) ≤ M for every x in I.

Examples. Page 234 numbers 2 and 4.

1
Definition. Let c be an interior point of the domain of the function f .
Then f (c) is a

(a) local maximum value if and only if f (x) ≤ f (c) for all x in some
open interval containing c

(b) local minimum value if and only if f (x) ≥ f (c) for all x in some
open interval containing c.

Theorem 2. Local Extreme Values.


If a function f has a local maximum value or a local minimum value at
an interior point c of its domain, and if f  exists at c, then f  (c) = 0.

Definition. A point in the domain of a function f at which f  = 0 or f 


does not exist is a critical point of f .

2
Note. How to Find the Absolute Extrema of a Continuous
Function f on a Closed Interval
To find extrema on a closed interval, we first find the critical points and
then:

Step 1. Evaluate f at all critical points and endpoints.

Step 2. Take the largest and smallest of these values.

Examples. Page 234 number 16, page 235 number 34, and page 236
number 52.

3
Chapter 3. Applications of Derivatives
3.2. The Mean Value Theorem and Differential Equations

Theorem 3. Rolle’s Theorem.


Suppose that y = f (x) is continuous at every point of [a, b] and differen-
tiable at every point of (a, b). If f (a) = f (b) = 0, then there is at least
one number c in (a, b) at which f  (c) = 0.

Proof. Since f is continuous by hypothesis, f assumes an absolute max-


imum and minimum for x ∈ [a, b] by Theorem 1 (the Extreme Value
Theorem). These extrema occur only

1. at interior points where f  is zero

2. at interior points where f  does not exist

3. at the endpoints of the function’s domain, a and b.

Since we have hypothesized that f is differentiable on (a, b), then Option


2 is not possible.
In the event of Option 1, the point at which an extreme occurs, say
c, must satisfy f  (c) = 0 by Theorem 2 of Section 3.1 (Local Extreme
Values). Therefore the theorem holds.
1
In the event of Option 3, the maximum and minimum occur at the end-
points a and b (where f is 0) and so f must be a constant of 0 throughout
the interval. Therefore f  (x) = 0 for all x ∈ (a, b), by Rule 1 page 149,
and the theorem holds. QED

Example. Page 244 number 38.

Theorem 4. The Mean Value Theorem


Suppose that y = f (x) is continuous on a closed interval [a, b] and differ-
entiable on the interval (a, b). Then there is at least one point c ∈ (a, b)
such that
f (b) − f (a)
f  (c) = .
b−a

Figure 3.2.13, page 238


2
Examples. Page 244 numbers 32, 43.

Corollary 1. Functions with Zero Derivatives Are Constant


Functions.
If f  (x) = 0 at each point of an interval I, then f (x) = k for all x ∈ I,
where k is a constant.

Note. Corollary 1 is the converse of Rule 1 from page 149.

Corollary 2. Functions with the Same Derivative Function


on an Interval Differ by a Constant Value There
If f  (x) = g  (x) at each point of an interval I, then there exists a constant
k such that f (x) = g(x) + k for all x ∈ I.

Proof. Consider the function h(x) = f (x)−g(x). Under our hypothesis,


h(x) is constant on I and so h(x) = 0 for all x ∈ I. So by Corollary 1,
h(x) = k in I. Therefore f (x) − g(x) = k and f (x) = g(x) + k. QED

Example. Page 243 number 14.

3
Definition. A differential equation is an equation relating an unknown
function and one or more of its derivatives. A function whose deriva-
tives satisfy a differential equation is called a solution of the differential
equation.

Example. Page 244 number 22.

4
Chapter 3. Applications of Derivatives
3.3. The Shape of a Graph

Definition. Let f be a function defined on an interval I. Then

1. f increases on I if for all points x1 and x2 in I, x1 < x2 ⇒ f (x1) <


f (x2).

2. f decreases on I if for all points x1 and x2 in I, x1 < x2 ⇒ f (x1) >


f (x2).

Corollary 3. The First Derivative Test for Increasing and


Decreasing.
Suppose that f is continuous on [a, b] and differentiable on (a, b)

If f  > 0 at each point of (a, b), then f increases on [a, b].

If f  < 0 at each point of (a, b), then f decreases on [a, b].

Proof. Suppose x1, x2 ∈ [a, b] with x1 < x2. The Mean Value Theorem
applied to f on [x1, x2] implies that f (x2) − f (x1) = f  (c)(x2 − x1) for
some c between x1 and x2. Since x2 − x1 > 0, then f (x2) − f (x1) and

1
f  (c) are of the same sign. Therefore f (x2) > f (x1) if f  is positive on
(a, b), and f (x2) < f (x1) if f  is negative on (a, b). QED

Example. Page 255 number 18 a,b.

Note. First Derivative Test for Local Extrema.


At a critical point x = c,

1. f has a local minimum if f  changes from negative to positive at c

2. f has a local maximum if f  changes from positive to negative at c

3. f has no local extreme if f  has the sign on both sides of c.

Example. Page 255 number 18 e.

Definition. The graph of a differentiable function y = f (x) is

(a) concave up on an open interval I if y  is increasing on I

(b) concave down on an open interval I if y  is decreasing on I.

2
Note. Second Derivative Test for Concavity.
The graph of a twice-differentiable function y = f (x) is
(a) concave up on any interval where y  > 0
(b) concave down on any interval where y  < 0.

Note. If f is concave up at point (x0, y0), then a tangent line to f at


(x0, y0) lies below the graph of f near (x0, y0). If f is concave down at
point (x0, y0), then a tangent line to f at (x0, y0 ) lies above the graph of
f near (x0, y0).

Definition. A point where the graph of a function has a tangent line


and where the concavity changes is a point of inflection.

Example. Page 255 number 34.

3
Theorem 5. Second Derivative Test for Local Extrema.

1. If f  (c) = 0 and f  (c) < 0, then f has a local maximum at x = c.

2. If f  (c) = 0 and f  (c) > 0, then f has a local minimum at x = c.

Example. Page 254 number 6, page 256 number 50.

4
Chapter 3. Applications of Derivatives
3.4. Graphical Solutions of Autonomous Differential
Equations

dy
Definition. An equation of the form = g(y) is an autonomous
dx
ordinary differential equation.

dy
Definition. If = g(y) is an autonomous differential equation, then
dx
dy
the values of y for which = 0 are called equilibrium values or rest
dx
points.

Example. Page 264 number 2a.

Note. We make use of a phase line for these types of differential equa-
tions. This is a plot on the y-axis that shows the equation’s equilibrium
values along with the intervals where dy/dx and d2y/dx2 are positive and
negative. Then we know where the solutions are increasing and decreasing
and the concavity of the solution curves.

1
Example. Page 264 number 2c, 2b.

Note. An equilibrium value is stable if we perturb the system slightly


and it returns to the equilibrium value. For example, a ball at the bottom
of a well is in a stable equilibrium state — perturb the ball slightly and
it returns to where it was. An equilibrium value is unstable if a small
perturbation of the system may yield solutions that do not return to the
equilibrium value. For example, a ball at the top of a hill is in an unstable
equilibrium — perturb the ball a little and it rolls away.

Example. Page 264 number 2a (continued).

Example. Page 263 Example 5. Consider the logistic equation


dP
= r(M − P )P.
dt

2
Chapter 3. Applications of Derivatives
3.5. Modeling and Optimization

Example. Page 276 number 2.

Note. Strategy for Solving Max–Min Problems

1. Draw a picture and label the variables and constants (if appropriate).

2. State the question in terms of the variables.

3. Find a relationship between the unkowns and write the desired quantity
as a function of one variable.

4. Maximize/Minimize the function.

Examples. Page 276 number 12, page 278 number 24, page 281 number
48.

1
Chapter 3. Applications of Derivatives
3.6. Linearization and Differentials

Definition. If f is differentiable at x = a, then the approximating


function
L(x) = f (a) + f  (a)(x − a)

is the linearization of f at a.

Example. Page 293 number 2.

Definition. Let y = f (x) be a differentiable function. The differential


dx is an independent variable. The differential dy is

dy = f  (x) dx.

Example. Page 294 number 24.

1
Note. Differential Estimate of Change.
Let f (x) be differentiable at x = a. The approximate change in the value
of f when x changes from a to a + dx is

df = f  (a) dx.

Section 3.6, page 294

2
Definition. We can compare actual changes in a function and the esti-
mated change which is calculated from the use of differentials. We consider
the absolute, relative, and percentage change:

True Estimated

Absolute change ∆f = f (a + dx) − f (a) df = f  (a) dx


∆f df
Relative change
f (a) f (a)
∆f df
Percentage change × 100% × 100%
f (a) f (a)

Example. Page 295 number 38.

3
Chapter 3. Applications of Derivatives
3.7. Newton’s Method

Note. In Newton’s Method, we try to approximate an x-intercept of a


graph. We use tangent lines to improve our approximations as follows:

Figure 3.7.61, page 297

1
Note. Procedure for Newton’s Method

1. Guess a first approximation to a solution of the equation f (x) = 0.

2. Use the first approximation to get a second, the second to get a third,
and so on, using the formula
f (xn)
xn+1 = xn − .
f  (xn)

Note. The equation of the tangent line to y = f (x) at (xn, f (xn )) is


y − f (xn ) = f  (xn)(x − xn). Therefore the intercept xn+1 satisfies

0 − f (xn) = f  (xn)(xn+1 − xn)


f (xn )
or xn+1 = xn −
f  (xn)
(see page 298 for the algebra steps).

2
Figure 3.7.62, page 298

Examples. Page 303 numbers 2 and 18.

3
Chapter 4. Integration
4.1. Indefinite Integrals, Differential Equations,
and Modeling

Definition. A function F (x) is an antiderivative of a function f (x) if


F (x) = f (x) for all x in the domain of f . The set of all antiderivatives of

f is the indefinite integral of f with respect to x, denoted by f (x) dx.

The symbol is an integral sign. The function f is the integrand of the
integral, and x is the variable of integration.

Note. We denote the indefinite integral SET as



f (x) dx = F (x) + C

where F is a specific antiderivative and C represents an “arbitrary con-


stant.” (In class, we will use “k” for a specific constant.)

Examples. Page 319 numbers 12 and 23.

1
Table 4.1. Integral Formulas

Indefinite Integral Derivative Formula

 
 xn+1 d  xn+1 
1. n
x dx = + C, n = −1, n rational   = x
n
n+1 dx n + 1
  d
dx = 1 dx = x + C (special case) [x] = 1
dx
 
 cos kx d  cos kx 
2. sin kx dx = − +C − = sin kx
k dx k
 
 sin kx d  sin kx 
3. cos kx dx = +C = cos kx
k dx k
 d
4. sec2 x dx = tan x + C [tan x] = sec2 x
dx
 d
5. csc2 x dx = − cot x + C [− cot x] = csc2 x
dx
 d
6. sec x tan x dx = sec x + C [sec x] = sec x tan x
dx
 d
7. csc x cot x dx = − csc x + C [− csc x] = csc x cot x
dx

Definition. The problem of finding a function y of x when we know y 


and a value of y at a particular point x0 is called an initial value problem.

Example. Page 321 numbers 54 and 58, page 320 number 36.

2
Chapter 4. Integration
4.2. Integral Rules; Integration by Substitution

Note. Rules for Indefinite Integration.


 
1. Constant Multiple Rule: kf (x) dx = k f (x) dx
 
2. Rule for Negatives: −f (x) dx = − f (x) dx
  
3. Sum and Difference Rule: [f (x) ± g(x)] dx = f (x) dx ± g(x) dx

Note. If u = u(x) is a differentiable function, then



n 

n un+1 (u(x))n+1
(u(x)) u (x) dx = u du = +C = +C
n+1 n+1
where n = −1 and n is rational.

Example. Page 328 numbers 10a.

1
Note. More generally, we have the method of u-substitution. If f and g 
are continuous then
 
f (g(x))g  (x) dx = f (u) du

where u = g(x) and du = g  (x) dx.

Examples. Page 328 number 20, pages 329 numbers 26 and 44.

2
Chapter 4. Integration
4.3. Estimating with Finite Sums

Note. In this section, we use finite sums to estimate quantities which we


will calculate precisely using integrals.

Examples. Page 334 Example 3, page 337 number 6.

Example. Page 338 number 10a.

1
Chapter 4. Integration
4.4. Riemann Sums and Definite Integrals

Note. We use the sigma notation to denote sums:


n

ak = a1 + a2 + · · · + an .
k=1

Definition. A partition of the interval [a, b] is a set

P = {x0, x1, . . . , xn} where a = x0 < x1 < · · · < xn = b.

partition P determines n closed subintervals

[x0, x1], [x1, x2], . . . , [xn−1, xn].

The length of the kth subinterval is ∆xk = xk − xk−1.

1
Note. We now estimate the area bounded between a function y = f (x)
and the x-axis. We make the convention that the area bounded above the
x-axis and below the function is positive, and the area bounded below
the x-axis and above the curve is negative. We estimate this “area” by
choosing a ck ∈ [xk−1, xk ] and we use f (ck ) as the “height” of a rectangle
with base [xk−1, xk ]. Then a partition P of [a, b] can be used to estimate
this “area” by adding up the “area” of these rectangles.

Figure 4.4.7, page 342

2
Definition. With the above notation, a Riemann sum of f on the
interval [a, b] is a sum of the form
n

sn = f (ck ) ∆xk .
k=1

Definition. The norm of a partition P = {x0, x1, . . . , xn} of interval


[a, b], denoted P , is largest subinterval:

P  = max ∆xk = max (xk − xk−1).


1≤k≤n 1≤k≤n

Note. If P  is “small,” then a Riemann sum is a “good” approximation


of the “area” described above.

Definition. Let f be a function defined on a closed interval [a, b]. For


any partition P of [a, b], let the numbers ck be chosen arbitrarily in the
subintervals [xk−1, xk ]. If there exists a number I such that
n

lim f (ck ) ∆xk = I
P →0 k=1

no matter how P and the ck ’s are chosen, then f is integrable on [a, b]


and I is the definite integral of f over [a, b]. This is denoted
 b
a
f (x) dx.
3
Note. When we deal with applications of integration, we will often think
of definite integrals as sums. Notice, however, that strictly speaking they
are not sums, but they are limits of sums.

Note. We have now introduced three ideas, each different from the other,

but each related to the other (as we will see when we state the Fundamental

Theorem of Calculus). We have:

Name of Object Type of Object


ANTIDERIVATIVE FUNCTION
INDEFINITE INTEGRAL SET
DEFINITE INTEGRAL NUMBER
Antiderivatives and indefinite integrals are related by the fact that the
indefinite integral of a function f is the set of all antiderivatives of f .
The Fundamental Theorem of Calculus, to be seen in the next section,
will relate antiderivatives and definite integrals (and therefore will relate
definite and indefinite integrals).

Example. Page 349 number 12.

4
Theorem 1. All continuous functions are integrable. That is, if a func-
tion f is continuous on an interval [a, b], then its definite integral over
[a, b] exists.

Note. If we partition [a, b] into n pieces of equal length (b − a)/n, then


the partition if regular. We can then evaluate the limit P  → 0 by
letting n → ∞. This can be used to evaluate definite integrals. If we do
so, then these formulas are useful:
 2
n
 n(n + 1) n
 n(n + 1)(2n + 1) n
 n(n + 1) 
i= , i2 = , i3 = 
  .
i=1 2 i=1 6 i=1 2

Example. Use a regular partition of [0, 1] with ck = xk to evaluate


 1
0
x2 dx.

Definition. If y = f (x) is nonnegative and integrable over a closed


interval [a, b], then the area under the curve y = f (x) from a to b is the
integral of f from a to b,
 b
A= a
f (x) dx.

Example. Page 350 number 18.


5
Definition. If f is integrable on [a, b], then its average (mean) value
of [a, b] is
1 b
av(f ) = f (x) dx.
b−a a

Example. Page 350 number 24.

Note. Rules for Definite Integrals.


 b  a
1. Order of Integration: a
f (x) dx = − b
f (x) dx (this in fact is a
definition)
 a
2. Zero: a
f (x) dx = 0 (this too is a definition)
 b  b
3. Constant Multiple: a
kf (x) dx = k a
f (x) dx
 b  b  b
4. Sum and Difference: a
(f (x) ± g(x)) dx = a
f (x) dx ± a
g(x) dx
 b  c  c
5. Additivity: a
f (x) dx + b
f (x) dx = a
f (x) dx

6. Max-Min Inequality: If max f and min f are the maximum and


minimum values of f on [a, b], then
 b
min f · (b − a) ≤ a
f (x) dx ≤ max f · (b − a).
 b  b
7. Domination: f (x) ≥ g(x) on [a, b] ⇒ a
f (x) dx ≥ a
g(x) dx.

Example. Page 350 number 28.


6
Chapter 4. Integration
4.5. The Mean Value Theorem and Fundamental Theorems

Theorem 2. The Mean Value Theorem for Definite Integrals.


If f is continuous on [a, b], then at some point c in [a, b],
1 b
f (c) = f (x) dx.
b−a a

Figure 4.5.13, page 352

1
Proof of Theorem 2. By the Max-Min Inequality from Section 4.5,
we have
1 b
min f ≤ f (x) dx ≤ max f.
b−a a
Since f is continuous, f must assume any value between min f and max f ,
1 b
including f (x) dx by the Intermediate Value Theorem. Q.E.D.
b−a a

Theorem 3. The Fundamental Theorem of Calculus, Part 1.


If f is continuous on [a, b] then the function
 x
F (x) = a
f (t) dt

has a derivative at every point x in [a, b] and


dF d  x 
= f (t) dt = f (x).
dx dx a

2
Proof. Notice that
 x+h  x  x+h
F (x + h) − F (x) = a
f (t) dt − a
f (t) dt = x
f (t) dt.

So
F (x + h) − F (x) 1 1  x+h
= [F (x + h) − F (x)] = x f (t) dt.
h h h
Since f is continuous, Theorem 2 implies that for some c ∈ [x, x + h] we
have
1  x+h
f (c) = x f (t) dt.
h
Since c ∈ [x, x + h], then lim f (c) = f (x) (since f is continuous at x).
h→0
Therefore
dF F (x + h) − F (x)
= lim
dx h→0 h

1 x+h
= lim x f (t) dt
h→0 h

= lim f (c) = f (x)


h→0

Q.E.D.

Example. Page 361 numbers 20, 18.

3
Theorem 3. The Fundamental Theorem of Calculus, Part 2.
If f is continuous at every point of [a, b] and if F is any antiderivative of
f on [a, b], then
 b
a
f (x) dx = F (b) − F (a).

Proof. We know from the first part of the Fundamental Theorem (The-
orem 3a) that
 x
G(x) = a
f (t) dt

defines an antiderivative of f . Therefore if F is any antiderivative of f ,


then F (x) = G(x) + k for some constant k. Therefore

F (b) − F (a) = [G(b) + k] − [G(a) + k]

= G(b) − G(a)
 b  a
= a
f (t) dt − a
f (t) dt
 b
= a
f (t) dt − 0
 b
= a
f (t) dt.

Q.E.D.

Examples. Page 361 numbers 8 and 26, page 362 numbers 38 and 48.

4
Chapter 4. Integration
4.6. Substitution in Definite Integrals

Note. We can use u-substitution in definite integrals:


 b  g(b)

a
f (g(x))g (x)dx = g(a)
f (u) du

where u = g(x), and du = g  (x) dx.

Examples. Page 371 numbers 10a and 14.

Definition. If f and g are continuous with f (x) ≥ g(x) throughout


[a, b], then the area of the region between the curves y = f (x) and y =
g(x) from a to b is the integral of [f − g] from a to b:
 b
A= a
[f (x) − g(x)] dx.

1
Figure 4.6.21, page 367

Note. We will take a heuristic shortcut and take “dx” slices.

Examples. Page 372 numbers 22 and 30.

2
Chapter 4. Integration
4.7. Numerical Integration
Note. If we start with a regular partition, then we can approximate
definite integrals using trapezoids instead of rectangles.

Figure 4.7.24, page 374


b−a
We let h = ∆xk = and the area of the kth trapezoid is
n
yk−1 + yk 1
(base) × (average height) = h = (yk−1 + yk )h.
2 2
So our estimate is
n
 1 h
T = (yk−1 + yk )h = (y0 + 2y1 + 2y2 + · · · + 2yn−1 + yn ).
k=1 2 2

1
 b
Definition. In the Trapezoid Rule, the integral a
f (x) dx, is approxi-
mated by
h
T = (y0 + 2y1 + 2y2 + · · · + 2yn−1 + yn ) .
2
This approximation is based on a regular partition of [a, b] where ∆xk =
h = (b − a)/n, xk = a + kh, and yk = f (xk ).

Note. We can estimate the error involved in using the Trapezoid Rule to
approximate a definite integral. If f  is continuous and M is any upper
bound for the values of |f  | on [a, b], then
 
 b


 b−a 2
|ET | = 
 a f (x) dx − T ≤

 hM
12
where h = (b − a)/n.

Note. If f (x) = mx + b then f  (x) ≡ 0 and ET = 0. So the Trapezoid


Rule gives exact values for such functions.

Examples. Page 381 number 8 I abc.

2
Note. Instead of approximating y = f (x) with straight line segments,
we can approximate it with parabolas. We then integrate to find the area
under the parabolas. This leads to Simpson’s Rule.

Figure 4.7.27, page 377

 b
Definition. In Simpson’s Rule, the integral a
f (x) dx, is approximated
by
h
S = (y0 + 4y1 + 2y2 + 4y3 + · · · + 2yn−2 + 4yn−1 + yn ).
3
This approximation is based on a regular partition of [a, b] of size n where
n is even, and where ∆xk = h = (b − a)/n, xk = a + kh, and yk = f (xk ).

3
Note. We can estimate the error involved in using Simpson’s Rule to
approximate a definite integral. If f (4) is continuous and M is any upper
bound for the values of |f (4)| on [a, b], then
 
 b


 b−a 4
|ES | = 
 a f (x) dx − S ≤

 hM
180
where h = (b − a)/n.

Note. If f is a third degree polynomial then f (4) (x) ≡ 0 and ES = 0.


So Simpson’s Rule gives exact values for such functions.

Examples. Page 381 number 8 II abc, and page 382 number 16a.

4
Chapter 5. Applications of Integrals
5.1. Volumes by Slicing and Rotation About an Axis

Definition. The volume of a solid of known integrable cross-section area


A(x) from x = a to x = b is the integral of A from a to b,
 b
V = a
A(x) dx.

Figure 5.1.1, page 393

Example. Page 402 number 2b.


1
Note. If we revolve an area about an axis, then we get cross setional
areas which are circular. We are lead to the “disk method.”

Note. Disk Method for Rotation about the x-Axis


The volume of the solid generated by revolving about the x-axis the region
between the x-axis and the graph of the continuous function y = R(x),
a ≤ x ≤ b, is
 b  b
2
V = a
π[radius] dx = a
π[R(x)]2 dx.

We can make a similar definition for x = R(y) and rotation about the
y-axis.

Figure 5.1.14, page 379 of 9th Edition

Example. Page 403 number 14.


2
Note. If the revolution does not result in disks, but results in disks with
holes in the centers, then we are lead to the “washer method.”

Note. Washer Method for Rotation about the x-Axis.


The volume of the solid generated by revolving about the x-axis the region
between y = r(x) and y = R(x) where 0 ≤ r(x) ≤ R(x) and r(x), R(x)
are continuous, for a ≤ x ≤ b is
 b  b
2 2
V = a
π[(outer radius) −(inner radius) ] dx = a
π[(R(x))2−(r(x))2] dx.

We can make a similar definition for a function of y and rotation about


the y-axis.

3
Figure 5.1.10, page 399

Example. Page 404 number 36, page 405 number 56.

4
Chapter 5. Applications of Integrals
5.2. Modeling Volume Using Cylindrical Shells

Note. If we take “dx slices” and revolve them about the y-axis, or if
we take “dy slices” and revolve them about the x-axis, then we generate
“cylindrical shells.”

Note. The Shell Formula for Revolution About the y-Axis.


The volume of the solid generated by revolving the region between the
x-axis and the graph of the continuous function y = f (x) ≥ 0, 0 ≤ a ≤
x ≤ b, about the y-axis is
 b  b
V = a
2π (shell radius) (shell height) dx = a
2πxf (x) dx.

We can make a similar definition for functions of y and rotation about the
x-axis.

1
Figure 5.2.27, page 388 of 9th Edition

Figure 5.2.19, page 408


2
Example. Page 413 number 28a, page 411 number 6.

Note. If we desire to revolve about a horizontal or vertical line other


than an axis, then we only need modify the radius term.

Example. Page 413 numbers 28c and 28d.

3
Chapter 5. Applications of Integrals
5.3. Lengths of Plane Curves

Definition. Function f is smooth if it’s derivative in continuous. If f


is smooth on [a, b], the length of the curve y = f (x) from a to b is the
number 

  2
 b  dy

L= a
 1 +   dx.
dx
If g is smooth on [c, d], the length of the curve x = g(y) from c to d is
the number 

  2
 d  dx

L= c
 1 +   dy.
dy

Figure 5.3.27, page 415


1
Figure 5.3.29, page 418

Example. Page 420 number 4.

Definition. If a curve C is described by the parametric equations


x = f (t), y = g(t), α ≤ t ≤ β, where f  and g  are continuous and
not simultaneously zero on [α, β] and if C is traverses exactly once as t
increases from α to β, then the length of C is

   
 β


 dx 2  dy 2
L= α
 + dt.
dt dt

Examples. Page 420 number 12, page 421 number 28.

Das könnte Ihnen auch gefallen