SlideShare una empresa de Scribd logo
1 de 22
3
The Gauge Integral




3.1 Introduction
Much of calculus deals with the interplay between differentiation and
integration. The antiquated term “antidifferentiation” emphasizes the fact
that differentiation and integration are inverses of one another. We will
take it for granted that readers are acquainted with the mechanics of in-
tegration. The current chapter develops just enough integration theory to
make our development of differentiation in Chap. 4 and the calculus of
variations in Chap. 17 respectable. It is only fair to warn readers that in
other chapters a few applications to probability and statistics will assume
familiarity with properties of the expectation operator not covered here.
   The first successful effort to put integration on a rigorous basis was un-
dertaken by Riemann. In the early twentieth century, Lebesgue defined
a more sophisticated integral that addresses many of the limitations of
the Riemann integral. However, even Lebesgue’s integral has its defects.
In the past few decades, mathematicians such as Henstock and Kurzweil
have expanded the definition of integration on the real line to include
a wider variety of functions. The new integral emerging from these in-
vestigations is called the gauge integral or generalized Riemann integral
[7, 68, 108, 193, 250, 255, 278]. The gauge integral subsumes the Riemann
integral, the Lebesgue integral, and the improper integrals met in tradi-
tional advanced calculus courses. In contrast to the Lebesgue integral, the
integrands of the gauge integral are not necessarily absolutely integrable.


K. Lange, Optimization, Springer Texts in Statistics 95,                 53
DOI 10.1007/978-1-4614-5838-8 3,
© Springer Science+Business Media New York 2013
54     3. The Gauge Integral

   It would take us too far afield to develop the gauge integral in full
generality. Here we will rest content with proving some of its elementary
properties. One of the advantages of the gauge integral is that many theo-
rems hold with fewer qualifications. The fundamental theorem of calculus is
a case in point. The commonly stated version of the fundamental theorem
concerns a differentiable function f (x) on an interval [a, b]. As all students
of calculus know,
                              b
                                  f (x) dx    =       f (b) − f (a).
                          a

Although this version is true for the gauge integral, it does not hold for the
Lebesgue integral because the mere fact that f (x) exists throughout [a, b]
does not guarantee that it is Lebesgue integrable.
  This quick description of the gauge integral is not intended to imply that
the gauge integral is uniformly superior to the Lebesgue integral and its
extensions. Certainly, probability theory would be severely handicapped
without the full flexibility of modern measure theory. Furthermore, the ad-
vanced theory of the gauge integral is every bit as difficult as the advanced
theory of the Lebesgue integral. For pedagogical purposes, however, one can
argue that a student’s first exposure to the theory of integration should fea-
ture the gauge integral. As we shall see, many of the basic properties of
the gauge integral flow directly from its definition. As an added dividend,
gauge functions provide an alternative approach to some of the material of
Chap. 2.


3.2 Gauge Functions and δ-Fine Partitions
The gauge integral is defined through gauge functions. A gauge function
is nothing more than a positive function δ(t) defined on a finite interval
[a, b]. In approximating the integral of a function f (t) over [a, b] by a finite
Riemann sum, it is important to sample the function most heavily in those
regions where it changes most rapidly. Now by a Riemann sum we mean a
sum
                                             n−1
                       S(f, π)        =            f (ti )(si+1 − si ),
                                             i=0

where the mesh points a = s0 < s1 < · · · < sn = b form a partition π of
[a, b], and the tags ti are chosen so that ti ∈ [si , si+1 ]. If δ(ti ) measures the
rapidity of change of f (t) near ti , then it makes sense to take δ(t) small in
regions of rapid change and to force si and si+1 to belong to the interval
(ti − δ(ti ), ti + δ(ti )). A tagged partition with this property is called a δ-
fine partition. Our first proposition relieves our worry that δ-fine partitions
exist.
3.2 Gauge Functions and δ-Fine Partitions         55

Proposition 3.2.1 (Cousin’s Lemma) For every gauge δ(t) on a finite
interval [a, b] there is a δ-fine partition.

Proof: Assume that [a, b] lacks a δ-fine partition. Since we can construct a
δ-fine partition of [a, b] by appending a δ-fine partition of the half-interval
[(a + b)/2, b] to a δ-fine partition of the half-interval [a, (a + b)/2], it fol-
lows that either [a, (a + b)/2] or [(a + b)/2, b] lacks a δ-fine partition. As in
Example 2.3.1, we choose one of the half-intervals based on this failure and
continue bisecting. This creates a nested sequence of intervals [ai , bi ] con-
verging to a point x. If i is large enough, then [ai , bi ] ⊂ (x − δ(x), x + δ(x)),
and the interval [ai , bi ] with tag x is a δ-fine partition of itself. This con-
tradicts the choice of [ai , bi ] and the assumption that the original interval
[a, b] lacks a δ-fine partition.
   Before launching into our treatment of the gauge integral, we pause to
gain some facility with gauge functions [108]. Here are three examples that
illustrate their value.
Example 3.2.1 A Gauge Proof of Weierstrass’ Theorem
Consider a real-valued continuous function f (t) with domain [a, b]. Suppose
that f (t) does not attain its supremum on [a, b]. Then for each t there exists
a point x ∈ [a, b] with f (t) < f (x). By continuity there exists δ(t) > 0 such
that f (y) < f (x) for all y ∈ [a, b] with |y − t| < δ(t). Using δ(t) as a
gauge, select a δ-fine partition a = s0 < s1 < · · · < sn = b with tags
ti ∈ [si , si+1 ] and designated points xi satisfying f (ti ) < f (xi ). Let xmax
be the point xi having the largest value f (xi ). Because xmax lies in some
interval [si , si+1 ], we have f (xmax ) < f (xi ). This contradiction discredits
our assumption that f (x) does not attain its supremum. A similar argument
applies to the infimum.

Example 3.2.2 A Gauge Proof of the Heine-Borel Theorem

One can use Cousin’s lemma to prove the Heine-Borel Theorem on the real
line [278]. This theorem states that if C is a compact set contained in the
union ∪α Oα of a collection of open sets Oα , then C is actually contained in
the union of a finite number of the Oα . Suppose C ⊂ [a, b]. Define a gauge
δ(t) so that the interval (t − δ(t), t + δ(t)) does not intersect C when t ∈ C
and (t − δ(t), t + δ(t)) is contained in some Oα when t ∈ C. Based on δ(t),
select a δ-fine partition a = s0 < s1 < · · · < sn = b with tags ti ∈ [si , si+1 ].
By definition C is contained in the union ∪ti ∈C Ui , where Ui is the set Oα
covering ti . The Heine-Borel theorem extends to compact sets in Rn .

Example 3.2.3 A Gauge Proof of the Intermediate Value Theorem

Under the assumption of the previous example, let c be an number strictly
between f (a) and f (b). If we assume that there is no t ∈ [a, b] with f (t) = c,
then there exists a positive number δ(t) such that either f (x) < c for all
56     3. The Gauge Integral

x ∈ [a, b] with |x − t| < δ(t) or f (x) > c for all x ∈ [a, b] with |x − t| < δ(t).
We now select a δ-fine partition a = s0 < s1 < · · · < sn = b and observe
that throughout each interval [si , si+1 ] either f (t) < c or f (t) > c. If to
start f (s0 ) = f (a) < c, then f (s1 ) < c, which implies f (s2 ) < c and so
forth until we get to f (sn ) = f (b) < c. This contradicts the assumption
that c lies strictly between f (a) and f (b). With minor differences, the same
proof works when f (a) > c.
   In preparation for our next example and for the fundamental theorem
of calculus later in this chapter, we must define derivatives. A real-valued
function f (t) defined on an interval [a, b] possesses a derivative f (c) at
c ∈ [a, b] provided the limit
                              f (t) − f (c)
                           lim                      =     f (c)              (3.1)
                          t→c     t−c
exists. At the endpoints a and b, the limit is necessarily one sided. Tak-
ing a sequential view of convergence, definition (3.1) means that for every
sequence tm converging to c we must have
                                 f (tm ) − f (c)
                         lim                          =       f (c).
                        m→∞          tm − c
In calculus, we learn the following rules for computing derivatives:
Proposition 3.2.2 If f (t) and g(t) are differentiable functions on (a, b),
then

                   αf (t) + βg(t)        =   αf (t) + βg (t)

                         f (t)g(t)       =   f (t)g(t) + f (t)g (t)
                                 1                 f (t)
                                         =   −            .
                               f (t)               f (t)2
In the third formula we must assume f (t) = 0. Finally, if g(t) maps into
the domain of f (t), then the functional composition f ◦ g(t) has derivative
                         [f ◦ g(t)]     =    f ◦ g(t)g (t).
Proof: We will prove the above sum, product, quotient, and chain rules in
a broader context in Chap. 4. Our proofs will not rely on integration.

Example 3.2.4 Strictly Increasing Functions
Let f (t) be a differentiable function on [c, d] with strictly positive derivative.
We now show that f (t) is strictly increasing. For each t ∈ [c, d] there exists
δ(t) > 0 such that
                                 f (x) − f (t)
                                                    >     0                  (3.2)
                                     x−t
3.3 Definition and Basic Properties of the Integral     57

for all x ∈ [a, b] with |x − t| < δ(t). According to Proposition 3.2.1, for any
two points a < b from [c, d], there exists a δ-fine partition

                        a = s0 < s1 < · · · < sn = b

of [a, b] with tags ti ∈ [si , si+1 ]. In view of inequality (3.2), at least one
of the two inequalities f (si ) ≤ f (ti ) ≤ f (si+1 ) must be strict. Thus, the
telescoping sum
                                        n−1
                    f (b) − f (a) =           [f (si+1 ) − f (si )]
                                        i=0

must be positive.



3.3 Definition and Basic Properties of the Integral
With later applications in mind, it will be convenient to define the gauge
integral for vector-valued functions f (x) : [a, b] → Rn . In this context, f (x)
is said to have integral I if for every > 0 there exists a gauge δ(x) on
[a, b] such that

                               S(f, π) − I       <                           (3.3)

for all δ-fine partitions π. Our first order of business is to check that the
integral is unique whenever it exists. Thus, suppose that the vector J is a
second possible value of the integral. Given > 0 choose gauges δI (x) and
δJ (x) leading to inequality (3.3). The minimum δ(x) = min{δI (x), δJ (x)}
is also a gauge, and any partition π that is δ-fine is also δI and δJ -fine.
Hence,

          I −J      ≤      I − S(f, π) + S(f, π) − J             < 2 .

Since is arbitrary, J = I.
  One can also define f (x) to be integrable if its Riemann sums are Cauchy
in an appropriate sense.
Proposition 3.3.1 (Cauchy criterion) A function f (x) : [a, b] → Rn is
integrable if and only if for every > 0 there exists a gauge δ(x) > 0 such
that

                           S(f, π1 ) − S(f, π2 )      <                      (3.4)

for any two δ-fine partitions π1 and π2 .
Proof: It is obvious that the Cauchy criterion is necessary for integrability.
To show that it is sufficient, consider the sequence m = m−1 and compat-
ible sequence of gauges δm (x) determined by condition (3.4). We can force
58      3. The Gauge Integral

the constraints δm (x) ≤ δm−1 (x) to hold by inductively replacing δm (x) by
min{δm−1 (x), δm (x)} whenever needed. Now select a δm -fine partition πm
for each m. Because the gauge sequence δm (x) is decreasing, every partition
π that is δm -fine is also δm−1 -fine. Hence, the sequence of Riemann sums
S(f, πm ) is Cauchy and has a limit I satisfying S(f, πm ) − I ≤ m−1 .
Finally, given the potential integral I, we take an arbitrary > 0 and
choose m so that m−1 < . If π is δm -fine, then the inequality

     S(f, π) − I         ≤         S(f, π) − S(f, πm ) + S(f, πm ) − I                                                 < 2 .

completes the proof.
  For two integrable functions f (x) and g(x), the gauge integral inherits
the linearity property
                b                                                                b                           b
                    [αf (x) + βg(x)] dx                 = α                          f (x) dx + β                g(x) dx
            a                                                                a                           a


from its approximating Riemann sums. To prove this fact, take                                                              > 0 and
choose gauges δf (x) and δg (x) so that
                              b                                                                   b
        S(f, πf ) −               f (x) dx < ,                    S(f, πg ) −                         g(x) dx <
                          a                                                                   a

whenever πf is δf -fine and πg is δg -fine. If the tagged partition π is δ-fine
for the gauge δ(x) = min{δf (x), δg (x)}, then
                                                             b                                    b
                      S(αf + βg, π) − α                          f (x) dx − β                         g(x) dx
                                                         a                                    a
                                                  b                                                              b
          ≤ |α| S(f, π) −                             f (x) dx + |β| S(g, π) −                                       g(x) dx
                                              a                                                              a
          ≤ (|α| + |β|) .

  The gauge integral also inherits obvious order properties. For example,
 b
 a f (x) dx ≥ 0 whenever the integrand f (x) ≥ 0 for all x ∈ [a, b]. In this
                                b
case, the inequality |S(f, π) − a f (x) dx| < implies
                                                                         b
                         0 ≤ S(f, π) ≤                                       f (x) dx + .
                                                                     a

Since can be made arbitrarily small for f (x) integrable, it follows that
 b
 a
   f (x) dx ≥ 0. This nonnegativity property translates into the
order property
                                         b                                           b
                                             f (x) dx            ≤                       g(x) dx
                                     a                                           a
3.3 Definition and Basic Properties of the Integral                            59

for two integrable functions f (x) ≤ g(x). In particular, when both f (x)
and |f (x)| are both integrable, we have
                                    b                              b
                                        f (x) dx       ≤               |f (x)| dx .
                                a                              a

For vector-valued functions, the analogous rule
                                    b                              b
                                        f (x) dx       ≤                f (x) dx                       (3.5)
                                a                              a

is also inherited from the approximating Riemann sums. The reader can
easily supply the proof using the triangle inequality of the Euclidean norm.
It does not take much imagination to extend the definition of the gauge
integral to matrix-valued functions, and inequality (3.5) applies in this
setting as well.
   One of the nicest features of the gauge integral is that one can perturb
an integrable function at a countable number of points without changing
the value of its integral. This property fails for the Riemann integral but is
exhibited by the Lebesgue integral. To validate the property, it suffices to
prove that a function that equals 0 except at a countable number of points
has integral 0. Suppose f (x) is such a function with exceptional points
x1 , x2 , . . . and corresponding exceptional values f 1 , f 2 , . . .. We now define
a gauge δ(x) with value 1 on the nonexceptional points and values

                                δ(xj ) =
                                                   2j+2 [ f j + 1]
at the exceptional points. If π is a δ-fine partition, then xj can serve as
a tag for at most two intervals [si , si+1 ] of π and each such interval has
length less than 2δ(xj ). It follows that
                                                                                         ∞
                                                        2                                     1
     S(f, π)    ≤ 2                 f (xj )       j+2 [ f
                                                                 ≤                                 =
                            j
                                                2         j + 1]                        j=1
                                                                                              2j

                           b
and therefore that a f (x) dx = 0.
  In practice, the interval additivity rule
                       c                               b                          c
                           f (x) dx         =              f (x) dx +                 f (x) dx         (3.6)
                   a                               a                          b

is obviously desirable. There are three separate issues in proving it. First,
given the existence of the integral over [a, c], do the integrals over [a, b]
and [b, c] exist? Second, if the integrals over [a, b] and [b, c] exist, does
the integral over [a, c] exist? Third, if the integrals over [a, b] and [b, c]
exist, are they additive? The first question is best approached through
Proposition 3.3.1. For > 0 there exists a gauge δ(x) such that
                                    S(f, π1 ) − S(f, π2 )                 <
60    3. The Gauge Integral

for any two δ-fine partitions π1 and π2 of [a, c]. Given δ(x), take any two
δ-fine partitions γ1 and γ2 of [a, b] and a single δ-fine partition ω of [b, c].
The concatenated partitions γ1 ∪ ω and γ2 ∪ ω are δ-fine throughout [a, c]
and satisfy

     S(f, γ1 ) − S(f, γ2 )           =        S(f, γ1 ∪ ω) − S(f, γ2 ∪ ω)                       <    .

According to the Cauchy criterion, the integral over [a, b] therefore exists.
A similar argument implies that the integral over [b, c] also exists. Finally,
the combination of these results shows that the integral exists over any
interval [u, v] contained within [a, b].
  For the converse, choose gauges δ1 (x) on [a, b] and δ2 (x) on [b, c] so that
                      b                                                      c
     S(f, γ) −            f (x) dx       <        ,    S(f, ω) −                 f (x) dx        <
                  a                                                      b

for any δ1 -fine partition γ of [a, b] and any δ2 -fine partition ω of [b, c]. The
concatenated partition π = γ ∪ ω satisfies
                                             b                    c
                          S(f, π) −              f (x) dx −           f (x) dx
                                         a                    b
                                          b                                           c
              ≤           S(f, γ) −              f (x) + S(f, ω) −                        f (x) dx       (3.7)
                                         a                                        b
              < 2

because the Riemann sums satisfy S(f, π) = S(f, γ)+S(f, ω). This suggests
defining a gauge δ(x) equal to δ1 (x) on [a, b] and equal to δ2 (x) on [b, c].
The problem with this tactic is that some partitions of [a, c] do not split
at b. However, we can ensure a split by redefining δ(x) by

                      ˜                  min{δ1 (b), δ2 (b)}                     x=b
                      δ(x)       =
                                         min{δ(x), 1 |x − b|}
                                                    2                            x = b.

This forces b to be the tag of its assigned interval, and we can if needed
split this interval at b and retain b as tag of both subintervals. With δ(x)
amended in this fashion, any δ-fine partition π can be viewed as a con-
catenated partition γ ∪ ω splitting at b. As such π obeys inequality (3.7).
This argument simultaneously proves that the integral over [a, c] exists and
satisfies the additivity property (3.6)
   If the function f (x) is vector-valued with n components, then the in-
tegrability of f (x) should imply the integrability of each its components
fi (x). Furthermore, we should be able to write
                                       ⎛ b            ⎞
                        b
                                             f (x) dx
                                           a 1
                                       ⎜        .     ⎟
                          f (x) dx = ⎝          .
                                                .     ⎠.
                             a                            b
                                                          a   fn (x) dx
3.3 Definition and Basic Properties of the Integral                           61

Conversely, if its components are integrable, then f (x) should be integrable
as well. The inequalities
                                        n
                                                                                  √
      S(f, π) − I              ≤              |S(fi , π) − Ii | ≤                     n S(f, π) − I .
                                        i=1


based on Example 2.5.6 and Problem 3 of Chap. 2 are instrumental in
proving this logical equivalence. Given that we can integrate component
by component, for the remainder of this chapter we will deal exclusively
with real-valued functions.
  We have not actually shown that any function is integrable. The most
obvious possibility is a constant. Fortunately, it is trivial to demonstrate
that
                                                b
                                                    c dx      = c(b − a).
                                            a

Step functions are one rung up the hierarchy of functions. If

                                                             n−1
                                    f (x)           =              ci 1(si ,si+1 ] (x)
                                                             i=0


for a = s0 < s1 < · · · < sn = b, then our nascent theory allows us to
evaluate
             b                          n−1          si+1                      n−1
                 f (x) dx =                                  ci dx =                  ci (si+1 − si ).
         a                              i=0         si                          i=0


This fact and the next technical proposition turn out to be the key to
showing that continuous functions are integrable.

Proposition 3.3.2 Let f (x) be a function with domain [a, b]. Suppose for
every > 0 there exist two integrable functions g(x) and h(x) satisfying
g(x) ≤ f (x) ≤ h(x) for all x and

                                    b                                    b
                                        h(x) dx              ≤               g(x) dx + .
                                a                                    a

Then f (x) is integrable.

Proof: For        > 0, choose gauges δg (x) and δh (x) on [a, b] so that

                         b                                                                b
    S(g, πg ) −              g(x) dx            <        ,       S(h, πh ) −                  h(x) dx   <
                     a                                                                a
62    3. The Gauge Integral

for any δg -fine partition πg and any δh -fine partition πh . If π is a δ-fine
partition for δ(x) = min{δg (x), δh (x)}, then the inequalities
                             b
                                 g(x) dx −         <     S(g, π)
                         a
                                                   ≤     S(f, π)
                                                   ≤     S(h, π)
                                                              b
                                                   <              h(x) +
                                                          a
                                                              b
                                                   ≤              g(x) dx + 2
                                                          a

trap S(f, π) in an interval of length 3 . Because the Riemann sum S(f, γ)
for any other δ-fine partition γ is trapped in the same interval, the integral
of f (x) exists by the Cauchy criterion.

Proposition 3.3.3 Every continuous function f (x) on [a, b] is integrable.
Proof: In view of the uniform continuity of f (x) on [a, b], for every > 0
there exists a δ > 0 with |f (x) − f (y)| < when |x − y| < δ. For the
constant gauge δ(x) = δ and a corresponding δ-fine partition π with mesh
points s0 , . . . , sn , let mi be the minimum and Mi be the maximum of f (x)
on [si , si+1 ]. The step functions
                n                                                     n
     g(x) =             mi 1(si ,si+1 ] (x),       h(x) =                  Mi 1(si ,si+1 ] (x)
                i=1                                                  i=1

then satisfy g(x) ≤ f (x) ≤ h(x) except at the single point a. Furthermore,
                   b                       b                         n
                       h(x) dx −               g(x) dx    ≤                (si+1 − si )
               a                       a                             i=1
                                                          =         (b − a).
Application of Proposition 3.3.2 now completes the proof.



3.4 The Fundamental Theorem of Calculus
The fundamental theorem of calculus divides naturally into two parts. For
the gauge integral, the first and more difficult part is easily proved by
invoking what is called the straddle inequality. Let f (x) be differentiable
at the point t ∈ [a, b]. Then there exists δ(t) > 0 such that
                                  f (x) − f (t)
                                                − f (t)             <
                                      x−t
3.4 The Fundamental Theorem of Calculus                  63

for all x ∈ [a, b] with |x − t| < δ(t). If u < t < v are two points straddling t
and located in [a, b] ∩ (t − δ(t), t + δ(t)), then
      |f (v) − f (u) − f (t)(v − u)| ≤ |f (v) − f (t) − f (t)(v − t)|
                                       + |f (t) − f (u) − f (t)(t − u)|
                                             ≤     (v − t) + (t − u)                    (3.8)
                                             =     (v − u).
Inequality (3.8) also clearly holds when either u = t or v = t.
Proposition 3.4.1 (Fundamental Theorem I) If f (x) is differentiable
throughout [a, b], then
                             b
                                 f (x) dx    =    f (b) − f (a).
                         a

Proof: Using the gauge δ(t) figuring in the straddle inequality (3.8), select
a δ-fine partition π with mesh points a = s0 < s1 < · · · < sn = b and tags
ti ∈ [si , si+1 ]. Application of the inequality and telescoping yield
                                       n−1
  |f (b) − f (a) − S(f , π)| =               [f (si+1 ) − f (si ) − f (ti )(si+1 − si )]
                                       i=0
                                      n−1
                                  ≤          |f (si+1 ) − f (si ) − f (ti )(si+1 − si )|
                                      i=0
                                      n−1
                                  ≤           (si+1 − si )
                                       i=0
                                  =    (b − a).
This demonstrates that f (x) has integral f (b) − f (a).
  The first half of the fundamental theorem remains valid for a continuous
function f (x) that is differentiable except on a countable set N [250]. Since
changing an integrand at a countable number of points does not alter its
integral, it suffices to prove that
                             b
                                                                     0       t∈N
     f (b) − f (a) =             g(t) dt, where g(t)         =
                         a                                           f (t)   t ∈ N.
Suppose > 0 is given. For t ∈ N define the gauge value δ(t) to satisfy
the straddle inequality. Enumerate the points tj of N, and define δ(tj ) > 0
so that |f (tj ) − f (tj + s)| < 2−j−2 whenever |s| < δ(tj ). Now select a
δ-fine partition π with mesh points a = s0 < s1 < · · · < sn = b and tags
ri ∈ [si , si+1 ]. Break the sum
                                      n−1
    f (b) − f (a) − S(g, π) =                f (si+1 ) − f (si ) − g(ri )(si+1 − si )
                                      i=0
64    3. The Gauge Integral

into two parts. Let S denote the sum of the terms with tags ri ∈ N, and
let S denote the sum of the terms with tags ri ∈ N . As noted earlier,
|S | ≤ (b − a). Because a tag is attached to at most two subintervals, the
second sum satisfies

            |S | ≤               |f (si+1 ) − f (si )|
                       ri ∈N

                  ≤               |f (si+1 ) − f (ri )| + |f (ri ) − f (si )|
                       ri ∈N
                           ∞
                  ≤    2         22−j−2        =           .
                           j=1


It follows that |S + S | ≤ (b − a + 1) and therefore that the stated integral
exists and equals f (b) − f (a).
   In demonstrating the second half of the fundamental theorem, we will
implicitly use the standard convention
                           c                                   d
                               f (x) dx    =    −                  f (x) dx
                       d                                   c

for c < d. This convention will also be in force in proving the substitution
formula.
Proposition 3.4.2 (Fundamental Theorem II) If a function f (x) is
integrable on [a, b], then its indefinite integral
                                                   t
                                F (t) =                f (x) dx
                                               a

has derivative F (t) = f (t) at any point t where f (x) is continuous. The
derivative is taken as one sided if t = a or t = b.
Proof: In deriving the interval additivity rule (3.6), we showed that the
integral F (t) exists. At a point t where f (x) is continuous, for any > 0
there is a δ > 0 such that − < f (x) − f (t) < when |x − t| < δ and
x ∈ [a, b]. Hence, the difference

           F (t + s) − F (t)                           1           t+s
                             − f (t) =                                   [f (x) − f (t)] dx
                   s                                   s       t

is less than and greater than − for |s| < δ. In the limit as s tends to 0,
we recover F (t) = f (t).
  The fundamental theorem of calculus has several important corollaries.
These are covered in the next three propositions on the substitution rule,
integration by parts, and finite Taylor expansions.
3.4 The Fundamental Theorem of Calculus                    65

Proposition 3.4.3 (Substitution Rule) Suppose f (x) is differentiable
on [a, b], g(x) is differentiable on [c, d], and the image of [c, d] under g(x)
is contained within [a, b]. Then
                              g(d)                               d
                                     f (y) dy           =            f [g(x)]g (x) dx.
                          g(c)                               c

Proof: Part I of the fundamental theorem and the chain rule identity

                                     {f [g(x)]}          =   f [g(x)]g (x)

imply that both integrals have value f [g(d)] − f [g(c)].

Proposition 3.4.4 (Integration by Parts) Suppose f (x) and g(x) are
differentiable on [a, b]. Then f (x)g(x) is integrable on [a, b] if and only if
f (x)g (x) is integrable on [a, b]. Furthermore, the two integrals are related
by the identity
            b                                b
                f (x)g(x) dx +                   f (x)g (x) dx         = f (b)g(b) − f (a)g(a),
        a                                a

Proof: The product rule for derivatives is

                      [f (x)g(x)]                 = f (x)g(x) + f (x)g (x).

If two of three members of this identity are integrable, then the third is as
well. Since part I of the fundamental theorem entails
                          b
                              [f (x)g(x)] dx             =   f (b)g(b) − f (a)g(a),
                      a

the proposition follows.
  The derivative of a function may itself be differentiable. Indeed, it makes
sense to speak of the kth-order derivative of a function f (x) if f (x) is
sufficiently smooth. Traditionally, the second-order derivative is denoted
f (x) and an arbitrary kth-order derivative by f (k) (x). We can use these
extra derivatives to good effect in approximating f (x) locally. The next
proposition makes this clear and offers an explicit estimate of the error in
a finite Taylor expansion of f (x).
Proposition 3.4.5 (Taylor Expansion) Suppose f (x) has a derivative
of order k + 1 on an open interval around the point y. Then for all x in the
interval, we have
                                                    k
                                                         1 (j)
                  f (x) =             f (y) +               f (y)(x − y)j + Rk (x),               (3.9)
                                                   j=1
                                                         j!
66    3. The Gauge Integral

where the remainder
                                                    1
                             (x − y)k+1
         Rk (x)    =                                    f (k+1) [y + t(x − y)](1 − t)k dt.
                                 k!             0

If |f (k+1) (z)| ≤ b for all z between x and y, then

                                                          b|x − y|k+1
                                 |Rk (x)|        ≤                    .                      (3.10)
                                                            (k + 1)!

Proof: When k = 0, the Taylor expansion (3.9) reads
                                                                 1
               f (x)         =   f (y) + (x − y)                     f [y + t(x − y)]dt
                                                             0

and follows from the fundamental theorem of calculus and the chain rule.
Induction and the integration-by-parts formula
                       1
                           f (k) [y + t(x − y)](1 − t)k−1 dt
                   0
                    1                            1
           =      − f (k) [y + t(x − y)](1 − t)k
                    k                            0
                    x − y 1 (k+1)
                  +            f     [y + t(x − y)](1 − t)k dt
                      k      0
                                                     1
                  1 (k)     x−y
           =        f (y) +                              f (k+1) [y + t(x − y)](1 − t)k dt
                  k          k                   0

now validate the general expansion (3.9). The error estimate follows directly
from the bound |f (k+1) (z)| ≤ b and the integral
                                      1
                                                                      1
                                          (1 − t)k dt       =            .
                                  0                                  k+1


3.5 More Advanced Topics in Integration
Within the confines of a single chapter, it is impossible to develop rigorously
all of the properties of the gauge integral. In this section we will discuss
briefly four topics: (a) integrals over unbounded intervals, (b) improper
integrals and Hake’s theorem, (c) the interchange of limits and integrals,
and (d) multidimensional integrals and Fubini’s theorem.
   Defining the integral of a function over an unbounded interval requires
several minor adjustments. First, the real line is extended to include the
points ±∞. Second, a gauge function δ(x) is now viewed as mapping x
to an open interval containing x. The associated interval may be infinite;
indeed, it must be infinite if x equals ±∞. In a δ-fine partition π, the
3.5 More Advanced Topics in Integration                   67

interval Ij containing the tag xj is contained in δ(xj ). The length of an
infinite interval Ij is defined to be 0 in an approximating Riemann sum
S(f, π) to avoid infinite contributions to the sum. Likewise, the integrand
f (x) is assigned the value 0 at x = ±∞.
   This extended definition carries with it all the properties we expect. Its
most remarkable consequence is that it obliterates the distinction between
proper and improper integrals. Hake’s theorem provides the link. If we
allow a and b to be infinite as well as finite, then Hake’s theorem says a
function f (x) is integrable over (a, b) if and only if either of the two limits
                                        b                                        c
                              lim           f (x) dx         or        lim           f (x) dx
                              c→a   c                              c→b       a

                                                             b
exists. If either limit exists, then                         a
                                                                 f (x) dx equals that limit. For instance,
the integral
                  ∞                                      c                                        c
                      1                                      1                                1
                         dx =                lim                dx =             lim −                = 1
              1       x2                    c→∞      1       x2              c→∞              x   1

exists and has the indicated limit by this reasoning.
                                                      ∞
Example 3.5.1 Existence of                            0      sinc(x) dx
Consider the integral of sinc(x) = sin(x)/x over the interval (0, ∞). Because
sinc(x) is continuous throughout [0, 1] with limit 1 as x approaches 0, the
integral over [0, 1] is defined. Hake’s theorem and integration by parts show
that the integral
                          ∞                                        c
                              sin x                        sin x
                                    dx          =     lim        dx
                      1         x                    c→∞ 1   x
                                                                         c
                                                           cos x c         cos x
                                                = lim −             −            dx
                                                  c→∞        x 1       1    x2
                                                             ∞
                                                               cos x
                                                = cos 1 −            dx
                                                           1     x2

exists provided the integral of x−2 cos x exists over (1, ∞). We will demon-
strate this fact in a moment. If we accept it, then it is clear that the integral
of sinc(x) over (0, ∞) exists as well. As we shall find in Example 3.5.4, this
integral equals π/2. In contrast to these positive results, sinc(x) is not
absolutely integrable over (0, ∞). Finally, we note in passing that the sub-
stitution rule gives
          ∞                                     ∞                                        ∞
              sin cx                                sin y −1                                 sin y      π
                     dx =                                 c dy =                                   dy =   .
      0         x                           0       c−1 y                            0         y        2

for any c > 0.
68      3. The Gauge Integral

     We now ask under what circumstances the formula
                                          b                                 b
                            lim               fn (x) dx         =                   lim fn (x) dx            (3.11)
                        n→∞           a                                  a n→∞
is valid. The two relevant theorems permitting the interchange of limits
and integrals are the monotone convergence theorem and the dominated
convergence theorem. In the monotone convergence theorem, we are given
an increasing sequence fn (x) of integrable functions that converge to a
finite limit for each x. Formula (3.11) is true in this setting provided
                                                      b
                                          sup             fn (x) dx         <            ∞.
                                              n   a

In the dominated convergence theorem, we assume the sequence fn (x) is
trapped between two integrable functions g(x) and h(x) in the sense that
                                     g(x) ≤ fn (x) ≤ h(x)
for all n and x. If limn→∞ fn (x) exists in this setting, then the inter-
change (3.11) is allowed. The choices
     fn (x) = 1[1,n] (x)x−2 cos x ,                         g(x) = −x−2 ,                       h(x) = x−2
in the dominated convergence theorem validate the existence of
                            ∞                                                        n
                                x−2 cos x dx               =          lim                x−2 cos x dx.
                        1                                            n→∞         1
We now consider two more substantive applications of the monotone and
dominated convergence theorems.
Example 3.5.2 Johann Bernoulli’s Integral
As example of delicate maneuvers in integration, consider the integral
                                 1                              1
                                      1
                                        dx        =                 e−x ln x dx
                             0       xx                     0
                                                                1 ∞
                                                                        (−x ln x)n
                                                  =                                dx
                                                            0       n=0
                                                                           n!
                                                            ∞               1
                                                               1
                                                  =                             (−x ln x)n dx .
                                                           n=0
                                                               n!       0

The reader will notice the application of the monotone convergence theorem
in passing from the second to the third line above. Further progress can be
made by applying the integration by parts result
                    1                                                               1
                                                                 n                             lnn−1 x
                        xm lnn x dx               =        −                            xm+1           dx
                0                                               m+1             0                 x
                                                                                    1
                                                              n
                                                  =        −                            xm lnn−1 x dx
                                                             m+1                0
3.5 More Advanced Topics in Integration                          69

recursively to evaluate
           1                                                       1
                                                n!                                               n!
               (−x ln x)n dx =                                         xn dx =                          .
       0                                     (n + 1)n          0                             (n + 1)n+1
The pleasant surprise
                                       1                       ∞
                                            1                        1
                                              dx     =
                                   0       xx                n=0
                                                                 (n + 1)n+1

emerges.
Example 3.5.3 Competing Definitions of the Gamma Function
The dominated convergence theorem allows us to derive Gauss’s represen-
tation
                                                             n!nz
                               Γ(z) =              lim
                                                  n→∞ z(z + 1) · · · (z + n)

of the gamma function from Euler’s representation
                                                             ∞
                                   Γ(z) =                        xz−1 e−x dx .
                                                         0

As students of statistics are apt to know from their exposure to the beta
distribution, repeated integration by parts and the fundamental theorem
of calculus show that
                          1
                                                                                n!
                              xz−1 (1 − x)n dx            =                                   .
                      0                                                z(z + 1) · · · (z + n)
The substitution rule yields
                          1                                                 n                         n
                                                                                                  y
                nz            xz−1 (1 − x)n dx            =                     y z−1 1 −                  dy .
                      0                                                 0                         n
Thus, it suffices to prove that
                      ∞                                                    n                          n
                                                                                              y
                          xz−1 e−x dx            =       lim                   y z−1 1 −                  dy .
                  0                                   n→∞              0                      n
Given the limit
                                                         y    n
                                           lim    1−                       =     e−y ,
                                       n→∞               n
we need an integrable function h(y) that dominates the nonnegative se-
quence
                                                                                         y    n
                               fn (y) =            1[0,n] (y)y z−1 1 −
                                                                                         n
70     3. The Gauge Integral

from above in order to apply the dominated convergence theorem. In light
of the inequality
                                  y n
                             1−          ≤ e−y ,
                                 n
the function h(y) = y z−1 e−y will serve.
   Finally, the gauge integral extends to multiple dimensions, where a ver-
sion of Fubini’s theorem holds for evaluating multidimensional integrals
via iterated integrals [278]. Consider a function f (x, y) defined over the
Cartesian product H × K of two multidimensional intervals H and K. The
intervals in question can be bounded or unbounded. If f (x, y) is integrable
over H × K, then Fubini’s theorem asserts that the integrals H f (x, y) dx
and K f (x, y) dy exist and can be integrated over the remaining variable
to give the full integral. In symbols,

           f (x, y) dx dy =                       f (x, y) dx dy =                                    f (x, y) dy dx .
     H×K                                 K    J                                         J         K
Conversely, if either iterated integral exists, one would like to conclude that
the full integral exists as well. This is true whenever f (x, y) is nonnega-
tive. Unfortunately, it is false in general, and two additional hypotheses
introduced by Tonelli are needed to rescue the situation. One hypothesis
is that f (x, y) is measurable. Measurability is a technical condition that
holds except for very pathological functions. The other hypothesis is that
|f (x, y)| ≤ g(x, y) for some nonnegative function g(x, y) for which the
iterated integral exists. This domination condition is shared with the dom-
inated convergence theorem and forces f (x, y) to be absolutely integrable.
                                                  ∞
Example 3.5.4 Evaluation of                       0
                                                      sinc(x) dx
According to Fubini’s theorem
                    n       nπ                                         nπ        n
                                 e−xy sin x dx dy     =                              e−xy sin x dy dx .           (3.12)
                0       0                                          0         0
The second of these iterated integrals
           nπ       n                                         nπ                             nπ
                                                                   sin x                                   sin x
                        e−xy sin x dy dx       =                         dx −                     e−nx           dx
       0       0                                          0          x                   0                   x
             ∞
tends to     0
                         as n tends to ∞ by a combination of Hake’s theorem
                sinc(x) dx
and the dominated convergence theorem. The inner integral of the left
iterated integral in (3.12) equals
           nπ                                                      nπ                                 nπ
                e−xy sin x dx           =    −e−xy cos x                    − ye−xy sin x
       0                                                           0                                  0
                                                          nπ
                                             − y2              e−xy sin x dx
                                                      0
                                                                                            nπ
                                                    −nπy
                                        =    1−e               cos nπ − y 2                       e−xy sin x dx
                                                                                        0
3.6 Problems   71

after two integrations by parts. It follows that
                       nπ
                                                           1 − e−nπy cos nπ
                               e−xy sin x dx           =                    .
                   0                                            1 + y2

Finally, application of the dominated convergence theorem gives
                                                                         ∞
                           n
                               1 − e−nπy cos nπ                                1
             lim                                dy            =                     dy
             n→∞       0            1 + y2                           0       1 + y2
                                                                    π
                                                              =       .
                                                                    2
Equating the limits of the right and left hand sides of the identity (3.12)
                                            ∞
therefore [278] yields the value of π/2 for 0 sinc(x) dx.



3.6 Problems
  1. Give an alternative proof of Cousin’s lemma by letting y be the supre-
     mum of the set of x ∈ [a, b] such that [a, x] possesses a δ-fine partition.

  2. Use Cousin’s lemma to prove that a continuous function f (x) defined
     on an interval [a, b] is uniformly continuous there [108]. (Hint: Given
       > 0 define a gauge δ(x) by the requirement that |f (y) − f (x)| < 1 2
     for all y ∈ [a, b] with |y − x| < 2δ(x).)

  3. A possibly discontinuous function f (x) has one-sided limits at each
     point x ∈ [a, b]. Show by Cousin’s lemma that f (x) is bounded on
     [a, b].

  4. Suppose f (x) has a nonnegative derivative f (x) throughout [a, b].
     Prove that f (x) is nondecreasing on [a, b]. Also prove that f (x) is
     constant on [a, b] if and only if f (x) = 0 for all x. (Hint: These yield
     easily to the fundamental theorem of calculus. Alternatively for the
     first assertion, consider the function

                                            f (x)      = f (x) + x

     for   > 0.)

  5. Using only the definition of the gauge integral, demonstrate that
                                        b                     −a
                                            f (t) dt   =           f (−t) dt
                                    a                        −b

     when either integral exists.
72      3. The Gauge Integral

     6. Based on the standard definition of the natural logarithm
                                                               y
                                                                   1
                                       ln y        =                 dx,
                                                           1       x
       prove that ln yz = ln y + ln z for all positive arguments y and z. Use
       this property to verify that ln y −1 = − ln y and that ln y r = r ln y for
       every rational number r.
     7. Apply Proposition 3.3.2 and demonstrate that every monotonic func-
        tion defined on an interval [a, b] is integrable on that interval.
     8. Let f (x) be a continuous real-valued function on [a, b]. Show that
        there exists c ∈ [a, b] with
                                  b
                                      f (x) dx         =       f (c)(b − a).
                              a


     9. In the Taylor expansion of Proposition 3.4.5, suppose f (k+1) (x) is
        continuous. Show that we can replace the remainder by
                                                   (x − y)k+1 (k+1)
                          Rk (x)         =                   f      (z)
                                                    (k + 1)!
       for some z between x and y.
 10. Suppose that f (x) is infinitely differentiable and that c and r are pos-
     itive numbers. If |f (k) (x)| ≤ ck!rk for all x near y and all nonnegative
     integers k, then use Proposition 3.4.5 to show that
                                                   ∞
                                                        f (k) (y)
                           f (x)        =                         (x − y)k
                                                           k!
                                                  k=0

       near y. Explicitly determine the infinite Taylor series expansion of the
       function f (x) = (1 + x)−1 around x = 0 and justify its convergence.
 11. Suppose the nonnegative continuous function f (x) satisfies
                                            b
                                                f (x) dx = 0.
                                        a

       Prove that f (x) is identically 0 on [a, b].
 12. Consider the function

                         f (x) =            x2 sin (x−2 ) x = 0
                                            0             x = 0.
                    1                                                      1
       Show that 0 f (x) dx = sin (1) and limt↓0 t |f (x)| dx = ∞. Hence,
       f (x) is integrable but not absolutely integrable on [0,1].
3.6 Problems   73

13. Prove that
                             ∞
                                                 β                 1           α+1
                                 xα e−x dx =                         Γ
                         0                                         β            β

    for α and β positive [82].

14. Justify the formula
                                   1                                     ∞
                                       ln(1 − x)                           1
                                                               = −            .
                               0           x                           n=1
                                                                           n2

15. Show that
                                   ∞
                                            xz−1
                                                  dx           = ζ(z)Γ(z),
                               0           ex − 1
                       ∞
    where ζ(z) =       n=1   n−z .

16. Prove that the functions
                                                      ∞
                                                            sin t
                       f (x)         =                             dt
                                                  1       x2  + t2
                                                      ∞
                       g(x)          =                    e−xt cos t dt,          x > 0,
                                                  0

    are continuous.

17. Let fn (x) be a sequence of integrable functions on [a, b] that converges
    uniformly to f (x). Demonstrate that f (x) is integrable and satisfies
                                           b                               b
                        lim                    fn (x) dx       =               f (x) dx .
                       n→∞             a                               a

    (Hints: For    > 0 small take n large enough so that

            fn (x) −                           ≤ f (x) ≤ fn (x) +
                       2(b − a)                                                     2(b − a)

    for all x.)

18. Let p and q be positive integers. Justify the series expansion
                                 1                                 ∞
                                        xp−1                           (−1)n
                                              dx =
                             0         1 + xq                      n=0
                                                                       p + nq

    by the monotone convergence theorem. Be careful since the series
    does not converge absolutely [278].
74   3. The Gauge Integral

 19. Suppose f (x) is a continuous function on R. Demonstrate that the
     sequence
                                                    n−1
                                               1                        k
                         fn (x)         =                 f       x+
                                               n                        n
                                                    k=0

     converges uniformly to a continuous function on every finite interval
     [a, b] [69].
 20. Prove that
                               1
                                   xb − xa                         b+1
                                           dx          = ln
                           0         ln x                          a+1

     for 0 < a < b [278] by showing that both sides equal the double
     integral

                                                   xy dx dy .
                                   [0,1]×[a,b]


 21. Integrate the function

                                                       y 2 − x2
                               f (x, y) =
                                                      (x2 + y 2 )2

     over the unit square [0, 1]× [0, 1]. Show that the two iterated integrals
     disagree, and explain why Fubini’s theorem fails.
                                                              2                 2
 22. Suppose the two partial derivatives ∂x∂∂x2 f (x) and ∂x∂∂x1 f (x) exist
                                              1               2
     and are continuous in a neighborhood of a point y ∈ R2 . Show that
     they are equal at the point. (Hints: If they are not equal, take a small
     box around the point where their difference has constant sign. Now
     apply Fubini’s theorem.)
 23. Demonstrate that
                                        ∞                         √
                                               2                    π
                                            e−x dx        =
                                    0                              2
                                                                        2   2
     by evaluating the integral of f (y) = y2 e−(1+y1 )y2 over the rectangle
     (0, ∞) × (0, ∞).

Más contenido relacionado

La actualidad más candente

Quantum Computation and the Stabilizer Formalism for Error Correction
Quantum Computation and the Stabilizer Formalism for Error CorrectionQuantum Computation and the Stabilizer Formalism for Error Correction
Quantum Computation and the Stabilizer Formalism for Error CorrectionDaniel Bulhosa Solórzano
 
Important Cuts
Important CutsImportant Cuts
Important CutsASPAK2014
 
Afm chapter 4 powerpoint
Afm chapter 4 powerpointAfm chapter 4 powerpoint
Afm chapter 4 powerpointvolleygurl22
 
Spectral sum rules for conformal field theories
Spectral sum rules for conformal field theoriesSpectral sum rules for conformal field theories
Spectral sum rules for conformal field theoriesSubham Dutta Chowdhury
 
Limits
LimitsLimits
Limitssarcia
 
Reciprocity Law For Flat Conformal Metrics With Conical Singularities
Reciprocity Law For Flat Conformal Metrics With Conical SingularitiesReciprocity Law For Flat Conformal Metrics With Conical Singularities
Reciprocity Law For Flat Conformal Metrics With Conical SingularitiesLukasz Obara
 
Jensen's inequality, EM 알고리즘
Jensen's inequality, EM 알고리즘 Jensen's inequality, EM 알고리즘
Jensen's inequality, EM 알고리즘 Jungkyu Lee
 
Inversion Theorem for Generalized Fractional Hilbert Transform
Inversion Theorem for Generalized Fractional Hilbert TransformInversion Theorem for Generalized Fractional Hilbert Transform
Inversion Theorem for Generalized Fractional Hilbert Transforminventionjournals
 
ABC based on Wasserstein distances
ABC based on Wasserstein distancesABC based on Wasserstein distances
ABC based on Wasserstein distancesChristian Robert
 
Nies cuny describing_finite_groups
Nies cuny describing_finite_groupsNies cuny describing_finite_groups
Nies cuny describing_finite_groupsAndre Nies
 

La actualidad más candente (19)

Bd32360363
Bd32360363Bd32360363
Bd32360363
 
Quantum Computation and the Stabilizer Formalism for Error Correction
Quantum Computation and the Stabilizer Formalism for Error CorrectionQuantum Computation and the Stabilizer Formalism for Error Correction
Quantum Computation and the Stabilizer Formalism for Error Correction
 
Approx
ApproxApprox
Approx
 
Important Cuts
Important CutsImportant Cuts
Important Cuts
 
Afm chapter 4 powerpoint
Afm chapter 4 powerpointAfm chapter 4 powerpoint
Afm chapter 4 powerpoint
 
Lesson 5: Continuity
Lesson 5: ContinuityLesson 5: Continuity
Lesson 5: Continuity
 
Spectral sum rules for conformal field theories
Spectral sum rules for conformal field theoriesSpectral sum rules for conformal field theories
Spectral sum rules for conformal field theories
 
Lecture26
Lecture26Lecture26
Lecture26
 
Limits
LimitsLimits
Limits
 
Quantum Noise and Error Correction
Quantum Noise and Error CorrectionQuantum Noise and Error Correction
Quantum Noise and Error Correction
 
10.1.1.226.4381
10.1.1.226.438110.1.1.226.4381
10.1.1.226.4381
 
Stochastic Processes Homework Help
Stochastic Processes Homework Help Stochastic Processes Homework Help
Stochastic Processes Homework Help
 
Reciprocity Law For Flat Conformal Metrics With Conical Singularities
Reciprocity Law For Flat Conformal Metrics With Conical SingularitiesReciprocity Law For Flat Conformal Metrics With Conical Singularities
Reciprocity Law For Flat Conformal Metrics With Conical Singularities
 
Jensen's inequality, EM 알고리즘
Jensen's inequality, EM 알고리즘 Jensen's inequality, EM 알고리즘
Jensen's inequality, EM 알고리즘
 
Thesis 6
Thesis 6Thesis 6
Thesis 6
 
Inversion Theorem for Generalized Fractional Hilbert Transform
Inversion Theorem for Generalized Fractional Hilbert TransformInversion Theorem for Generalized Fractional Hilbert Transform
Inversion Theorem for Generalized Fractional Hilbert Transform
 
Pairing scott
Pairing scottPairing scott
Pairing scott
 
ABC based on Wasserstein distances
ABC based on Wasserstein distancesABC based on Wasserstein distances
ABC based on Wasserstein distances
 
Nies cuny describing_finite_groups
Nies cuny describing_finite_groupsNies cuny describing_finite_groups
Nies cuny describing_finite_groups
 

Destacado

A course on mathematical logic
A course on mathematical logicA course on mathematical logic
A course on mathematical logicSpringer
 
Periodic differential operators
Periodic differential operatorsPeriodic differential operators
Periodic differential operatorsSpringer
 
Handbook of children with special health care needs
Handbook of children with special health care needsHandbook of children with special health care needs
Handbook of children with special health care needsSpringer
 
Regularity and complexity in dynamical systems
Regularity and complexity in dynamical systemsRegularity and complexity in dynamical systems
Regularity and complexity in dynamical systemsSpringer
 
Complex models and computational methods in statistics
Complex models and computational methods in statisticsComplex models and computational methods in statistics
Complex models and computational methods in statisticsSpringer
 
Applied bayesian statistics
Applied bayesian statisticsApplied bayesian statistics
Applied bayesian statisticsSpringer
 
Complex analysis and differential equation
Complex analysis and differential equationComplex analysis and differential equation
Complex analysis and differential equationSpringer
 
Organizational learning
Organizational learningOrganizational learning
Organizational learningSpringer
 

Destacado (8)

A course on mathematical logic
A course on mathematical logicA course on mathematical logic
A course on mathematical logic
 
Periodic differential operators
Periodic differential operatorsPeriodic differential operators
Periodic differential operators
 
Handbook of children with special health care needs
Handbook of children with special health care needsHandbook of children with special health care needs
Handbook of children with special health care needs
 
Regularity and complexity in dynamical systems
Regularity and complexity in dynamical systemsRegularity and complexity in dynamical systems
Regularity and complexity in dynamical systems
 
Complex models and computational methods in statistics
Complex models and computational methods in statisticsComplex models and computational methods in statistics
Complex models and computational methods in statistics
 
Applied bayesian statistics
Applied bayesian statisticsApplied bayesian statistics
Applied bayesian statistics
 
Complex analysis and differential equation
Complex analysis and differential equationComplex analysis and differential equation
Complex analysis and differential equation
 
Organizational learning
Organizational learningOrganizational learning
Organizational learning
 

Similar a Optimization (20)

L 4 4
L 4 4L 4 4
L 4 4
 
The Fundamental theorem of calculus
The Fundamental theorem of calculus The Fundamental theorem of calculus
The Fundamental theorem of calculus
 
1543 integration in mathematics b
1543 integration in mathematics b1543 integration in mathematics b
1543 integration in mathematics b
 
azEssay2
azEssay2azEssay2
azEssay2
 
Mth3101 Advanced Calculus Chapter 1
Mth3101 Advanced Calculus Chapter 1Mth3101 Advanced Calculus Chapter 1
Mth3101 Advanced Calculus Chapter 1
 
publish paper
publish paperpublish paper
publish paper
 
Lar calc10 ch01_sec4
Lar calc10 ch01_sec4Lar calc10 ch01_sec4
Lar calc10 ch01_sec4
 
deepak real.ppt
deepak real.pptdeepak real.ppt
deepak real.ppt
 
The integral
The integralThe integral
The integral
 
On the Fixed Point Extension Results in the Differential Systems of Ordinary ...
On the Fixed Point Extension Results in the Differential Systems of Ordinary ...On the Fixed Point Extension Results in the Differential Systems of Ordinary ...
On the Fixed Point Extension Results in the Differential Systems of Ordinary ...
 
SURF 2012 Final Report(1)
SURF 2012 Final Report(1)SURF 2012 Final Report(1)
SURF 2012 Final Report(1)
 
01_AJMS_185_19_RA.pdf
01_AJMS_185_19_RA.pdf01_AJMS_185_19_RA.pdf
01_AJMS_185_19_RA.pdf
 
01_AJMS_185_19_RA.pdf
01_AJMS_185_19_RA.pdf01_AJMS_185_19_RA.pdf
01_AJMS_185_19_RA.pdf
 
MEAN VALUE THEOREM
MEAN VALUE THEOREMMEAN VALUE THEOREM
MEAN VALUE THEOREM
 
2_GLMs_printable.pdf
2_GLMs_printable.pdf2_GLMs_printable.pdf
2_GLMs_printable.pdf
 
Treewidth and Applications
Treewidth and ApplicationsTreewidth and Applications
Treewidth and Applications
 
StewartCalcET7e_05_04.ppt
StewartCalcET7e_05_04.pptStewartCalcET7e_05_04.ppt
StewartCalcET7e_05_04.ppt
 
Logistics Management Assignment Help
Logistics Management Assignment Help Logistics Management Assignment Help
Logistics Management Assignment Help
 
AJMS_403_22.pdf
AJMS_403_22.pdfAJMS_403_22.pdf
AJMS_403_22.pdf
 
Chap4_Sec2.ppt
Chap4_Sec2.pptChap4_Sec2.ppt
Chap4_Sec2.ppt
 

Más de Springer

The chemistry of the actinide and transactinide elements (set vol.1 6)
The chemistry of the actinide and transactinide elements (set vol.1 6)The chemistry of the actinide and transactinide elements (set vol.1 6)
The chemistry of the actinide and transactinide elements (set vol.1 6)Springer
 
Transition metal catalyzed enantioselective allylic substitution in organic s...
Transition metal catalyzed enantioselective allylic substitution in organic s...Transition metal catalyzed enantioselective allylic substitution in organic s...
Transition metal catalyzed enantioselective allylic substitution in organic s...Springer
 
Total synthesis of natural products
Total synthesis of natural productsTotal synthesis of natural products
Total synthesis of natural productsSpringer
 
Solid state nmr
Solid state nmrSolid state nmr
Solid state nmrSpringer
 
Mass spectrometry
Mass spectrometryMass spectrometry
Mass spectrometrySpringer
 
Higher oxidation state organopalladium and platinum
Higher oxidation state organopalladium and platinumHigher oxidation state organopalladium and platinum
Higher oxidation state organopalladium and platinumSpringer
 
Principles and applications of esr spectroscopy
Principles and applications of esr spectroscopyPrinciples and applications of esr spectroscopy
Principles and applications of esr spectroscopySpringer
 
Inorganic 3 d structures
Inorganic 3 d structuresInorganic 3 d structures
Inorganic 3 d structuresSpringer
 
Field flow fractionation in biopolymer analysis
Field flow fractionation in biopolymer analysisField flow fractionation in biopolymer analysis
Field flow fractionation in biopolymer analysisSpringer
 
Thermodynamics of crystalline states
Thermodynamics of crystalline statesThermodynamics of crystalline states
Thermodynamics of crystalline statesSpringer
 
Theory of electroelasticity
Theory of electroelasticityTheory of electroelasticity
Theory of electroelasticitySpringer
 
Tensor algebra and tensor analysis for engineers
Tensor algebra and tensor analysis for engineersTensor algebra and tensor analysis for engineers
Tensor algebra and tensor analysis for engineersSpringer
 
Springer handbook of nanomaterials
Springer handbook of nanomaterialsSpringer handbook of nanomaterials
Springer handbook of nanomaterialsSpringer
 
Shock wave compression of condensed matter
Shock wave compression of condensed matterShock wave compression of condensed matter
Shock wave compression of condensed matterSpringer
 
Polarization bremsstrahlung on atoms, plasmas, nanostructures and solids
Polarization bremsstrahlung on atoms, plasmas, nanostructures and solidsPolarization bremsstrahlung on atoms, plasmas, nanostructures and solids
Polarization bremsstrahlung on atoms, plasmas, nanostructures and solidsSpringer
 
Nanostructured materials for magnetoelectronics
Nanostructured materials for magnetoelectronicsNanostructured materials for magnetoelectronics
Nanostructured materials for magnetoelectronicsSpringer
 
Nanobioelectrochemistry
NanobioelectrochemistryNanobioelectrochemistry
NanobioelectrochemistrySpringer
 
Modern theory of magnetism in metals and alloys
Modern theory of magnetism in metals and alloysModern theory of magnetism in metals and alloys
Modern theory of magnetism in metals and alloysSpringer
 
Mechanical behaviour of materials
Mechanical behaviour of materialsMechanical behaviour of materials
Mechanical behaviour of materialsSpringer
 

Más de Springer (20)

The chemistry of the actinide and transactinide elements (set vol.1 6)
The chemistry of the actinide and transactinide elements (set vol.1 6)The chemistry of the actinide and transactinide elements (set vol.1 6)
The chemistry of the actinide and transactinide elements (set vol.1 6)
 
Transition metal catalyzed enantioselective allylic substitution in organic s...
Transition metal catalyzed enantioselective allylic substitution in organic s...Transition metal catalyzed enantioselective allylic substitution in organic s...
Transition metal catalyzed enantioselective allylic substitution in organic s...
 
Total synthesis of natural products
Total synthesis of natural productsTotal synthesis of natural products
Total synthesis of natural products
 
Solid state nmr
Solid state nmrSolid state nmr
Solid state nmr
 
Mass spectrometry
Mass spectrometryMass spectrometry
Mass spectrometry
 
Higher oxidation state organopalladium and platinum
Higher oxidation state organopalladium and platinumHigher oxidation state organopalladium and platinum
Higher oxidation state organopalladium and platinum
 
Principles and applications of esr spectroscopy
Principles and applications of esr spectroscopyPrinciples and applications of esr spectroscopy
Principles and applications of esr spectroscopy
 
Inorganic 3 d structures
Inorganic 3 d structuresInorganic 3 d structures
Inorganic 3 d structures
 
Field flow fractionation in biopolymer analysis
Field flow fractionation in biopolymer analysisField flow fractionation in biopolymer analysis
Field flow fractionation in biopolymer analysis
 
Thermodynamics of crystalline states
Thermodynamics of crystalline statesThermodynamics of crystalline states
Thermodynamics of crystalline states
 
Theory of electroelasticity
Theory of electroelasticityTheory of electroelasticity
Theory of electroelasticity
 
Tensor algebra and tensor analysis for engineers
Tensor algebra and tensor analysis for engineersTensor algebra and tensor analysis for engineers
Tensor algebra and tensor analysis for engineers
 
Springer handbook of nanomaterials
Springer handbook of nanomaterialsSpringer handbook of nanomaterials
Springer handbook of nanomaterials
 
Shock wave compression of condensed matter
Shock wave compression of condensed matterShock wave compression of condensed matter
Shock wave compression of condensed matter
 
Polarization bremsstrahlung on atoms, plasmas, nanostructures and solids
Polarization bremsstrahlung on atoms, plasmas, nanostructures and solidsPolarization bremsstrahlung on atoms, plasmas, nanostructures and solids
Polarization bremsstrahlung on atoms, plasmas, nanostructures and solids
 
Nanostructured materials for magnetoelectronics
Nanostructured materials for magnetoelectronicsNanostructured materials for magnetoelectronics
Nanostructured materials for magnetoelectronics
 
Nanobioelectrochemistry
NanobioelectrochemistryNanobioelectrochemistry
Nanobioelectrochemistry
 
Modern theory of magnetism in metals and alloys
Modern theory of magnetism in metals and alloysModern theory of magnetism in metals and alloys
Modern theory of magnetism in metals and alloys
 
Mechanical behaviour of materials
Mechanical behaviour of materialsMechanical behaviour of materials
Mechanical behaviour of materials
 
Magnonics
MagnonicsMagnonics
Magnonics
 

Optimization

  • 1. 3 The Gauge Integral 3.1 Introduction Much of calculus deals with the interplay between differentiation and integration. The antiquated term “antidifferentiation” emphasizes the fact that differentiation and integration are inverses of one another. We will take it for granted that readers are acquainted with the mechanics of in- tegration. The current chapter develops just enough integration theory to make our development of differentiation in Chap. 4 and the calculus of variations in Chap. 17 respectable. It is only fair to warn readers that in other chapters a few applications to probability and statistics will assume familiarity with properties of the expectation operator not covered here. The first successful effort to put integration on a rigorous basis was un- dertaken by Riemann. In the early twentieth century, Lebesgue defined a more sophisticated integral that addresses many of the limitations of the Riemann integral. However, even Lebesgue’s integral has its defects. In the past few decades, mathematicians such as Henstock and Kurzweil have expanded the definition of integration on the real line to include a wider variety of functions. The new integral emerging from these in- vestigations is called the gauge integral or generalized Riemann integral [7, 68, 108, 193, 250, 255, 278]. The gauge integral subsumes the Riemann integral, the Lebesgue integral, and the improper integrals met in tradi- tional advanced calculus courses. In contrast to the Lebesgue integral, the integrands of the gauge integral are not necessarily absolutely integrable. K. Lange, Optimization, Springer Texts in Statistics 95, 53 DOI 10.1007/978-1-4614-5838-8 3, © Springer Science+Business Media New York 2013
  • 2. 54 3. The Gauge Integral It would take us too far afield to develop the gauge integral in full generality. Here we will rest content with proving some of its elementary properties. One of the advantages of the gauge integral is that many theo- rems hold with fewer qualifications. The fundamental theorem of calculus is a case in point. The commonly stated version of the fundamental theorem concerns a differentiable function f (x) on an interval [a, b]. As all students of calculus know, b f (x) dx = f (b) − f (a). a Although this version is true for the gauge integral, it does not hold for the Lebesgue integral because the mere fact that f (x) exists throughout [a, b] does not guarantee that it is Lebesgue integrable. This quick description of the gauge integral is not intended to imply that the gauge integral is uniformly superior to the Lebesgue integral and its extensions. Certainly, probability theory would be severely handicapped without the full flexibility of modern measure theory. Furthermore, the ad- vanced theory of the gauge integral is every bit as difficult as the advanced theory of the Lebesgue integral. For pedagogical purposes, however, one can argue that a student’s first exposure to the theory of integration should fea- ture the gauge integral. As we shall see, many of the basic properties of the gauge integral flow directly from its definition. As an added dividend, gauge functions provide an alternative approach to some of the material of Chap. 2. 3.2 Gauge Functions and δ-Fine Partitions The gauge integral is defined through gauge functions. A gauge function is nothing more than a positive function δ(t) defined on a finite interval [a, b]. In approximating the integral of a function f (t) over [a, b] by a finite Riemann sum, it is important to sample the function most heavily in those regions where it changes most rapidly. Now by a Riemann sum we mean a sum n−1 S(f, π) = f (ti )(si+1 − si ), i=0 where the mesh points a = s0 < s1 < · · · < sn = b form a partition π of [a, b], and the tags ti are chosen so that ti ∈ [si , si+1 ]. If δ(ti ) measures the rapidity of change of f (t) near ti , then it makes sense to take δ(t) small in regions of rapid change and to force si and si+1 to belong to the interval (ti − δ(ti ), ti + δ(ti )). A tagged partition with this property is called a δ- fine partition. Our first proposition relieves our worry that δ-fine partitions exist.
  • 3. 3.2 Gauge Functions and δ-Fine Partitions 55 Proposition 3.2.1 (Cousin’s Lemma) For every gauge δ(t) on a finite interval [a, b] there is a δ-fine partition. Proof: Assume that [a, b] lacks a δ-fine partition. Since we can construct a δ-fine partition of [a, b] by appending a δ-fine partition of the half-interval [(a + b)/2, b] to a δ-fine partition of the half-interval [a, (a + b)/2], it fol- lows that either [a, (a + b)/2] or [(a + b)/2, b] lacks a δ-fine partition. As in Example 2.3.1, we choose one of the half-intervals based on this failure and continue bisecting. This creates a nested sequence of intervals [ai , bi ] con- verging to a point x. If i is large enough, then [ai , bi ] ⊂ (x − δ(x), x + δ(x)), and the interval [ai , bi ] with tag x is a δ-fine partition of itself. This con- tradicts the choice of [ai , bi ] and the assumption that the original interval [a, b] lacks a δ-fine partition. Before launching into our treatment of the gauge integral, we pause to gain some facility with gauge functions [108]. Here are three examples that illustrate their value. Example 3.2.1 A Gauge Proof of Weierstrass’ Theorem Consider a real-valued continuous function f (t) with domain [a, b]. Suppose that f (t) does not attain its supremum on [a, b]. Then for each t there exists a point x ∈ [a, b] with f (t) < f (x). By continuity there exists δ(t) > 0 such that f (y) < f (x) for all y ∈ [a, b] with |y − t| < δ(t). Using δ(t) as a gauge, select a δ-fine partition a = s0 < s1 < · · · < sn = b with tags ti ∈ [si , si+1 ] and designated points xi satisfying f (ti ) < f (xi ). Let xmax be the point xi having the largest value f (xi ). Because xmax lies in some interval [si , si+1 ], we have f (xmax ) < f (xi ). This contradiction discredits our assumption that f (x) does not attain its supremum. A similar argument applies to the infimum. Example 3.2.2 A Gauge Proof of the Heine-Borel Theorem One can use Cousin’s lemma to prove the Heine-Borel Theorem on the real line [278]. This theorem states that if C is a compact set contained in the union ∪α Oα of a collection of open sets Oα , then C is actually contained in the union of a finite number of the Oα . Suppose C ⊂ [a, b]. Define a gauge δ(t) so that the interval (t − δ(t), t + δ(t)) does not intersect C when t ∈ C and (t − δ(t), t + δ(t)) is contained in some Oα when t ∈ C. Based on δ(t), select a δ-fine partition a = s0 < s1 < · · · < sn = b with tags ti ∈ [si , si+1 ]. By definition C is contained in the union ∪ti ∈C Ui , where Ui is the set Oα covering ti . The Heine-Borel theorem extends to compact sets in Rn . Example 3.2.3 A Gauge Proof of the Intermediate Value Theorem Under the assumption of the previous example, let c be an number strictly between f (a) and f (b). If we assume that there is no t ∈ [a, b] with f (t) = c, then there exists a positive number δ(t) such that either f (x) < c for all
  • 4. 56 3. The Gauge Integral x ∈ [a, b] with |x − t| < δ(t) or f (x) > c for all x ∈ [a, b] with |x − t| < δ(t). We now select a δ-fine partition a = s0 < s1 < · · · < sn = b and observe that throughout each interval [si , si+1 ] either f (t) < c or f (t) > c. If to start f (s0 ) = f (a) < c, then f (s1 ) < c, which implies f (s2 ) < c and so forth until we get to f (sn ) = f (b) < c. This contradicts the assumption that c lies strictly between f (a) and f (b). With minor differences, the same proof works when f (a) > c. In preparation for our next example and for the fundamental theorem of calculus later in this chapter, we must define derivatives. A real-valued function f (t) defined on an interval [a, b] possesses a derivative f (c) at c ∈ [a, b] provided the limit f (t) − f (c) lim = f (c) (3.1) t→c t−c exists. At the endpoints a and b, the limit is necessarily one sided. Tak- ing a sequential view of convergence, definition (3.1) means that for every sequence tm converging to c we must have f (tm ) − f (c) lim = f (c). m→∞ tm − c In calculus, we learn the following rules for computing derivatives: Proposition 3.2.2 If f (t) and g(t) are differentiable functions on (a, b), then αf (t) + βg(t) = αf (t) + βg (t) f (t)g(t) = f (t)g(t) + f (t)g (t) 1 f (t) = − . f (t) f (t)2 In the third formula we must assume f (t) = 0. Finally, if g(t) maps into the domain of f (t), then the functional composition f ◦ g(t) has derivative [f ◦ g(t)] = f ◦ g(t)g (t). Proof: We will prove the above sum, product, quotient, and chain rules in a broader context in Chap. 4. Our proofs will not rely on integration. Example 3.2.4 Strictly Increasing Functions Let f (t) be a differentiable function on [c, d] with strictly positive derivative. We now show that f (t) is strictly increasing. For each t ∈ [c, d] there exists δ(t) > 0 such that f (x) − f (t) > 0 (3.2) x−t
  • 5. 3.3 Definition and Basic Properties of the Integral 57 for all x ∈ [a, b] with |x − t| < δ(t). According to Proposition 3.2.1, for any two points a < b from [c, d], there exists a δ-fine partition a = s0 < s1 < · · · < sn = b of [a, b] with tags ti ∈ [si , si+1 ]. In view of inequality (3.2), at least one of the two inequalities f (si ) ≤ f (ti ) ≤ f (si+1 ) must be strict. Thus, the telescoping sum n−1 f (b) − f (a) = [f (si+1 ) − f (si )] i=0 must be positive. 3.3 Definition and Basic Properties of the Integral With later applications in mind, it will be convenient to define the gauge integral for vector-valued functions f (x) : [a, b] → Rn . In this context, f (x) is said to have integral I if for every > 0 there exists a gauge δ(x) on [a, b] such that S(f, π) − I < (3.3) for all δ-fine partitions π. Our first order of business is to check that the integral is unique whenever it exists. Thus, suppose that the vector J is a second possible value of the integral. Given > 0 choose gauges δI (x) and δJ (x) leading to inequality (3.3). The minimum δ(x) = min{δI (x), δJ (x)} is also a gauge, and any partition π that is δ-fine is also δI and δJ -fine. Hence, I −J ≤ I − S(f, π) + S(f, π) − J < 2 . Since is arbitrary, J = I. One can also define f (x) to be integrable if its Riemann sums are Cauchy in an appropriate sense. Proposition 3.3.1 (Cauchy criterion) A function f (x) : [a, b] → Rn is integrable if and only if for every > 0 there exists a gauge δ(x) > 0 such that S(f, π1 ) − S(f, π2 ) < (3.4) for any two δ-fine partitions π1 and π2 . Proof: It is obvious that the Cauchy criterion is necessary for integrability. To show that it is sufficient, consider the sequence m = m−1 and compat- ible sequence of gauges δm (x) determined by condition (3.4). We can force
  • 6. 58 3. The Gauge Integral the constraints δm (x) ≤ δm−1 (x) to hold by inductively replacing δm (x) by min{δm−1 (x), δm (x)} whenever needed. Now select a δm -fine partition πm for each m. Because the gauge sequence δm (x) is decreasing, every partition π that is δm -fine is also δm−1 -fine. Hence, the sequence of Riemann sums S(f, πm ) is Cauchy and has a limit I satisfying S(f, πm ) − I ≤ m−1 . Finally, given the potential integral I, we take an arbitrary > 0 and choose m so that m−1 < . If π is δm -fine, then the inequality S(f, π) − I ≤ S(f, π) − S(f, πm ) + S(f, πm ) − I < 2 . completes the proof. For two integrable functions f (x) and g(x), the gauge integral inherits the linearity property b b b [αf (x) + βg(x)] dx = α f (x) dx + β g(x) dx a a a from its approximating Riemann sums. To prove this fact, take > 0 and choose gauges δf (x) and δg (x) so that b b S(f, πf ) − f (x) dx < , S(f, πg ) − g(x) dx < a a whenever πf is δf -fine and πg is δg -fine. If the tagged partition π is δ-fine for the gauge δ(x) = min{δf (x), δg (x)}, then b b S(αf + βg, π) − α f (x) dx − β g(x) dx a a b b ≤ |α| S(f, π) − f (x) dx + |β| S(g, π) − g(x) dx a a ≤ (|α| + |β|) . The gauge integral also inherits obvious order properties. For example, b a f (x) dx ≥ 0 whenever the integrand f (x) ≥ 0 for all x ∈ [a, b]. In this b case, the inequality |S(f, π) − a f (x) dx| < implies b 0 ≤ S(f, π) ≤ f (x) dx + . a Since can be made arbitrarily small for f (x) integrable, it follows that b a f (x) dx ≥ 0. This nonnegativity property translates into the order property b b f (x) dx ≤ g(x) dx a a
  • 7. 3.3 Definition and Basic Properties of the Integral 59 for two integrable functions f (x) ≤ g(x). In particular, when both f (x) and |f (x)| are both integrable, we have b b f (x) dx ≤ |f (x)| dx . a a For vector-valued functions, the analogous rule b b f (x) dx ≤ f (x) dx (3.5) a a is also inherited from the approximating Riemann sums. The reader can easily supply the proof using the triangle inequality of the Euclidean norm. It does not take much imagination to extend the definition of the gauge integral to matrix-valued functions, and inequality (3.5) applies in this setting as well. One of the nicest features of the gauge integral is that one can perturb an integrable function at a countable number of points without changing the value of its integral. This property fails for the Riemann integral but is exhibited by the Lebesgue integral. To validate the property, it suffices to prove that a function that equals 0 except at a countable number of points has integral 0. Suppose f (x) is such a function with exceptional points x1 , x2 , . . . and corresponding exceptional values f 1 , f 2 , . . .. We now define a gauge δ(x) with value 1 on the nonexceptional points and values δ(xj ) = 2j+2 [ f j + 1] at the exceptional points. If π is a δ-fine partition, then xj can serve as a tag for at most two intervals [si , si+1 ] of π and each such interval has length less than 2δ(xj ). It follows that ∞ 2 1 S(f, π) ≤ 2 f (xj ) j+2 [ f ≤ = j 2 j + 1] j=1 2j b and therefore that a f (x) dx = 0. In practice, the interval additivity rule c b c f (x) dx = f (x) dx + f (x) dx (3.6) a a b is obviously desirable. There are three separate issues in proving it. First, given the existence of the integral over [a, c], do the integrals over [a, b] and [b, c] exist? Second, if the integrals over [a, b] and [b, c] exist, does the integral over [a, c] exist? Third, if the integrals over [a, b] and [b, c] exist, are they additive? The first question is best approached through Proposition 3.3.1. For > 0 there exists a gauge δ(x) such that S(f, π1 ) − S(f, π2 ) <
  • 8. 60 3. The Gauge Integral for any two δ-fine partitions π1 and π2 of [a, c]. Given δ(x), take any two δ-fine partitions γ1 and γ2 of [a, b] and a single δ-fine partition ω of [b, c]. The concatenated partitions γ1 ∪ ω and γ2 ∪ ω are δ-fine throughout [a, c] and satisfy S(f, γ1 ) − S(f, γ2 ) = S(f, γ1 ∪ ω) − S(f, γ2 ∪ ω) < . According to the Cauchy criterion, the integral over [a, b] therefore exists. A similar argument implies that the integral over [b, c] also exists. Finally, the combination of these results shows that the integral exists over any interval [u, v] contained within [a, b]. For the converse, choose gauges δ1 (x) on [a, b] and δ2 (x) on [b, c] so that b c S(f, γ) − f (x) dx < , S(f, ω) − f (x) dx < a b for any δ1 -fine partition γ of [a, b] and any δ2 -fine partition ω of [b, c]. The concatenated partition π = γ ∪ ω satisfies b c S(f, π) − f (x) dx − f (x) dx a b b c ≤ S(f, γ) − f (x) + S(f, ω) − f (x) dx (3.7) a b < 2 because the Riemann sums satisfy S(f, π) = S(f, γ)+S(f, ω). This suggests defining a gauge δ(x) equal to δ1 (x) on [a, b] and equal to δ2 (x) on [b, c]. The problem with this tactic is that some partitions of [a, c] do not split at b. However, we can ensure a split by redefining δ(x) by ˜ min{δ1 (b), δ2 (b)} x=b δ(x) = min{δ(x), 1 |x − b|} 2 x = b. This forces b to be the tag of its assigned interval, and we can if needed split this interval at b and retain b as tag of both subintervals. With δ(x) amended in this fashion, any δ-fine partition π can be viewed as a con- catenated partition γ ∪ ω splitting at b. As such π obeys inequality (3.7). This argument simultaneously proves that the integral over [a, c] exists and satisfies the additivity property (3.6) If the function f (x) is vector-valued with n components, then the in- tegrability of f (x) should imply the integrability of each its components fi (x). Furthermore, we should be able to write ⎛ b ⎞ b f (x) dx a 1 ⎜ . ⎟ f (x) dx = ⎝ . . ⎠. a b a fn (x) dx
  • 9. 3.3 Definition and Basic Properties of the Integral 61 Conversely, if its components are integrable, then f (x) should be integrable as well. The inequalities n √ S(f, π) − I ≤ |S(fi , π) − Ii | ≤ n S(f, π) − I . i=1 based on Example 2.5.6 and Problem 3 of Chap. 2 are instrumental in proving this logical equivalence. Given that we can integrate component by component, for the remainder of this chapter we will deal exclusively with real-valued functions. We have not actually shown that any function is integrable. The most obvious possibility is a constant. Fortunately, it is trivial to demonstrate that b c dx = c(b − a). a Step functions are one rung up the hierarchy of functions. If n−1 f (x) = ci 1(si ,si+1 ] (x) i=0 for a = s0 < s1 < · · · < sn = b, then our nascent theory allows us to evaluate b n−1 si+1 n−1 f (x) dx = ci dx = ci (si+1 − si ). a i=0 si i=0 This fact and the next technical proposition turn out to be the key to showing that continuous functions are integrable. Proposition 3.3.2 Let f (x) be a function with domain [a, b]. Suppose for every > 0 there exist two integrable functions g(x) and h(x) satisfying g(x) ≤ f (x) ≤ h(x) for all x and b b h(x) dx ≤ g(x) dx + . a a Then f (x) is integrable. Proof: For > 0, choose gauges δg (x) and δh (x) on [a, b] so that b b S(g, πg ) − g(x) dx < , S(h, πh ) − h(x) dx < a a
  • 10. 62 3. The Gauge Integral for any δg -fine partition πg and any δh -fine partition πh . If π is a δ-fine partition for δ(x) = min{δg (x), δh (x)}, then the inequalities b g(x) dx − < S(g, π) a ≤ S(f, π) ≤ S(h, π) b < h(x) + a b ≤ g(x) dx + 2 a trap S(f, π) in an interval of length 3 . Because the Riemann sum S(f, γ) for any other δ-fine partition γ is trapped in the same interval, the integral of f (x) exists by the Cauchy criterion. Proposition 3.3.3 Every continuous function f (x) on [a, b] is integrable. Proof: In view of the uniform continuity of f (x) on [a, b], for every > 0 there exists a δ > 0 with |f (x) − f (y)| < when |x − y| < δ. For the constant gauge δ(x) = δ and a corresponding δ-fine partition π with mesh points s0 , . . . , sn , let mi be the minimum and Mi be the maximum of f (x) on [si , si+1 ]. The step functions n n g(x) = mi 1(si ,si+1 ] (x), h(x) = Mi 1(si ,si+1 ] (x) i=1 i=1 then satisfy g(x) ≤ f (x) ≤ h(x) except at the single point a. Furthermore, b b n h(x) dx − g(x) dx ≤ (si+1 − si ) a a i=1 = (b − a). Application of Proposition 3.3.2 now completes the proof. 3.4 The Fundamental Theorem of Calculus The fundamental theorem of calculus divides naturally into two parts. For the gauge integral, the first and more difficult part is easily proved by invoking what is called the straddle inequality. Let f (x) be differentiable at the point t ∈ [a, b]. Then there exists δ(t) > 0 such that f (x) − f (t) − f (t) < x−t
  • 11. 3.4 The Fundamental Theorem of Calculus 63 for all x ∈ [a, b] with |x − t| < δ(t). If u < t < v are two points straddling t and located in [a, b] ∩ (t − δ(t), t + δ(t)), then |f (v) − f (u) − f (t)(v − u)| ≤ |f (v) − f (t) − f (t)(v − t)| + |f (t) − f (u) − f (t)(t − u)| ≤ (v − t) + (t − u) (3.8) = (v − u). Inequality (3.8) also clearly holds when either u = t or v = t. Proposition 3.4.1 (Fundamental Theorem I) If f (x) is differentiable throughout [a, b], then b f (x) dx = f (b) − f (a). a Proof: Using the gauge δ(t) figuring in the straddle inequality (3.8), select a δ-fine partition π with mesh points a = s0 < s1 < · · · < sn = b and tags ti ∈ [si , si+1 ]. Application of the inequality and telescoping yield n−1 |f (b) − f (a) − S(f , π)| = [f (si+1 ) − f (si ) − f (ti )(si+1 − si )] i=0 n−1 ≤ |f (si+1 ) − f (si ) − f (ti )(si+1 − si )| i=0 n−1 ≤ (si+1 − si ) i=0 = (b − a). This demonstrates that f (x) has integral f (b) − f (a). The first half of the fundamental theorem remains valid for a continuous function f (x) that is differentiable except on a countable set N [250]. Since changing an integrand at a countable number of points does not alter its integral, it suffices to prove that b 0 t∈N f (b) − f (a) = g(t) dt, where g(t) = a f (t) t ∈ N. Suppose > 0 is given. For t ∈ N define the gauge value δ(t) to satisfy the straddle inequality. Enumerate the points tj of N, and define δ(tj ) > 0 so that |f (tj ) − f (tj + s)| < 2−j−2 whenever |s| < δ(tj ). Now select a δ-fine partition π with mesh points a = s0 < s1 < · · · < sn = b and tags ri ∈ [si , si+1 ]. Break the sum n−1 f (b) − f (a) − S(g, π) = f (si+1 ) − f (si ) − g(ri )(si+1 − si ) i=0
  • 12. 64 3. The Gauge Integral into two parts. Let S denote the sum of the terms with tags ri ∈ N, and let S denote the sum of the terms with tags ri ∈ N . As noted earlier, |S | ≤ (b − a). Because a tag is attached to at most two subintervals, the second sum satisfies |S | ≤ |f (si+1 ) − f (si )| ri ∈N ≤ |f (si+1 ) − f (ri )| + |f (ri ) − f (si )| ri ∈N ∞ ≤ 2 22−j−2 = . j=1 It follows that |S + S | ≤ (b − a + 1) and therefore that the stated integral exists and equals f (b) − f (a). In demonstrating the second half of the fundamental theorem, we will implicitly use the standard convention c d f (x) dx = − f (x) dx d c for c < d. This convention will also be in force in proving the substitution formula. Proposition 3.4.2 (Fundamental Theorem II) If a function f (x) is integrable on [a, b], then its indefinite integral t F (t) = f (x) dx a has derivative F (t) = f (t) at any point t where f (x) is continuous. The derivative is taken as one sided if t = a or t = b. Proof: In deriving the interval additivity rule (3.6), we showed that the integral F (t) exists. At a point t where f (x) is continuous, for any > 0 there is a δ > 0 such that − < f (x) − f (t) < when |x − t| < δ and x ∈ [a, b]. Hence, the difference F (t + s) − F (t) 1 t+s − f (t) = [f (x) − f (t)] dx s s t is less than and greater than − for |s| < δ. In the limit as s tends to 0, we recover F (t) = f (t). The fundamental theorem of calculus has several important corollaries. These are covered in the next three propositions on the substitution rule, integration by parts, and finite Taylor expansions.
  • 13. 3.4 The Fundamental Theorem of Calculus 65 Proposition 3.4.3 (Substitution Rule) Suppose f (x) is differentiable on [a, b], g(x) is differentiable on [c, d], and the image of [c, d] under g(x) is contained within [a, b]. Then g(d) d f (y) dy = f [g(x)]g (x) dx. g(c) c Proof: Part I of the fundamental theorem and the chain rule identity {f [g(x)]} = f [g(x)]g (x) imply that both integrals have value f [g(d)] − f [g(c)]. Proposition 3.4.4 (Integration by Parts) Suppose f (x) and g(x) are differentiable on [a, b]. Then f (x)g(x) is integrable on [a, b] if and only if f (x)g (x) is integrable on [a, b]. Furthermore, the two integrals are related by the identity b b f (x)g(x) dx + f (x)g (x) dx = f (b)g(b) − f (a)g(a), a a Proof: The product rule for derivatives is [f (x)g(x)] = f (x)g(x) + f (x)g (x). If two of three members of this identity are integrable, then the third is as well. Since part I of the fundamental theorem entails b [f (x)g(x)] dx = f (b)g(b) − f (a)g(a), a the proposition follows. The derivative of a function may itself be differentiable. Indeed, it makes sense to speak of the kth-order derivative of a function f (x) if f (x) is sufficiently smooth. Traditionally, the second-order derivative is denoted f (x) and an arbitrary kth-order derivative by f (k) (x). We can use these extra derivatives to good effect in approximating f (x) locally. The next proposition makes this clear and offers an explicit estimate of the error in a finite Taylor expansion of f (x). Proposition 3.4.5 (Taylor Expansion) Suppose f (x) has a derivative of order k + 1 on an open interval around the point y. Then for all x in the interval, we have k 1 (j) f (x) = f (y) + f (y)(x − y)j + Rk (x), (3.9) j=1 j!
  • 14. 66 3. The Gauge Integral where the remainder 1 (x − y)k+1 Rk (x) = f (k+1) [y + t(x − y)](1 − t)k dt. k! 0 If |f (k+1) (z)| ≤ b for all z between x and y, then b|x − y|k+1 |Rk (x)| ≤ . (3.10) (k + 1)! Proof: When k = 0, the Taylor expansion (3.9) reads 1 f (x) = f (y) + (x − y) f [y + t(x − y)]dt 0 and follows from the fundamental theorem of calculus and the chain rule. Induction and the integration-by-parts formula 1 f (k) [y + t(x − y)](1 − t)k−1 dt 0 1 1 = − f (k) [y + t(x − y)](1 − t)k k 0 x − y 1 (k+1) + f [y + t(x − y)](1 − t)k dt k 0 1 1 (k) x−y = f (y) + f (k+1) [y + t(x − y)](1 − t)k dt k k 0 now validate the general expansion (3.9). The error estimate follows directly from the bound |f (k+1) (z)| ≤ b and the integral 1 1 (1 − t)k dt = . 0 k+1 3.5 More Advanced Topics in Integration Within the confines of a single chapter, it is impossible to develop rigorously all of the properties of the gauge integral. In this section we will discuss briefly four topics: (a) integrals over unbounded intervals, (b) improper integrals and Hake’s theorem, (c) the interchange of limits and integrals, and (d) multidimensional integrals and Fubini’s theorem. Defining the integral of a function over an unbounded interval requires several minor adjustments. First, the real line is extended to include the points ±∞. Second, a gauge function δ(x) is now viewed as mapping x to an open interval containing x. The associated interval may be infinite; indeed, it must be infinite if x equals ±∞. In a δ-fine partition π, the
  • 15. 3.5 More Advanced Topics in Integration 67 interval Ij containing the tag xj is contained in δ(xj ). The length of an infinite interval Ij is defined to be 0 in an approximating Riemann sum S(f, π) to avoid infinite contributions to the sum. Likewise, the integrand f (x) is assigned the value 0 at x = ±∞. This extended definition carries with it all the properties we expect. Its most remarkable consequence is that it obliterates the distinction between proper and improper integrals. Hake’s theorem provides the link. If we allow a and b to be infinite as well as finite, then Hake’s theorem says a function f (x) is integrable over (a, b) if and only if either of the two limits b c lim f (x) dx or lim f (x) dx c→a c c→b a b exists. If either limit exists, then a f (x) dx equals that limit. For instance, the integral ∞ c c 1 1 1 dx = lim dx = lim − = 1 1 x2 c→∞ 1 x2 c→∞ x 1 exists and has the indicated limit by this reasoning. ∞ Example 3.5.1 Existence of 0 sinc(x) dx Consider the integral of sinc(x) = sin(x)/x over the interval (0, ∞). Because sinc(x) is continuous throughout [0, 1] with limit 1 as x approaches 0, the integral over [0, 1] is defined. Hake’s theorem and integration by parts show that the integral ∞ c sin x sin x dx = lim dx 1 x c→∞ 1 x c cos x c cos x = lim − − dx c→∞ x 1 1 x2 ∞ cos x = cos 1 − dx 1 x2 exists provided the integral of x−2 cos x exists over (1, ∞). We will demon- strate this fact in a moment. If we accept it, then it is clear that the integral of sinc(x) over (0, ∞) exists as well. As we shall find in Example 3.5.4, this integral equals π/2. In contrast to these positive results, sinc(x) is not absolutely integrable over (0, ∞). Finally, we note in passing that the sub- stitution rule gives ∞ ∞ ∞ sin cx sin y −1 sin y π dx = c dy = dy = . 0 x 0 c−1 y 0 y 2 for any c > 0.
  • 16. 68 3. The Gauge Integral We now ask under what circumstances the formula b b lim fn (x) dx = lim fn (x) dx (3.11) n→∞ a a n→∞ is valid. The two relevant theorems permitting the interchange of limits and integrals are the monotone convergence theorem and the dominated convergence theorem. In the monotone convergence theorem, we are given an increasing sequence fn (x) of integrable functions that converge to a finite limit for each x. Formula (3.11) is true in this setting provided b sup fn (x) dx < ∞. n a In the dominated convergence theorem, we assume the sequence fn (x) is trapped between two integrable functions g(x) and h(x) in the sense that g(x) ≤ fn (x) ≤ h(x) for all n and x. If limn→∞ fn (x) exists in this setting, then the inter- change (3.11) is allowed. The choices fn (x) = 1[1,n] (x)x−2 cos x , g(x) = −x−2 , h(x) = x−2 in the dominated convergence theorem validate the existence of ∞ n x−2 cos x dx = lim x−2 cos x dx. 1 n→∞ 1 We now consider two more substantive applications of the monotone and dominated convergence theorems. Example 3.5.2 Johann Bernoulli’s Integral As example of delicate maneuvers in integration, consider the integral 1 1 1 dx = e−x ln x dx 0 xx 0 1 ∞ (−x ln x)n = dx 0 n=0 n! ∞ 1 1 = (−x ln x)n dx . n=0 n! 0 The reader will notice the application of the monotone convergence theorem in passing from the second to the third line above. Further progress can be made by applying the integration by parts result 1 1 n lnn−1 x xm lnn x dx = − xm+1 dx 0 m+1 0 x 1 n = − xm lnn−1 x dx m+1 0
  • 17. 3.5 More Advanced Topics in Integration 69 recursively to evaluate 1 1 n! n! (−x ln x)n dx = xn dx = . 0 (n + 1)n 0 (n + 1)n+1 The pleasant surprise 1 ∞ 1 1 dx = 0 xx n=0 (n + 1)n+1 emerges. Example 3.5.3 Competing Definitions of the Gamma Function The dominated convergence theorem allows us to derive Gauss’s represen- tation n!nz Γ(z) = lim n→∞ z(z + 1) · · · (z + n) of the gamma function from Euler’s representation ∞ Γ(z) = xz−1 e−x dx . 0 As students of statistics are apt to know from their exposure to the beta distribution, repeated integration by parts and the fundamental theorem of calculus show that 1 n! xz−1 (1 − x)n dx = . 0 z(z + 1) · · · (z + n) The substitution rule yields 1 n n y nz xz−1 (1 − x)n dx = y z−1 1 − dy . 0 0 n Thus, it suffices to prove that ∞ n n y xz−1 e−x dx = lim y z−1 1 − dy . 0 n→∞ 0 n Given the limit y n lim 1− = e−y , n→∞ n we need an integrable function h(y) that dominates the nonnegative se- quence y n fn (y) = 1[0,n] (y)y z−1 1 − n
  • 18. 70 3. The Gauge Integral from above in order to apply the dominated convergence theorem. In light of the inequality y n 1− ≤ e−y , n the function h(y) = y z−1 e−y will serve. Finally, the gauge integral extends to multiple dimensions, where a ver- sion of Fubini’s theorem holds for evaluating multidimensional integrals via iterated integrals [278]. Consider a function f (x, y) defined over the Cartesian product H × K of two multidimensional intervals H and K. The intervals in question can be bounded or unbounded. If f (x, y) is integrable over H × K, then Fubini’s theorem asserts that the integrals H f (x, y) dx and K f (x, y) dy exist and can be integrated over the remaining variable to give the full integral. In symbols, f (x, y) dx dy = f (x, y) dx dy = f (x, y) dy dx . H×K K J J K Conversely, if either iterated integral exists, one would like to conclude that the full integral exists as well. This is true whenever f (x, y) is nonnega- tive. Unfortunately, it is false in general, and two additional hypotheses introduced by Tonelli are needed to rescue the situation. One hypothesis is that f (x, y) is measurable. Measurability is a technical condition that holds except for very pathological functions. The other hypothesis is that |f (x, y)| ≤ g(x, y) for some nonnegative function g(x, y) for which the iterated integral exists. This domination condition is shared with the dom- inated convergence theorem and forces f (x, y) to be absolutely integrable. ∞ Example 3.5.4 Evaluation of 0 sinc(x) dx According to Fubini’s theorem n nπ nπ n e−xy sin x dx dy = e−xy sin x dy dx . (3.12) 0 0 0 0 The second of these iterated integrals nπ n nπ nπ sin x sin x e−xy sin x dy dx = dx − e−nx dx 0 0 0 x 0 x ∞ tends to 0 as n tends to ∞ by a combination of Hake’s theorem sinc(x) dx and the dominated convergence theorem. The inner integral of the left iterated integral in (3.12) equals nπ nπ nπ e−xy sin x dx = −e−xy cos x − ye−xy sin x 0 0 0 nπ − y2 e−xy sin x dx 0 nπ −nπy = 1−e cos nπ − y 2 e−xy sin x dx 0
  • 19. 3.6 Problems 71 after two integrations by parts. It follows that nπ 1 − e−nπy cos nπ e−xy sin x dx = . 0 1 + y2 Finally, application of the dominated convergence theorem gives ∞ n 1 − e−nπy cos nπ 1 lim dy = dy n→∞ 0 1 + y2 0 1 + y2 π = . 2 Equating the limits of the right and left hand sides of the identity (3.12) ∞ therefore [278] yields the value of π/2 for 0 sinc(x) dx. 3.6 Problems 1. Give an alternative proof of Cousin’s lemma by letting y be the supre- mum of the set of x ∈ [a, b] such that [a, x] possesses a δ-fine partition. 2. Use Cousin’s lemma to prove that a continuous function f (x) defined on an interval [a, b] is uniformly continuous there [108]. (Hint: Given > 0 define a gauge δ(x) by the requirement that |f (y) − f (x)| < 1 2 for all y ∈ [a, b] with |y − x| < 2δ(x).) 3. A possibly discontinuous function f (x) has one-sided limits at each point x ∈ [a, b]. Show by Cousin’s lemma that f (x) is bounded on [a, b]. 4. Suppose f (x) has a nonnegative derivative f (x) throughout [a, b]. Prove that f (x) is nondecreasing on [a, b]. Also prove that f (x) is constant on [a, b] if and only if f (x) = 0 for all x. (Hint: These yield easily to the fundamental theorem of calculus. Alternatively for the first assertion, consider the function f (x) = f (x) + x for > 0.) 5. Using only the definition of the gauge integral, demonstrate that b −a f (t) dt = f (−t) dt a −b when either integral exists.
  • 20. 72 3. The Gauge Integral 6. Based on the standard definition of the natural logarithm y 1 ln y = dx, 1 x prove that ln yz = ln y + ln z for all positive arguments y and z. Use this property to verify that ln y −1 = − ln y and that ln y r = r ln y for every rational number r. 7. Apply Proposition 3.3.2 and demonstrate that every monotonic func- tion defined on an interval [a, b] is integrable on that interval. 8. Let f (x) be a continuous real-valued function on [a, b]. Show that there exists c ∈ [a, b] with b f (x) dx = f (c)(b − a). a 9. In the Taylor expansion of Proposition 3.4.5, suppose f (k+1) (x) is continuous. Show that we can replace the remainder by (x − y)k+1 (k+1) Rk (x) = f (z) (k + 1)! for some z between x and y. 10. Suppose that f (x) is infinitely differentiable and that c and r are pos- itive numbers. If |f (k) (x)| ≤ ck!rk for all x near y and all nonnegative integers k, then use Proposition 3.4.5 to show that ∞ f (k) (y) f (x) = (x − y)k k! k=0 near y. Explicitly determine the infinite Taylor series expansion of the function f (x) = (1 + x)−1 around x = 0 and justify its convergence. 11. Suppose the nonnegative continuous function f (x) satisfies b f (x) dx = 0. a Prove that f (x) is identically 0 on [a, b]. 12. Consider the function f (x) = x2 sin (x−2 ) x = 0 0 x = 0. 1 1 Show that 0 f (x) dx = sin (1) and limt↓0 t |f (x)| dx = ∞. Hence, f (x) is integrable but not absolutely integrable on [0,1].
  • 21. 3.6 Problems 73 13. Prove that ∞ β 1 α+1 xα e−x dx = Γ 0 β β for α and β positive [82]. 14. Justify the formula 1 ∞ ln(1 − x) 1 = − . 0 x n=1 n2 15. Show that ∞ xz−1 dx = ζ(z)Γ(z), 0 ex − 1 ∞ where ζ(z) = n=1 n−z . 16. Prove that the functions ∞ sin t f (x) = dt 1 x2 + t2 ∞ g(x) = e−xt cos t dt, x > 0, 0 are continuous. 17. Let fn (x) be a sequence of integrable functions on [a, b] that converges uniformly to f (x). Demonstrate that f (x) is integrable and satisfies b b lim fn (x) dx = f (x) dx . n→∞ a a (Hints: For > 0 small take n large enough so that fn (x) − ≤ f (x) ≤ fn (x) + 2(b − a) 2(b − a) for all x.) 18. Let p and q be positive integers. Justify the series expansion 1 ∞ xp−1 (−1)n dx = 0 1 + xq n=0 p + nq by the monotone convergence theorem. Be careful since the series does not converge absolutely [278].
  • 22. 74 3. The Gauge Integral 19. Suppose f (x) is a continuous function on R. Demonstrate that the sequence n−1 1 k fn (x) = f x+ n n k=0 converges uniformly to a continuous function on every finite interval [a, b] [69]. 20. Prove that 1 xb − xa b+1 dx = ln 0 ln x a+1 for 0 < a < b [278] by showing that both sides equal the double integral xy dx dy . [0,1]×[a,b] 21. Integrate the function y 2 − x2 f (x, y) = (x2 + y 2 )2 over the unit square [0, 1]× [0, 1]. Show that the two iterated integrals disagree, and explain why Fubini’s theorem fails. 2 2 22. Suppose the two partial derivatives ∂x∂∂x2 f (x) and ∂x∂∂x1 f (x) exist 1 2 and are continuous in a neighborhood of a point y ∈ R2 . Show that they are equal at the point. (Hints: If they are not equal, take a small box around the point where their difference has constant sign. Now apply Fubini’s theorem.) 23. Demonstrate that ∞ √ 2 π e−x dx = 0 2 2 2 by evaluating the integral of f (y) = y2 e−(1+y1 )y2 over the rectangle (0, ∞) × (0, ∞).