4 Integration
The theory of infinite series addresses how to sum infinitely many finite quantities. Integration, by
contrast, is the business of summing infinitely many infinitesimal quantities. Attempts to do both have
been part of mathematics for well over 2000 years, and the philosophical objections are just as old.
17
The development and increased application of calculus from the late 1600s onward spurred mathe-
maticians to put the theory on a firmer footing, though from Newton and Leibniz it took another 150
years before Bernhard Riemann (1856) provided a thorough development of the integral.
4.32 The Riemann Integral
The basic idea behind Riemann integration is to approximate area using a sequence of rectangles
whose width tends to zero. The following discussion illustrates the essential idea, which should be
familiar from elementary calculus.
Example 4.1. Suppose f (x) = x
2
is defined on [0, 1].
For each n N, let x =
1
n
and define x
i
= ix.
Above each subinterval [x
i1
, x
i
], raise a rectangle of height
f (x
i
) = x
2
i
. The sum of the areas of these rectangles is the Rie-
mann sum with right-endpoints
18
R
n
=
n
i=1
f (x
i
)x =
n
i=1
i
2
n
3
=
n(n + 1)(2n + 1)
6n
3
=
1
3
+
3n + 1
6n
2
The Riemann sum with left-endpoints is defined similarly:
L
n
=
n
i=1
f (x
i1
)x =
n
i=1
(i 1)
2
n
3
=
1
3
3n 1
6n
2
Since f is an increasing function, the area A under the curve
plainly satisfies
L
n
A R
n
By the squeeze theorem, we conclude that A =
1
3
.
0
1
0 1
n =
16
R
n
=
0.365234
0
1
0 1
n =
16
L
n
=
0.302734
The example should feel convincing, though perhaps this is due to the simplicity of the function. To
apply this approach to more general functions, we need to be significantly more rigorous.
17
Two of Zeno’s ancient paradoxes are relevant here: Achilles and the Tortoise concerns a convergent infinite series,
while the Arrow Paradox toys with integration by questioning whether time can be viewed as a sum of instants. Perhaps
the most famous contemporary criticism comes from Bishop George Berkeley, who gave his name to the city and first
UC campus: in 1734’s The Analyst, Berkeley savaged the foundations of calculus, describing the infinitesimal increments
required in Newton’s theory of fluxions (derivatives) as merely the “ghosts of departed quantities.”
18
Recall some basic identities:
n
i=1
i =
1
2
n(n + 1),
n
i=1
i
2
=
1
6
n(n + 1)(2n + 1),
n
i=1
i
3
=
1
4
n
2
(n + 1)
2
70
Definition 4.2. A partition P = {x
0
, . . . , x
n
} of an interval [a, b] is a finite sequence for which
a = x
0
< x
1
< ··· < x
n1
< x
n
= b
Choosing a sample point x
i
in each subinterval [x
i1
, x
i
] results in a tagged partition.
The mesh of the partition is mesh(P) := max x
i
, the width x
i
= x
i
x
i1
of the largest subinterval.
If f : [a, b] R, the Riemann sum
n
i=1
f (x
i
) x
i
evaluates the area of a family of n rectangles, as
pictured. The heights f (x
i
) and thus areas can be negative or zero.
f (x)
x
x
1
x
1
x
2
x
2
x
3
x
3
x
4
x
4
x
5
x
5
x
6
x
6
b = x
7
x
7
a = x
0
b = x
7
In elementary calculus, one typically computes Riemann sums for equally-spaced partitions with left,
right or middle sample points. The flexibility of tagged partitions makes applying Riemann’s defini-
tion a challenge, so we instead consider two special families of rectangles.
Definition 4.3. Given a partition P of [a, b] and a bounded function f on [a, b], define
M
i
= sup
x[x
i1
,x
i
]
f (x) U( f , P) =
n
i=1
M
i
x
i
m
i
= inf
x[x
i1
,x
i
]
f (x) L( f , P) =
n
i=1
m
i
x
i
U( f , P) and L( f , P) are the upper and lower Darboux sums for
f with respect to P. The upper and lower Darboux integrals are
U( f ) = inf U( f , P) L( f ) = sup L( f , P)
where the supremum/infimum are taken over all partitions.
Necessarily both integrals are finite.
We say that f is (Riemann) integrable on [a, b] if U( f ) = L( f ).
We denote this value by
Z
b
a
f or
Z
b
a
f (x) dx
a x
1
x
2
x
3
x
4
x
5
x
6
x
7
b
Upper Darboux sum U( f , P)
a x
1
x
2
x
3
x
4
x
5
x
6
x
7
b
Lower Darboux sum L( f , P)
If the interval is understood or irrelevant, one often simply says that f is integrable and writes
R
f .
Intuitively, L( f , P) is the sum of the areas of rectangles built on P which just fit under the graph of f .
It is also the infimum of all Riemann sums on P. If f is discontinuous, then L( f , P) need not itself be
a Riemann sum, as there might not exist suitable sample points!
71
Examples 4.4. 1. We revisit Example 4.1 in this language.
Given a partition Q = {x
0
, . . . , x
n
} of [0, 1] and sample points x
i
[x
i1
, x
i
], we compute the
Riemann sum for f (x) = x
2
n
i=1
f (x
i
) x
i
=
n
i=1
(x
i
)
2
(x
i
x
i1
)
Since f is increasing, we have x
2
i1
(x
i
)
2
x
2
i
on each interval, whence
L( f , Q) =
n
i=1
(x
i1
)
2
(x
i
x
i1
)
n
i=1
(x
i
)
2
(x
i
x
i1
)
n
i=1
(x
i
)
2
(x
i
x
i1
) = U( f , Q)
The Darboux sums are therefore the Riemann sums for left- and right-endpoints.
If we take Q
n
to be the partition with subintervals of equal width x =
1
n
, then
U( f ) = inf
P
U( f , P) U( f , Q
n
) =
n
i=1
i
n
2
x = R
n
is the right Riemann sum discussed originally. Similarly L( f ) L
n
. Since L
n
and R
n
both
converge to
1
3
as n , the squeeze theorem forces
L
n
L( f ) U( f ) R
n
= L( f ) = U( f ) =
1
3
Otherwise said, f is integrable on [0, 1] with
R
1
0
x
2
dx =
1
3
.
2. Suppose f (x) = kx + c on [a, b], and that k > 0. Take
the evenly spaced partition P
n
where x
i
= a +
ba
n
i.
Since f is increasing, the upper Darboux sum is again
the Riemann sum with right-endpoints:
U( f , P
n
) = R
n
=
n
i=1
f (x
i
)x
=
b a
n
n
i=1
k(b a)
n
i + ak + c
0
ba
c
ak + c
bk + c
U( f , P
n
)
=
b a
n
k(b a)
n
·
1
2
n(n + 1) + (ak + c)n
n
1
2
k(b a)
2
+ (b a)(ak + c) =
k
2
(b
2
a
2
) + c(b a)
Similarly, the lower Darboux sum is the Riemann sum with left-endpoints:
L( f , P
n
) = L
n
=
b a
n
k(b a)
n
·
1
2
n(n 1) + (ak + c)n
n
k
2
(b
2
a
2
) + c(b a)
As above, L
n
L( f ) U( f ) R
n
and the squeeze theorem prove that f is integrable on [a, b]
with
R
b
a
f =
k
2
(b
2
a
2
) + c(b a).
72
Now we have some examples, a few remarks are in order.
Riemann versus Darboux Definition 4.3 is really that of the Darboux integral. Here is Riemann’s defi-
nition: f : [a, b] R being integrable with integral
R
b
a
f means
ϵ > 0, δ such that (P, x
i
) mesh(P) < δ =
n
i=1
f (x
i
)x
i
Z
b
a
f
< ϵ
This is significantly more difficult to work with, though it can be shown to be equivalent to the
Darboux integral. We won’t pursue Riemann’s formulation further, except to observe that if
a function is integrable and mesh(P
n
) 0, then
R
b
a
f = lim
n
n
i=1
f (x
i
)x
i
: this allows us to
approximate integrals using any sample points we choose, hence why right-endpoints (x
i
= x
i
)
are so common in Freshman calculus.
Monotone Functions Darboux sums are easy to compute for monotone functions. As in the examples,
if f is increasing, then each M
i
= f (x
i
), from which U( f , P) is the Riemann sum with right-
endpoints. Similarly, L( f , P) is the Riemann sum with left-endpoints.
Area If f is positive and continuous,
19
the Riemann integral
R
b
a
f serves as a definition for the area
under the curve y = f (x). This should make intuitive sense:
1. In the second example where we have a straight line, we obtain the same value for the
area by computing directly as the sum of a rectangle and a triangle!
2. For any partition P, the area under the curve should satisfy the inequalities
L( f , P) Area U( f , P)
But these are precisely the same inequalities satisfied by the integral itself!
L( f , P) L( f ) =
Z
b
a
f = U( f ) U( f , P)
In the examples we exhibited a sequence of partitions (P
n
) where U( f , P
n
) and L( f , P
n
) converged to
the same limit. The remaining results in this section develop some basic properties of partitions and
make this limiting process rigorous.
Definition 4.5. If P Q are both partitions of [a, b], we call Q a refinement of P.
To refine a partition, we simply throw some more points in!
Lemma 4.6. Suppose f : [a, b] R is bounded.
1. If Q is a refinement of P (on [a, b]), then
L( f , P) L( f , Q) U( f , Q) U( f , P)
2. For any partitions P, Q of [a, b], we have L( f , P) U( f , Q).
3. L( f ) U( f )
19
We’ll see in Theorem 4.17 that every continuous function is integrable.
73
Proof. 1. We prove inductively. Suppose first that Q = P {t} contains exactly one additional point
t (x
k1
, x
k
). Write
m
1
= inf
f (x) : x [x
k1
, t]
m
2
= inf
f (x) : x [t, x
k1
]
m = inf
f (x) : x [x
k1
, x
k
]
= min{m
1
, m
2
}
The Darboux sums L( f , P) and L( f , Q) are identical ex-
cept for the terms involving t. This results in extra area:
x
k1
x
k
t
Extra area!
··· ···
m
1
m
2
L( f , Q) L( f , P) = m
1
(t x
k1
) + m
2
(x
k
t) m(x
k
x
k1
)
= (m
1
m)(t x
k1
) + (m
2
m)(x
k
t) 0
More generally, since a refinement Q is obtained by adding finitely many new points, induction
tells us that P Q = L( f , P) L( f , Q).
The argument for U( f , Q) U( f , P) is similar, and the middle inequality is trivial.
2. If P and Q are partitions, then P Q is a refinement of both P and Q. By part 1,
L( f , P) L( f , P Q) U( f , P Q) U( f , Q) ()
3. This is an exercise.
Theorem 4.7. Suppose f : [a, b] R is bounded.
1. (Cauchy criterion) f is integrable ϵ > 0, P such that U( f , P) L( f , P) < ϵ.
2. f is integrable (P
n
)
nN
such that U( f , P
n
) L( f , P
n
) 0. In such a situation, both
sequences U( f , P
n
) and L( f , P
n
) converge to
R
b
a
f .
Part 1 is termed a ‘Cauchy’ criterion since it doesn’t mention the integral (limit).
Proof. We prove the Cauchy criterion, leaving part 2 as an exercise.
() Suppose f is integrable and that ϵ > 0 is given. Since inf U( f , Q) =
R
f = sup L( f , R), there
exist partitions Q, R such that
U( f , Q) <
Z
f +
ϵ
2
and L( f , R) >
Z
f
ϵ
2
Let P = Q R and apply (): L( f , R) L( f , P) U( f , P) U( f , Q). But then
U( f , P) L( f , P) U( f , Q) L( f , R) = U( f , Q)
Z
f +
Z
f L( f , R) < ϵ
() Assume the right hand side. For every partition, L( f , P) L( f ) U( f ) U( f , P). Thus
0 U( f ) L( f ) U( f , P) L( f , P) < ϵ
Since this holds for all ϵ > 0, we see that U( f ) = L( f ): that is, f is integrable.
74
Examples 4.8. 1. Consider f (x) =
x on the interval [0, b]. We choose a sequence of partitions (P
n
)
that evaluate nicely when fed to this function:
P
n
= {x
0
, . . . , x
n
} where x
i
=
i
n
2
b
= x
i
= x
i
x
i1
=
b
n
2
i
2
(i 1)
2
=
(2i 1)b
n
2
Since f is increasing on [0, b], we see that
U( f , P
n
) =
n
i=1
f (x
i
)x
i
=
n
i=1
i
b
n
·
(2i 1)b
n
2
=
b
3/2
n
3
n
i=1
2i
2
i
=
b
3/2
n
3
1
3
n(n + 1)(2n + 1)
1
2
n(n + 1)
n
2
3
b
3/2
Similarly
L( f , P
n
) =
n
i=1
f (x
i1
)x
i
=
n
i=1
(i 1)
b
n
·
(2i 1)b
n
2
=
b
3/2
n
3
n
i=1
2i
2
3i + 1
=
b
3/2
n
3
1
3
n(n + 1)(2n + 1)
3
2
n(n + 1) + n
n
2
3
b
3/2
Since the limits are equal, we conclude that f is integrable and
R
b
0
x dx =
2
3
b
3/2
.
0
0 b
b
Upper Sum U( f , P
n
)
0
0 b
b
Lower Sum L( f , P
n
)
2. Here is the classic example of a non-integrable function. Let f : [a, b] R to be the indicator
function of the irrational numbers,
f (x) =
(
1 if x Q
0 if x Q
Suppose P = {x
0
, . . . , x
n
} is any partition of [a, b]. Since any interval of positive length contains
both rational and irrational numbers, we see that
sup
f (x) : x [x
i1
, x
i
]
= 1 = U( f , P) =
n
i=1
(x
i
x
i1
) = b a = U( f ) = b a
inf
f (x) : x [x
i1
, x
i
]
= 0 = L( f , P) = 0 = L( f ) = 0
Since the upper and lower Darboux integrals differ, f is not (Riemann) integrable.
75
As any freshman calculus student can attest, if you can find an anti-derivative, then the fundamen-
tal theorem of calculus (Section 4.34) makes evaluating integrals far easier. For instance, you are
probably desperate to write
d
dx
2
3
x
3/2
= x
1/2
=
Z
b
0
x dx =
2
3
x
3/2
b
0
=
2
3
b
3/2
rather than computing Riemann/Darboux sums as in the previous example! However, in most prac-
tical situations, no easy-to-compute anti-derivative exists; the best we can do is to approximate using
Riemann sums for progressively finer partitions. Thankfully computers excel at such tedious work!
Exercises 4.32. Key concepts: Darboux sums/integrals, Partitions, sample points & refinements,
Cauchy & sequential criteria for integrability
1. Use partitions to find the upper and lower Darboux integrals on the interval [0, b] . Hence prove
that the function is integrable and compute its integral.
(a) f (x) = x
3
(b) g(x) =
3
x
2. Repeat question 1 for the following two functions. You cannot simply compute Riemann sums
for left and right endpoints and take limits: why not?
(a) h(x) = x(2 x) on [0, 2]
(Hint: choose a partition with 2n subintervals such that x
n
= 1 and observe that h(2 x) = h(x))
(b) On the interval [0, 3], let k(x) =
(
2x if x 1
5 x if x > 1
(Hint: this time try a partition with 3n subintervals)
3. Let f (x) = x for rational x and f (x) = 0 for irrational x. Calculate the upper and lower
Darboux integrals for f on the interval [0, b]. Is f integrable on [0, b]?
4. Prove part 3 of Lemma 4.6: L( f ) U( f ).
5. Prove part 2 of Theorem 4.7.
f is integrable (P
n
)
nN
such that lim
n
U( f , P
n
) L( f , P
n
)
= 0
Moreover, prove that both U( f , P
n
) and L( f , P
n
) converge to
R
f .
6. (a) Reread Definition 4.3. What happens if we allow f : [a, b] R to be unbounded?
(b) (Hard) Read Riemann versus Darboux on page 73. Explain why being Riemann integrable
also forces f to be bounded.
(c) (Hard) Explain the observation that L( f , P) is the infimum of the set of all Riemann sums
on P.
7. (If you like coding) Write a short program to estimate
R
b
a
f (x) dx using Riemann sums. This
can be very simple (equal partitions with right endpoints), or more complex (random partition
and sample points given a mesh). Apply your program to estimate
R
5
0
sin(x
2
e
x
) dx.
76
4.33 Properties of the Riemann Integral
The rough take-away of this long section is that everything you think is integrable probably is! Ex-
amples will be few, since we have not established many explicit values for integrals.
Theorem 4.9 (Linearity). If f , g are integrable and k, l are constant, then k f + lg is integrable and
Z
k f + lg = k
Z
f + l
Z
g
Example 4.10. Thanks to examples in the previous section, we can now calculate, e.g.,
Z
2
0
5x
3
3
x dx = 5 ·
1
4
·2
4
3 ·
2
3
·2
3/2
= 20 4
2
Proof. Suppose ϵ > 0 is given. By the Cauchy criterion (Theorem 4.7, part 1), there exist partitions
R, S such that
U( f , R) L( f , R) <
ϵ
2
and U(g, S) L(g, S) <
ϵ
2
If P = R S, then both inequalities are satisfied by P (Lemma 4.6). On each subinterval,
inf f (x) + inf g(x) inf
f (x) + g(x)
and sup
f (x) + g(x)
sup f (x) + sup g(x)
since the individual suprema/infima could be ‘evaluated’ at different places. Thus
L( f , P) + L(g, P) L( f + g, P) U( f + g, P) U( f , P) + U(g, P)
whence U( f + g, P) L( f + g, P) < ϵ and f + g is integrable. Moreover,
Z
( f + g)
Z
f
Z
g
U( f , P)
Z
f
+
U(g, P)
Z
g
< ϵ
Using lower Darboux integrals similarly obtains the other half of the inequality
ϵ <
Z
( f + g)
Z
f
Z
g < ϵ
Since this holds for all ϵ > 0, we conclude that
R
( f + g) =
R
f +
R
g.
That k f is integrable with
R
k f = k
R
f is an exercise. Put these together for the result.
Corollary 4.11 (Changing endvalues). Suppose f is integrable on [a, b] and g : [a, b] R satisfies
f (x) = g(x) on (a, b). Then g is also integrable on [a, b] and
R
b
a
g =
R
b
a
f .
Definition 4.12 (Integration on an open interval). A bounded function g : (a, b) R is integrable if
it has an integrable extension f : [a, b] R where f (x) = g(x) on (a, b). In such a case, we define
R
b
a
g :=
R
b
a
f .
The Corollary (its proof is an exercise) shows that the choice of extension is irrelevant.
77
Theorem 4.13 (Basic integral comparisons). Suppose f and g are integrable on [a, b]. Then:
1. f (x) g(x) =
R
f
R
g
2. m f (x) M = m(b a)
R
b
a
f M(b a)
3. f g is integrable.
4.
|
f
|
is integrable and
R
f
R
|
f
|
5. max( f , g) and min( f , g) are both integrable.
Part 3 is not integration by parts since it doesn’t tell us how
R
f g relates to
R
f and
R
g!
Proof. 1. Since g f is positive and integrable, L(g f , P) 0 for all partitions P. But then
0 inf L(g f , P) = L(g f ) =
Z
g f =
Z
g
Z
f
2. Apply part 1 twice.
3. This is an exercise.
4. The integrability is an exercise. For the comparison, apply part 1 to
|
f
|
f
|
f
|
.
5. Use max( f , g) =
1
2
( f + g) +
1
2
|
f g
|
, etc., together with the previous parts.
Theorem 4.14 (Domain splitting). Suppose f : [a, b] R and
let c (a, b). If f is integrable on both [a, c] and [c, b], then it is
integrable on [a, b] and
Z
b
a
f =
Z
c
a
f +
Z
b
c
f
f (x)
a c b x
R
c
a
f
R
b
c
f
In light of this result, it is conventional to allow integral limits to be reversed: if a < b, then
Z
a
b
f :=
Z
b
a
f is consistent with
Z
a
a
f = 0
Proof. Let ϵ > 0 be given, then R, S partitions of [a, c], [c, b] such that
U( f , R) L( f , R) <
ϵ
2
, U( f , S) L( f , S) <
ϵ
2
Choose P = R S to partition [a, b], then
U( f , P) L( f , P) = U( f , R) + U( f , S) L( f , R) L( f , S) < ϵ
Moreover
f (x)
a c b x
R
z
}| {
S
z
}| {
Z
b
a
f
Z
c
a
f
Z
b
c
f U( f , P) L( f , R) L( f , S) = U( f , P) L( f , P) < ϵ
Showing that this expression is greater than ϵ is similar.
78
Example 4.15. If f (x) =
x on [0, 1] and f (x) = 1 on [1, 2], then
Z
2
0
f =
Z
1
0
x dx +
Z
2
1
1 dx =
2
3
+ 1 =
5
3
Monotonic & Continuous Functions We establish the integrability of two large classes of functions.
Definition 4.16. A function f : [a, b] R is:
Monotonic if it is either increasing (x < y = f (x) f (y)) or decreasing.
Piecewise monotonic if there is a partition P = {x
0
, . . . , x
n
} (finite!) of [a, b] such that f is monotonic
on each open subinterval (x
k1
, x
k
).
Piecewise continuous if there is a partition such that f is uniformly continuous on each (x
k1
, x
k
).
Theorem 4.17. If f is monotonic or continuous on [a, b], then it is integrable.
Examples 4.18. 1. Since sine is continuous, we can approximate via a sequence of Riemann sums
Z
π
0
sin x dx =
π
n
lim
n
n
i=1
sin
πi
n
Evaluating this limit is another matter entirely, one best handled in the next section...
2. Similarly, e
x
is integrable and therefore may be approximated via Riemann sums:
Z
1
0
e
x
dx =
1
n
lim
n
n
i=1
exp
r
i
n
= lim
n
n
j=1
2j 1
n
exp
j
n
Both sums use right endpoints: the first has equal subintervals, while the second is analogous
to Example 4.8.1. These limits would typically be estimated using a computer.
Proof. Since [a, b] is closed and bounded, a continuous function f is uniformly so. Let ϵ > 0 be given:
δ > 0 such that x, y [a, b],
|
x y
|
< δ =
|
f (x) f (y)
|
<
ϵ
b a
Let P be a partition with mesh P < δ. Since f attains its bounds on each [x
i1
, x
i
],
x
i
, y
i
[x
i1
, x
i
] such that M
i
m
i
= f (x
i
) f (y
i
) <
ϵ
b a
from which
U( f , P) L( f , P) <
n
i=1
ϵ
b a
(x
i
x
i1
) = ϵ
The monotonicity argument is an exercise.
Combining the proof with Definition 4.12: every uniformly continuous f : (a, b) R is integrable.
79
Corollary 4.19. Piecewise continuous and bounded piecewise monotonic functions are integrable.
Proof. If f is piecewise continuous, then the restriction of f to (x
k1
, x
k
) has a continuous extension
g
k
: [x
k1
, x
k
] R; this is integrable by Theorem 4.17. By Corollary 4.11, f is integrable on [x
k1
, x
k
]
with
R
x
k
x
k1
f =
R
x
k
x
k1
g
k
. Theorem 4.14 (n 1 times!) finishes things off:
Z
b
a
f =
n
k=1
Z
x
k
x
k1
f
The argument for piecewise monotonicity is similar.
Example 4.20. The ‘fractional part’ function f (x) = x x
is both piecewise continuous and piecewise monotone on any
bounded interval. It is therefore integrable on any such interval.
0
1
0 1 2 3 4 5
For a final corollary, here is one more incarnation of the intermediate value theorem.
Corollary 4.21 (IVT for integrals). If f is continuous on [a, b], then ξ (a, b) for which
f (ξ) =
1
b a
Z
b
a
f
Proof. Since f is continuous, it is integrable on [a, b]. By the extreme value theorem it is also bounded
and attains its bounds: p, q [a, b] such that
f (p) := inf
x[a,b]
f (x), f (q) = sup
x[a,b]
f (x)
Applying Theorem 4.13, part 2, with m = f (p) and
M = f (q), we see that
(b a) f (p)
Z
b
a
f (b a) f (q)
f
av
M
m
ξa bp q
Divide by b a and apply the usual intermediate value theorem for f to see that the required ξ exists
between p and q.
In the picture, when f is positive and continuous, the grey area equals that under the curve; imagine
levelling off the blue hill with a bulldozer. . . The notation f
av
=
1
ba
R
b
a
f indicates the average value
of f on [a, b]: to see why this interpretation is sensible, take a sequence of Riemann sums on equally-
spaced partitions P
n
to see that
1
b a
Z
b
a
f = lim
n
n
i=1
f (x
i
)x = lim
n
f (x
1
) + ···+ f (x
n
)
n
is the limit of a sequence of averages of equally-spaced samples f (x
i
).
80
What can/cannot be integrated?
We now know a great many examples of integrable functions:
Piecewise continuous & monotonic functions are integrable.
Linear combinations, products, absolute values, maximums and minimums of (already) inte-
grable functions.
By contrast, we’ve only seen one non-integrable function (Example 4.8.2). After so many positive
integrability conditions, it is reasonable to ask precisely which functions are Riemann integrable.
Here is the answer, though it is quite tricky to understand.
Theorem 4.22 (Lebesgue). Suppose f : [a, b] R is bounded. Then
f is Riemann integrable it is continuous except on a set of measure zero
Na
¨
ıvely, the measure of a set is the sum of the lengths of its maximal subintervals, though unfortu-
nately this doesn’t make for a very useful definition.
20
Any countable subset has measure zero, so
Lebesgue’s result is almost as if we can extend Corollary 4.19 to allow for infinite sums. For instance,
Exercise 1.17.8 describes a function which is continuous only on the irrationals: it is thus Riemann
integrable (indeed
R
b
a
f = 0 for any a < b). There are also uncountable sets with measure zero such
as Cantor’s middle-third set C: the function
f (x) =
(
1 if x C
0 otherwise
is continuous except on C and therefore Riemann integrable; again
R
1
0
f (x) dx = 0.
Exercises 4.33. Key concepts: Linear combinations, products, etc., of integrable functions are integrable,
Continuous and monotone functions are integrable, Integrability on open intervals
1. Explain why
R
2π
0
x
2
sin
8
(e
x
) dx
8
3
π
3
2. If f is integrable on [a, b] prove that it is integrable on any interval [c, d] [a, b].
3. We complete the proof of Theorem 4.9 (linearity of integration).
(a) Suppose k > 0, let A R and define kA := {kx : x A}. Prove that sup kA = k sup A
and inf kA = k inf A.
(b) If k > 0 prove that k f is integrable on any interval and that
R
k f = k
R
f .
(c) How should you modify your argument if k < 0?
20
Formally, the length of an open interval (a, b) is b a and a set A R has measure zero if
ϵ > 0, open intervals I
n
such that A
[
n=1
I
n
and
i=1
length(I
n
) < ϵ
More generally, the Lebesgue measure of a set (subject to a technical condition) is the infimum of the sum of the lengths of
any countable collection of open covering intervals. Measure theory is properly a matter for graduate study. Surprisingly,
there exist sets with positive measure that contain no subintervals, and even sets which are non-measurable!
81
4. Give an example of an integrable but discontinuous function on a closed bounded interval [a, b]
for which the conclusion of the Intermediate Value Theorem for Integrals is false.
5. Use Darboux sums to compute the value of the integral
R
15/2
1/2
x xdx (Example 4.20).
6. We prove and extend Corollary 4.11. Suppose f is integrable on [a, b].
(a) If g : [a, b] R satisfies f (x) = g(x) for all x (a, b), prove that g is integrable and
R
b
a
g =
R
b
a
f .
(Hint: consider h = f g and show that
R
h = 0)
(b) Now suppose g : [a, b] R satisfies f (x) = g(x) for all x [a, b] except at finitely many
points. Prove that g is integrable and
R
b
a
g =
R
b
a
f .
7. Show that an increasing function on [a, b] is integrable and thus complete Theorem 4.17.
(Hint: Choose a partition with mesh P <
ϵ
f (b)f (a)
)
8. Suppose f and g are integrable on [a, b].
(a) Define h(x) =
f (x)
2
. We know:
f is bounded: K such that
|
f (x)
|
K on [a, b] .
Given ϵ > 0, P such that U( f , P) L( f , P) <
ϵ
2K
. For each subinterval [x
i1
, x
i
], let
M
i
= sup f (x), m
i
= inf f (x), M
i
= sup h(x), m
i
= inf h(x)
Prove that M
i
m
i
2(M
i
m
i
)K. Hence conclude that h is integrable.
(b) Prove that f g is integrable.
(Hint: f g =
1
4
( f + g)
2
1
4
( f g)
2
)
(c) Prove that U(
|
f
|
, P) L(
|
f
|
, P) U( f , P) L( f , P) for any partition P. Hence conclude
that
|
f
|
is integrable.
(One can extend these arguments to show that if j is continuous, then j f is integrable. Parts (a) and
(c) correspond, respectively, to j(x) = x
2
and j(x) =
|
x
|
.)
9. (Hard) Let f (x) =
x if x = 0 and sin
1
x
> 0
x if x = 0 and sin
1
x
< 0
0 if x = 0
(a) Show that f is not piecewise continuous on [0, 1].
(b) Show that f is not piecewise monotonic on [0, 1].
(c) Show that f is integrable on [0, 1].
(Hint: given ϵ, hunt for a suitable partition to make U( f , P) L( f , P) < ϵ by considering [0, x
1
]
differently to the other subintervals)
(d) Make a similar argument which proves that g = sin
1
x
is integrable on ( 0, 1].
(Hint: Show that g has an integrable extension on [0, 1])
82
4.34 The Fundamental Theorem of Calculus
The key result linking integration and differentiation is usually presented in two parts. While there
are significant subtleties, the rough statements are as follows (we follow the traditional numbering):
Part I Differentiation reverses integration:
d
dx
R
x
a
f (t) dt = f (x)
Part II Integration reverses differentiation:
R
b
a
F
(x) dx = F(b) F(a)
These facts seemed intuitively obvious to early practitioners of calcu-
lus. Given a continuous positive function f :
Let F(x) denote the area under y = f (x) between 0 and x.
A small increase x results in the area increasing by F.
F f (x)x is approximately the area of a rectangle, whence
F
x
f (x). This is part I.
F(b) F(a)
F
i
f (x
i
)x
i
. Since F
= f , this is part II.
When Leibniz introduced the symbols
R
and d in the late 1600s, it was partly to reflect the fundamen-
tal theorem.
21
If you’re happy with non-rigorous notions of limit, rate of change, area, and (infinite)
sums, the above is all you need!
Of course we are very much concerned with the details: What must we assume about f and F, and
how are these properties used in the proof?
Theorem 4.23 (FTC, part I). Suppose f is integrable on [a, b]. For any x [a, b], define
F(x) :=
Z
x
a
f (t) dt
Then:
1. F is uniformly continuous on [a, b];
2. If f is continuous at c [a, b], then F is differentiable
22
at c with F
(c) = f (c).
Compare this with the na
¨
ıve version above where we assumed f was continuous. We now require
only the integrability of f , and its continuity at one point for the full result.
21
R
is a stylized S for sum, while d stands for difference. Given a sequence F = (F
0
, F
1
, F
2
, . . . , F
n
), construct a new
sequence of differences
dF = (F
1
F
0
, F
2
F
1
, . . . , F
n
F
n1
)
which can then be summed:
Z
dF = (F
1
F
0
) + (F
2
F
1
) + ···(F
n
F
n1
) = F
n
F
0
()
Viewing a function as an ‘infinite sequence’ of values spaced along an interval, dF becomes a sequence of infinitesimals and
() is essentially the fundamental theorem:
R
dF = F(b) F(a). It is the concept of function that is suspect here, not the
essential relationship between sums and differences.
22
Strictly: if c = a, then F is right-differentiable, etc.
83
Examples 4.24. Examples in every elementary calculus course.
1. Since f (x) = sin
2
(x
3
7) is continuous on any bounded interval, we conclude that
d
dx
Z
x
4
sin
2
(t
3
7) dt = sin
2
(x
3
7)
If one follows Theorem 4.14 and its conventions, then this is valid for all x R.
2. The chain rule permits more complicated examples. For instance: f (t) = sin
t is continuous
on its domain [0, ) and y(x) = x
2
+ 3 has range [ 3, ) dom( f ), whence
d
dx
Z
x
2
+3
0
sin
t dt =
dy
dx
d
dy
Z
y
0
sin
t dt = 2x sin
p
x
2
+ 3
3. For a final positive example, we consider when
d
dx
Z
e
x
sin x
tan(t
2
) dt = e
x
tan(e
2x
) cos x tan(sin
2
x)
Makes sense. To evaluate this, first choose any constant a and write
Z
e
x
sin x
=
Z
e
x
a
+
Z
a
sin x
=
Z
e
x
a
Z
sin x
a
before differentiating. This is valid provided sin x, e
x
and a all lie in the same subinterval of
dom tan(t
2
) = R \ {±
q
π
2
, ±
q
3π
2
, ±
q
5π
2
, . . .}
Since
|
sin x
|
1 <
p
π
2
, this requires
e
2x
<
π
2
x <
1
2
ln
π
2
Choosing a = 1 would certainly suffice.
4. Now consider why the theorem requires continuity. The piecewise
continuous function
f : [0, 2] R : x 7
(
2x if x 1
1
2
if x > 1
has a jump discontinuity at x = 1. We can still compute
F(x) =
(
R
x
0
2t dt = x
2
if x 1
R
1
0
2t dt +
R
x
1
1
2
dt =
1
2
(x + 1) if x > 1
This is continuous, indeed uniformly so! However the discontinu-
ity of f results in F having a corner and thus being non-differentiable
at x = 1. Indeed F
(x) = f (x) whenever x = 1: that is, at all values
of x where f is continuous.
0
1
2
f (x)
0 1 2
x
0
1
F(x)
0 1 2
x
84
Proving FTC I Neither half of the theorem is particularly difficult once you write down what you
know and what you need to prove. Here are the key ingredients:
1. Uniform continuity for F means we must control the size of
|
F(y) F(x)
|
=
Z
y
a
f (t) dt
Z
x
a
f (t) dt
=
Z
y
x
f (t) dt
Z
y
x
|
f (t)
|
dt
But the boundedness of f allows us to control this last integral. . .
2. F
(c) = f (c) means showing that lim
xc
F(x)F(c)
xc
= f (c), which means controlling the size of
F(x) F(c)
x c
f (c)
=
1
x c
Z
x
c
f (t) dt f (c)
The trick here will is to bring the constant f (c) inside the integral as
1
xc
R
x
c
f (c) dt so that the
above becomes
1
|
xc
|
R
x
c
|
f (t) f (c)
|
dt. This may now be controlled via the continuity of f . . .
Proof. 1. Since f is integrable, it is bounded: M > 0 such that
|
f (x)
|
M for all x.
Let ϵ > 0 be given and define δ =
ϵ
M
. Then, for any x, y [a, b],
0 < y x < δ =
|
F(y) F(x)
|
=
Z
y
x
f (t) dt
Z
y
x
|
f (t)
|
dt (Theorem 4.13, part 4)
M(y x) (Theorem 4.13, part 2)
< Mδ = ϵ
We conclude that F is uniformly continuous on [a, b].
2. Let ϵ > 0 be given. Since f is continuous at c, δ > 0 such that, for all t [a, b],
|
t c
|
< δ =
|
f (t) f (c)
|
<
ϵ
2
Now for all x [a, b] (except c),
0 <
|
x c
|
< δ =
F(x) F(c)
x c
f (c)
=
1
x c
Z
x
c
f (t) f (c) dt
(Theorem 4.9)
1
|
x c
|
Z
x
c
|
f (t) f (c)
|
dt (Theorem 4.13)
1
|
x c
|
ϵ
2
|
x c
|
=
ϵ
2
< ϵ
Clearly lim
xc
F(x)F(c)
xc
= f (c). Otherwise said, F is differentiable at c with F
(c) = f (c).
85
The Fundamental Theorem, part II As with part I, the formulaic part of the result should be familiar,
though we are more interested in the assumptions and where they are needed.
Theorem 4.25 (FTC, part II). Suppose g is continuous on [a, b], differentiable on (a, b), and moreover
that g
is integrable on (a, b) (recall Definition 4.12). Then,
Z
b
a
g
= g(b) g(a)
Part II is often expressed in terms of anti-derivatives: F being an anti-derivative of f if F
= f . Com-
bined with FTC, part I, we recover the familiar +c result and a simpler version of the fundamental
theorem often seen in elementary calculus.
Corollary 4.26. Let f be continuous on [a, b].
If F is an anti-derivative of f , then
R
b
a
f = F(b) F(a).
Every anti-derivative of f has the form F(x) =
R
x
a
f (t) dt + c for some constant c.
Examples 4.27. Again, basic examples should be familiar.
1. Plainly g(x) = x
2
+ 2x
3/2
is continuous on [1, 4] and differentiable on (1, 4) with derivative
g
(x) = 2x + 3
x; this last is continuous (and thus integrable) on (1, 4). We conclude that
Z
4
1
2x + 3
x dx = x
2
+ 2x
3/2
4
1
= (16 + 16) (1 + 2) = 29
2. If g(x) = sin( 3x
2
), then g
(x) = 6x cos(3x
2
). Certainly g satisfies the hypotheses of the theorem
on any bounded interval [a, b]. We conclude
Z
b
a
6x cos(3x
2
) dx = sin(3b
2
) sin(3a
2
)
Moreover, every anti-derivative of f (x) = 6x cos(3x
2
) has the form F(x) = sin( 3x
2
) + c.
3. Recall Example 4.24.4 where the discontinuity of f at x = 1 led to the non-differentiability of
F(x) =
R
x
0
f (t) dt. The function F therefore fails the hypotheses of FTC II on the interval [0, 2].
It almost, however, satisfies the conclusions of FTC II, though this is somewhat tautological
given the definition of F: except at x = 1, F is certainly an anti-derivative of f , and moreover
R
2
0
f (x) dx = F(2) F(0).
In case you’re worried that this makes the theorem trivial, note that other anti-derivatives
ˆ
F of
f exist (except at x = 1) which fail to satisfy the conclusion. For instance
ˆ
F(x) =
(
x
2
if x < 1
1
2
x if x > 1
=
ˆ
F(2)
ˆ
F(0) = 1 =
3
2
=
Z
2
0
f (x) dx
86
Proving FTC II Exercise 10 offers a relatively easy proof when g
= f is continuous. For the real
McCoy, we can only rely on the integrability of g
: the trick is to use the mean value theorem to write
g(b) g(a) as a Riemann sum over a suitable partition.
Proof. Suppose ϵ > 0 is given. Since g
is integrable, we may choose some partition P satisfying
U(g
, P) L(g
, P) < ϵ. Since g satisfies the mean value theorem on each subinterval,
ξ
i
(x
i1
, x
i
) such that g
(ξ
i
) =
g(x
i
) g(x
i1
)
x
i
x
i1
from which
g(b) g(a) =
n
i=1
g(x
i
) g(x
i1
) =
n
i=1
g
(ξ
i
)(x
i
x
i1
)
This is a Riemann sum for g
associated to the partition P. Since the upper and lower Darboux sums
are the supremum and infimum of these, we see that
L(g
, P) g(b) g(a) U(g
, P)
However
R
b
a
g
satisfies the same inequality: L(g
, P)
R
b
a
g
U(g
, P). Since these inequalities
hold for all ϵ > 0, we conclude that
R
b
a
g
= g(b) g(a).
While we certainly used the integrability of g
in the proof, it might seem strange that we assumed it
at all: shouldn’t every derivative be integrable? Perhaps surprisingly, the answer is no! If you want
a challenge, look up the Volterra function, which is differentiable everywhere but whose derivative is
non-integrable!
The Rules of Integration
If one wants to evaluate an integral, rather than merely show it exists, there are really only two options:
1. Evaluate Riemann sums and take limits. This is often difficult if not impossible to do explicitly.
2. Use FTC II. The problem now becomes the finding of anti-derivatives, for which the core method
is essentially guess and differentiate. To obtain general rules, we can attempt to reverse the rules
of differentiation.
Integration by Parts Recall the product rule: the product g = uv of two differentiable functions is
differentiable with g
= u
v + uv
. Now apply Theorems 4.9, 4.13 and FTC II.
Corollary 4.28 (Integration by Parts). Suppose u, v are continuous on [a, b], differentiable on (a, b),
and that u
, v
are integrable on (a, b). Then
Z
b
a
u
(x)v(x)dx = u(b)v(b) u(a)v(a)
Z
b
a
u(x)v
(x)dx
This is significantly less useful than the product rule since it merely transforms the integral of one
product into the integral of another.
87
Examples 4.29. With practice, there is no need to explicitly state u and v.
1. Let u(x) = x and v
(x) = cos x. Then u
(x) = 1 and v(x) = sin x. These certainly satisfy the
hypotheses. We conclude
Z
π/2
0
x cos x dx =
[
x sin x
]
π/2
0
Z
π/2
0
sin x dx =
π
2
sin
π
2
0
[
cos x
]
π/2
0
=
π
2
+ cos
π
2
cos 0 =
π
2
1
2. Let u(x) = ln x and v
(x) = 1. Then u
(x) =
1
x
and v(x) = x, whence
Z
e
2
e
ln x dx =
[
x ln x
]
e
2
e
Z
e
2
e
x
x
dx = e
2
ln e
2
e ln e
[
x
]
e
2
e
= 2e
2
e e
2
+ e = e
2
Change of Variables/Substitution We now turn our attention to the chain rule. If g(x) = F
u(x)
,
where F and u are differentiable, then g is differentiable with
g
(x) =
dg
dx
=
dF
du
du
dx
= F
u(x)
u
(x)
Now integrate both sides; the only issue is what assumptions are needed to invoke FTC II.
Theorem 4.30 (Substitution Rule). Suppose u : [a, b] R and f : range(u) R are continuous.
Suppose also that u is differentiable on (a, b) with integrable derivative u
. Then
Z
b
a
f
u(x)
u
(x) dx =
Z
u(b)
u(a)
f (u) du
This is the famous u-sub’/change-of-variables formula from elementary calculus.
Proof. We leave as an exercise the verification that both integrals exist. By the intermediate and
extreme value theorems, range(u) is a closed bounded interval. Assume range(u) has positive length
for otherwise both integrals are trivially zero.
Choose any c range(u) and define
F : range(u) R by F(v) :=
Z
v
c
f (t) dt
Since f is continuous, by FTC I says that F is differentiable with F
(u) = f (u). But now
Z
b
a
f
u(x)
u
(x) dx =
Z
b
a
d
dx
F
u(x)
dx (chain rule)
= F
u(b)
F
u(a)
(FTC II)
=
Z
u(b)
u(a)
f (u) du
88
Examples 4.31. Successfully applying the substitution rule can require significant creativity.
23
1. To evaluate
R
π
0
2x sin x
2
dx, we consider the substitution u(x) = x
2
defined on [0,
π].
Certainly u is continuous; moreover its derivative u
(x) = 2x is integrable on (0,
π). Finally
f (u) = sin u is continuous on range(u) = [0, π]. The hypotheses are satisfied, whence
Z
π
0
2x sin x
2
dx =
Z
π
0
f
u(x)
u
(x) dx =
Z
u(π)
u(0)
f (u) du =
Z
π
0
sin u du
= cos u
π
0
= 2
2. For the following integral, a simple factorization suggests the substitution u(x) = x
2
2.
Plainly u : [
2,
3] [0, 1] and u
(x) = 2x is integrable. Moreover, f (u) =
1
u
2
+1
is continuous
on range(u) = [0, 1]. We conclude
Z
3
2
2x
x
4
4x
2
+ 5
dx =
Z
3
2
2x
(x
2
2)
2
+ 1
dx =
Z
1
0
1
u
2
+ 1
du = arctan u
1
0
=
π
4
3. The hypotheses on u really are all that’s necessary. In particular, u need not be left-/right-
differentiable at the endpoints of [a, b]. For instance, with f (u) = u
2
and u(x) =
x on [0, 4],
we easily verify
8
3
=
Z
4
0
1
2
x dx =
Z
4
0
x
2
x
dx =
Z
4
0
f
u(x)
u
(x) dx =
Z
2
0
f (u) du =
Z
2
0
u
2
du =
8
3
4. Sloppy ‘substitutions’ might lead to utter nonsense. For instance, u(x) = x
2
suggests
Z
2
1
1
x
dx =
Z
2
1
1
2x
2
2x dx =
Z
4
1
1
2u
du =
1
2
(ln 4 ln 1) = ln 2
This is total gibberish: the first integral does not exist since
1
x
is undefined at 0 (1, 2).
Thankfully, the hypotheses of the substitution rule prevent this: f (u) =
1
2u
is not continuous
on range(u) = [0, 4].
While you are very unlikely to make precisely this mistake, the risk is real in more complicated
or abstract situations. . .
23
Hence the old adage, “Differentiation is a science, whereas integration is an art.” To illustrate by example, consider
f (x) = tan(e
x
cos(3x
2
) + 4x
3
). The derivative is easily found using the product and chain rules:
d f
dx
=
1
1 + (e
x
cos(3x
2
) + 4x
3
)
2
e
x
cos(3x
2
) 6xe
x
sin(3x
2
) + 12x
2
By contrast, if you want to find an explicit anti-derivative of f (x), the integration analogues (parts/substitution) are essen-
tially useless. Similarly, the integral
Z
1
0
tan(e
x
cos(3x
2
) + 4x
3
) dx
is likely impossible to evaluate explicitly and can only be approximated, say by using Riemann sums.
89
Exercises 4.34. Key concepts: Complete statements of FTC parts I & II, Integration by Parts/Substitution
1. Calculate the following limits:
(a) lim
x0
1
x
Z
x
0
e
t
2
dt (b) lim
h0
1
h
Z
3+h
3
e
t
2
dt
2. Let f (t) =
0 if t < 0
t if 0 t 1
4 if t > 1
(a) Determine the function F(x) =
R
x
0
f (t) dt and sketch it. Where is F continuous?
(b) Where is F differentiable? Calculate F
at the points of differentiability.
3. Let f be continuous on R.
(a) Define F(x) =
R
x+1
x1
f (t) dt. Carefully show that F is differentiable on R and compute F
.
(b) Repeat for G(x) =
R
sin x
0
f (t) dt.
4. Recall Examples 4.24.4 and 4.27.3. Describe all anti-derivatives F of f on [0, 1) (1, 2]. Which
satisfy
R
2
0
f (x) dx = F(2) F(0)?
5. Suppose u, v satisfy the hypotheses of integration by parts. By FTC I,
R
x
a
u
(t)v(t) dt is an anti-
derivative of u
(x)v(x): what does integration by parts say is another?
6. Use a substitution to integrate
R
1
0
x
1 x
2
dx
7. Use integration by parts and the substitution rule to evaluate
R
b
0
arcsin x dx for any b < 1.
8. Use integration by parts to evaluate
R
b
0
x arctan x dx for any b > 0
9. If f and u satisfy the hypotheses of the substitution rule, explain why both ( f u)u
and f are
integrable on the required intervals.
10. We prove a simpler version of the fundamental theorem when f : [a, b] R is continuous.
Part I Define F(x) =
R
x
a
f (t) dt. If c, x [a, b] where c = x, prove that
m
F(x) F(c)
x c
M
where m, M are the maximum and minimum values of f (t) on the closed interval with
endpoints c, x; why do m, M exist? Now deduce that F
(c) = f (c).
Part II Now suppose F is any anti-derivative of f on [a, b]. Use part (a) and the mean value
theorem to prove that
R
b
a
f (t) dt = F(b) F(a).
90
4.36 Improper Integrals
The Riemann integral has several limitations. Even allowing for functions to be integrable on open
intervals (Definition 4.12), the existence of
R
b
a
f (x) dx requires both:
That (a, b) be a bounded interval.
That f be bounded on (a, b).
Limits provide a natural way to extend the Riemann integral to unbounded intervals and functions.
Definition 4.32. Suppose f : [a, b) R satisfies the following properties:
f is integrable on every closed bounded subinterval [a, t] [a, b).
If b is finite, then f is unbounded at b (b can be !)
The improper integral of f on [a, b) is
Z
b
a
f (x) dx := lim
tb
Z
t
a
f (x) dx
This is convergent or divergent as is the limit.
If an integral is improper at its lower limit (f : (a, b] R, etc.), then
R
b
a
f (x) dx := lim
sa
+
R
b
s
f (x) dx.
If an integral is improper at both ends, choose any c (a, b) and define
Z
b
a
f (x) dx = lim
sa
+
Z
c
s
f (x) dx + lim
tb
Z
t
c
f (x) dx
provided both one-sided improper integrals exist and the limit sum makes sense.
Theorem 4.14 says that the choice of c for a doubly-improper integral is irrelevant.
Many properties of the Riemann integral transfer naturally to improper integrals, though not every-
thing. . . For example, part 1 of Theorem 4.13 extends:
Theorem 4.33. If 0 f (x) g(x) on [a, b) , then
R
b
a
f
R
b
a
g whenever the integrals exist (standard
or improper). In particular:
R
b
a
f = =
R
b
a
g =
R
b
a
g convergent =
R
b
a
f converges to some value
R
b
a
g
We leave some of the detail to Exercise 7.
91
Examples 4.34. 1.
R
t
0
x
2
dx =
1
3
t
3
for any t > 0. Clearly
Z
0
x
2
dx = lim
t
1
3
t
3
=
More formally, the improper integral
R
0
x
2
dx diverges to infinity.
2. With f (x) = x
4/3
defined on [1, ),
Z
1
x
4/3
dx = lim
t
Z
t
1
x
4/3
dx = lim
t
h
3x
1/3
i
t
1
= lim
t
3 3t
1/3
= 3
3. Consider f (x) =
|
x
|
e
x
2
/2
on (, ). On any bounded interval [0, t],
Z
t
0
f (x) dx =
Z
t
0
xe
x
2
/2
dx =
h
e
x
2
/2
i
t
0
= 1 e
t
2
/2
t
1
By symmetry,
Z
|
x
|
e
x
2
/2
dx = 1 + 1 = 2
This example arises naturally in probability: multiplying by
1
2π
computes the expectation of
|
X
|
when X is a standard normally-distributed random variable
E(
|
X
|
) =
Z
1
2π
|
x
|
e
x
2
/2
dx =
r
2
π
4. Our knowledge of derivatives
d
dx
sin
1
x =
1
1x
2
(or the substitution rule) allows us to evaluate
Z
1
0
1
1 x
2
dx = lim
t1
Z
t
0
1
1 x
2
dx = lim
t1
sin
1
t =
π
2
By symmetry,
R
1
1
1
1x
2
dx = π. By comparison, we obtain bounds on another improper inte-
gral:
1
1 x
4
1
1 x
2
=
Z
1
1
1
1 x
4
dx
Z
1
1
1
1 x
2
dx = π
5. Improper integrals need not exist. For instance,
lim
t
Z
t
0
sin x dx = lim
t
1 cos t
diverges by oscillation.
92
Exercises 4.36. Key concepts: Formal definition and careful calculation of Improper Integrals
1. Use your answers from Section 4.34 to decide whether the improper integrals
R
1
0
arcsin x dx
and
R
0
x arctan x dx exist. If so, what are their values?
2. Let p be a positive constant. Prove:
Z
1
0
1
x
p
dx =
(
1
1p
if p < 1
if p 1
Z
1
1
x
p
dx =
(
1
p1
if p > 1
if p 1
(The first of these justifies the convergence/divergence properties of p-series via the integral test)
3. Suppose f is integrable on [a, b]. Explain why
R
b
a
f (x) dx = lim
tb
R
t
a
f (x) dx is still true, even
though the integral is not improper.
4. State a version of integration by parts modified for when
R
b
a
u
(x)v(x) dx is improper at b. Now
evaluate
R
0
xe
4x
dx.
5. What is wrong with the following calculation?
Z
x dx = lim
t
1
2
x
2
t
t
= lim
t
1
2
(t
2
t
2
) = lim
t
0 = 0
6. Prove or disprove: if
R
f and
R
g are convergent improper integrals, so is
R
f g.
7. Prove part of Theorem 4.33. Suppose 0 f (x) g(x) for all x [a, b), and that
R
b
a
g is a
convergent improper integral. Prove that
R
b
a
f converges and that
R
b
a
f
R
b
a
g.
93
Extensions of the Riemann Integral (just for fun)
In the 1890s, Thomas Stieltjes
24
offered a generalization of the Riemann integral.
Definition 4.35. Let f : [a, b] R be bounded and α : [a, b] R monotonically increasing. Given a
partition P = {x
0
, . . . , x
n
} of [a, b], define the sequence of differences
α
i
= α(x
i
) α(x
i1
)
The upper/lower Darboux–Stieltjes sums/integrals are defined analogously to the pure Riemann case:
U( f , P, α) =
n
i=1
sup
[x
i1
,x
i
]
f (x) α
i
L( f , P, α) =
n
i=1
inf
[x
i1
,x
i
]
f (x) α
i
U( f , α) = inf
P
U( f , P, α) L( f , α) = sup
P
L( f , P, α)
If U( f , α) = L( f , α), we say that f is Riemann–Stieltjes integrable of class R(α) and denote its value
R
b
a
f (x) dα.
The standard Riemann integral corresponds to α(x) = x. It is the ability to choose other functions α
that makes the Riemann–Stieltjes integral both powerful and applicable.
Standard Properties Most results in sections 4.32 and 4.33 hold with suitable modifications, as does
the discussion of improper integrals. For instance,
f R(α) P such that U( f , P, α) L( f , P, α) < ϵ
The result regarding the piecewise continuity of f is a notable exception: depending on α, a
piecewise continuous f might not lie in R(α).
Weighted integrals If α is differentiable, we obtain a standard Riemann integral
Z
b
a
f (x) dα =
Z
b
a
f (x)α
(x) dx
weighted so that f (x) contributes more when α is increasing rapidly.
Probability If α (a) = 0 and α(b) = 1, then α may be viewed as a probability distribution function and its
derivative α
as the corresponding probability density function. For example:
1. The uniform distribution on [a, b] has α =
1
ba
(x a) so that
Z
b
a
f (x) dα =
1
b a
Z
b
a
f (x) dx
Since α
is constant, the integrals weigh all values of x uniformly.
2. The standard normal distribution has α(x) =
R
x
1
2π
e
t
2
/2
dt. The fact that α
=
1
2π
e
x
2
/2
is maximal when x = 0 reflects the fact that a normally distributed variable is clustered
near its mean.
In all cases,
R
f (x) dα = E( f (X)) computes an expectation (see, e.g., Example 4.34.3).
24
Stieltjes was Dutch; the pronunciation is roughly ‘steelchez.’
94
Non-differentiable or continuous α This provides major flexibility! For example, if Q = {s
0
, . . . , s
n
}
partitions [a, b], and (c
k
)
n
k=1
is a positive sequence, then
α(x) =
0 if x = a
k
i=1
c
i
if x (s
k1
, s
k
]
defines an increasing step function, and the Riemann–Stieltjes integral a weighted sum
Z
b
a
f (x) dα =
n
i=1
c
i
f (s
i
)
Taking an infinite increasing sequence (s
n
) [a, b] results in an infinite series, which helps
explain why so many results for series and integrals look similar!
This also touches on probability. For example, let p [0, 1], n N, and s
k
= k on the interval
[0, n]. If c
k
=
(
n
k
)
p
k
(1 p)
nk
, then
Z
f (x) dα =
n
k=0
n
k
p
k
(1 p)
nk
f (x) = E( f (X))
is the expectation of f (X) when X B(n, p) is binomially distributed.
Lebesgue Integration: Integrals and Convergence
Lebesgue’s extension essentially uses rectangles whose heights tend to zero: cutting up the area under
a curve using horizontal instead of vertical strips. One of its major purposes is to permit a more general
interchange of limits and integration in many cases of pointwise (non-uniform) convergence. To see
the problem, consider the sequence of piecewise continuous functions
f
n
: [0, 1] R : x 7→
(
1 if x =
p
q
Q with q n
0 otherwise
Each f
n
is Riemann integrable with
R
1
0
f
n
(x) dx = 0. However, the pointwise limit
f (x) =
(
1 if x Q
0 if x Q
is not Riemann integrable (compare Example 4.8.2). In the Lebesgue theory, the limit f turns out to
be integrable with integral 0, so that
lim
n
Z
1
0
f
n
(x) dx =
Z
1
0
lim
n
f
n
(x) dx
Recall (Theorem 2.19) that the interchange of limits and integrals would be automatic if the conver-
gence f
n
f were uniform: of course the convergence isn’t uniform here.
Like measure theory (recall Theorem 4.22), Lebesgue integration is a central topic in graduate analysis.
95