Bounded Derivatives Which Are Not Riemann Integrable
Bounded Derivatives Which Are Not Riemann Integrable
Bounded Derivatives Which Are Not Riemann Integrable
by
Elliot M. Granath
Whitman College
2017
Certificate of Approval
Russell A. Gordon
Whitman College
May 10, 2017
Contents
Acknowledgements iv
Abstract v
iii
Acknowledgements
iv
Abstract
v
1 Finding an antiderivative of a continuous
1
F (x)
f (x)
2
Consider the function
g(x),
if a ≤ x < c;
f (x) =
d1 x + d2 , if c ≤ x ≤ b,
1
F (c + h) − F (c) d1 (c + h)2 + d2 (c + h) − 21 d1 c2 − d2 c
lim+ = lim+ 2
h→0 h h→0 h
1 2
d1 ch + 2 d1 h + d2 h
= lim+
h→0 h
= d1 c + d2 ,
= d1 c + d2 .
3
Having established that continuous piecewise linear functions have an-
tiderivatives, we claim that any continuous function on a closed interval can
be estimated with a sequence of such functions. Figure 2 shows a simple con-
tinuous function, together with a linear approximation in six parts. As our
intuition suggests, we can choose increasingly fine linear approximations that
will converge to the desired function. Using the Mean Value Theorem, we
can show that the corresponding sequence of antiderivatives converges to an
antiderivative of f .
f (x)
Theorem 1.4. If f : [a, b] → R is continuous on [a, b], then there exists a se-
quence {fn } of continuous piecewise linear functions that converges uniformly
to f on [a, b].
4
We claim that {fn } converges uniformly to f on [a, b]. Let > 0 and pick
N such that 1/N < /2. Let x ∈ [a, b], and identify an interval [xi , xi+1 ]
containing x. Then |xi − x| < δn , and without loss of generality, for all
n ≥ N , δn < δN . Also note that, for all 1 ≤ i ≤ n, if x, y ∈ [xi−1 , xi ], then
|f (y)−f (x)| ≤ max{x : x ∈ [xi−1 , xi ]}−min{x ∈ [xi−1 , xi ]} = |f (xi )−f (xi−1 )|.
Using this fact together with the uniform continuity of f ,
< .
In general, it is not necessarily true that lim lim f (x, y) = lim lim f (x, y).
x→x0 y→y0 y→y0 x→x0
For example,
x + 2y x + 2y
1 = lim lim 6= lim lim = 2.
x→0 y→0 x + y y→0 x→0 x + y
Before continuing, we will establish Lemma 1.5, which shows that, in the
right circumstances, we may simply interchange limits. In the same way that
uniform continuity preserves continuity, it also preserves limits.
5
Proof. Denote yn = lim fn (x). Since {fn } converges uniformly, given > 0,
x→c
there exists N such that if n, m > N , then |fn (x) − fm (x)| < for all points
x ∈ [a, b] \ {c} (this is the Cauchy Criterion for Uniform Convergence). Then,
since |fn (x) − fm (x)| < , we have |yn − ym | = | lim(fn (x) − fm (x))| ≤ . Thus,
x→c
< 3.
Example 1.6. This example is adapted from Gordon [4]. For each n ∈ Z+ ,
define Fn : [0, 1] → R where
if x ≥ n1 ,
x,
Fn (x) =
3 nx2 − 1 n3 x4 , if x < n1 .
2 2
6
Furthermore, the right-sided derivative is 1 and the left-sided derivative is
3
3n n1 − 2n3 n1
= 1, so Fn is differentiable on [0, 1]. For all n > 1, if
x ∈ [0, 1/n],
3 2 1 3 4
|Fn (x) − x| ≤ nx + n x + |x|
2 2
3 1 2
≤ + +
2n 2n 2n
3
= .
n
Proof. (Rudin [5]) Let > 0. Choose N such that if n, m ≥ N , then we have
|Fn (c) − Fm (c)| < and |Fn0 (x) − Fm0 (x)| < for all x ∈ [a, b]. Let x, y ∈ [a, b]
7
where x < y. By the Mean Value Theorem, for some d ∈ (x, y),
(Fn − Fm )(x) − (Fn − Fm )(y)
|(Fn − Fm )(x) − (Fn − Fm )(y)| = |x − y|
x−y
(1)
= |x − y||(Fn − Fm )0 (d)|
< |x − y|
≤ (b − a).
|Fn (x) − Fm (x)| ≤ |(Fn − Fm )(x) − (Fn − Fm )(c)| + |Fn (c) − Fm (c)|
< (b − a) +
= (b − a + 1).
Thus {Fn } converges uniformly to some F on [a, b]. Fix c ∈ [a, b] and define
the functions
(Fn − Fm )(x) − (Fn − Fm )(c)
|φn (x) − φm (x)| =
<
x−c
8
uniformly to φ on [a, b] \ {c}. By Lemma 1.5,
= lim fn (c)
n→∞
= f (c).
Proof. Let f be a continuous function on [a, b]. By Theorem 1.4, there exists
a sequence {fn } of continuous piecewise linear functions that converges uni-
formly to f on [a, b]. By Theorem 1.3, there exists a sequence {Fn } of functions
on [a, b] such that Fn0 (x) = fn (x) for all x ∈ [a, b]. For convenience, choose
each Fn such that Fn (a) = 0. Then by Theorem 1.7, {Fn } converges uniformly
to a function F on [a, b], F is differentiable on [a, b], and F 0 (x) = f (x) for all
x ∈ [a, b].
9
1 ≤ i ≤ n; additionally, ti ∈ [xi−1 , xi ] for 1 ≤ i ≤ n. Denote the norm of a
tagged partition tP by ||tP || = max{xi − xi−1 : 1 ≤ i ≤ n}. Finally, define the
Riemann sum of f with tagged partition tP by
n
X
S(f,tP ) = f (ti )(xi − xi−1 ).
i=1
Proof. By Theorem 1.8, f has an antiderivative F on [a, b]. Let > 0. Since
f is uniformly continuous, there is δ > 0 such that, if |y − x| < δ, then
|f (y) − f (x)| < . Let the set tP = {([xi−1 , xi ], ti ) : 1 ≤ i ≤ n} be a partition
of [a, b] where ||tP || < δ. By the Mean Value Theorem, for each i ∈ {1, . . . , n},
there is si ∈ (xi−1 , xi ) satisfying F (xi ) − F (xi−1 ) = f (si )(xi − xi−1 ). Then, for
each i ∈ {1, . . . , n}, we have
|f (ti )(xi − xi−1 ) − (F (xi ) − F (xi−1 ))| = |f (ti )(xi − xi−1 ) − f (si )(xi − xi−1 )|
10
It follows that
Xn n
X
|S(f,tP ) − (F (b) − F (a))| = f (ti )(xi − xi−1 ) − (F (xi ) − F (xi−1 ))
i=1 i=1
n
X
≤ |f (ti )(xi − xi−1 ) − (F (xi ) − F (xi−1 ))|
i=1
n
X
< (xi − xi−1 )
i=1
= (b − a).
Rb
Thus, f is Riemann integrable on [a, b] and a
f = F (b) − F (a).
In the proof of Theorem 1.7, we used the Mean Value Theorem to show that, if
a sequence of functions {fn } uniformly converges to f , then an appropriately
chosen corresponding sequence of antiderivatives {Fn } uniformly converges to
F where F 0 = f . However, we can provide a similar version of Rudin’s proof
without using the Mean Value Theorem. We use instead the following result.
11
F is an antiderivative for f on [a, b], then
1
|F (b) − F (a)| = (b − a) m(a + b) + n
2
1
≤ (b − a) · · (|ma + n| + |mb + n|)
2
= (b − a)|mb + n|
where G is an antiderivative of g.
12
Then
integrable
13
integral.” In this case, we might write
Z 1 Z 1
0
f = lim+ f0 = 1
0 a→0 a
This example uses an open interval, and takes advantage of the fact that
f 0 is unbounded at an endpoint. Suppose we restrict ourselves to functions
which are continuous and differentiable on a closed interval [a, b] (with one-
sided derivatives at a and b). It turns out that the derivative may still be
unbounded. Consider the function g : [−1, 1] → R where
xk sin 1 ,
x
if x 6= 0,
g(x) =
0,
if x = 0,
for 1 < k < 2. Then, away from the origin, we can differentiate g using
elementary methods. At the origin, we have
hk sin h1
k−1 1
lim = lim h sin
h→0 h h→0 h
= 0.
Thus
kxk−1 sin 1 − xk−2 cos 1 ,
x x
if x 6= 0,
g 0 (x) =
0,
if x = 0.
On the domain [−1, 1], g 0 is defined. However, while the term kxk−1 sin x1 is
bounded, xk−2 cos x1 is not. Thus the derivative of g is not Riemann integrable,
despite the fact that g is differentiable everywhere.
In both of the examples we just explored, we find a way to ensure that the
14
derivative of a function is unbounded, which shows that the derivative is not
Riemann integrable without further thought. What if we add an additional
requirement that f 0 be bounded? Is it always integrable? The answer is
still no. The most widely known counter-example was constructed in 1881
by Vito Volterra [6]. We will provide an overview of his function, before
focusing rigorously on another example by Dimitrie Pompeiu with certain
other properties.
15
Then we have
2x sin 1 − cos 1 ,
x x
if x 6= 0,
f 0 (x) =
0,
if x = 0.
Note that f 0 is bounded on a finite interval containing the origin, but f 0 is not
continuous at 0. Now recall the Cantor set, which is constructed by taking the
interval [0, 1] and removing the middle third, then removing the middle third
of the remaining intervals, and so on. The resulting set has measure zero, so
we will use a modified version of the Cantor set that begins with removing the
middle fourth of the interval [0, 1], etc. (the details are omitted here). Note
that this so-called thick Cantor set does not have measure zero (the sum total
of the lengths of intervals removed is 1/2).
16
and define the reflection of a across x = 1/8:
x20 sin x10 , if 1/8 ≤ x ≤ 1
+ x0 ;
8
1
2
1 1
b(x) = 4
− x sin 1 , if 8
+ x0 < x < 14 ;
4
−x
0, if x = 41 .
Finally, define the first term in our sequence f1 . Remember that a and b are
functions, not constants.
if 0 ≤ x ≤ 83 ;
0,
a x − 3 , 3
< x ≤ 12 ;
8
if 8
f1 (x) =
3 1
< x ≤ 58 ;
b x − 8
, if 2
5
0,
if 8
≤ x ≤ 1.
17
f1 (x)
In Volterra’s function, the guiding idea was to find a function whose set
of discontinuities does not have have measure zero. Consider the fact that,
if a derivative f 0 is Riemann integrable on [a, b], then its integral must be
f (b) − f (a). On the other hand, given a sequence of partitions of [a, b] whose
norm converges to 0, we must also have that respective sequence of upper
and lower Riemann sums of f 0 converge to f (b) − f (a). Pompeiu’s example
hinges on the construction of a positive, strictly increasing function h whose
derivative vanishes on a dense subset. In this manner, if h0 were integrable,
then its integral must be the positive number h(b) − h(a). However, given any
partition of its domain, the lower Riemann sum of h0 is zero, since h0 (x) = 0
on a dense subset, and thus the integral must be 0. We conclude that h0 is not
18
Riemann integrable.
This function was originally developed by Pompeiu, and the proof we use
here was outlined by Bruckner et al. [1]. Throughout Section 2.2, the meaning
of the functions f and h, will not change, nor will the sequences {ak }, {qk },
and {fk }.
p(h) h1/3
lim = lim =∞
h→0 h h→0 h
and, away from the origin, p0 (x) is simply the positive number 13 x−2/3 . Now
let {qk } be a listing of Q ∩ [0, 1], and for each k ∈ Z+ , define
fk (x) = (x − qk )1/3 .
∞
P 3/5
Let {ak } be a sequence of positive numbers such that ak converges. The
k=1
reason for the exponent 3/5 becomes evident in the proof of Theorem 2.2. For
∞
P
now, we simply care that the sum ak converges. This will be sufficient to
k=1
19
∞
P
prove that the sum ak fk (x) converges uniformly on [0, 1].
k=1
∞
P ∞
P
Lemma 2.1. If ak converges, then ak fk (x) converges uniformly on
k=1 k=1
[0, 1].
1 1 1 1 1
6 5 4 3 2
The intention with the construction of f thus far was to ensure that the deriva-
tive of f is ∞ on a dense subset. Knowing that f 0 “blows up” on a dense subset,
20
the following result is surprising.
∞
P 3/5
Theorem 2.2. Let {ak } be a sequence of positive numbers such that ak
k=1
∞
ak fk0 (x)
P
converges. Then converges almost everywhere.
k=1
∞
ak fk0 (x) diverges. Let
P
Proof. Let D be the set of points x ∈ [a, b] where
k=1
> 0 and define the interval
3/5 3/5
Ak = q k − ak , qk + ak .
∞
S
Let A = Ak and note that Q ∩ [0, 1] ⊆ A. Denoting the length of Ak by
k=1
3/5
`(Ak ), we have `(Ak ) = 2ak and
∞ ∞
3/5
X X
`(Ak ) = 2 ak = 2L,
k=1 k=1
3/5
where L is a real number. Let c ∈ [0, 1] \ A. Note that |c − qk | ≥ ak , so
3/5
ak ak a
2/3
≤ 3/5
= k2/3 .
3|c − qk | 3(ak )2/3 3
∞
P 1 3/5
Since a
32/3 k
converges and ak is positive for k ≥ 1, by the comparison
k=1
test,
∞ ∞
X ak X
= ak fk0 (c).
k=1
3(c − qk )2/3 k=1
∞
ak fk0 (c) converges;
P
converges. We have shown that, for any c ∈ [a, b] \ A,
k=1
thus, D ⊆ A. Since > 0 was arbitrary and the sum of the lengths of the
∞
ak fk0 (x) converges almost
P
intervals in A is 2L, D has measure zero and
k=1
everywhere on [0, 1].
Remark. It follows from Theorem 3.5 that we can use many different se-
quences for {ak }. We can use any geometric sequence {rk }, where 0 < r < 1,
21
since (rk )3/5 = (r3/5 )k and 0 < r3/5 < 1. When outlining the basics of Pom-
peiu’s function, then, we can simply pick ak = 1/2k . Additionally, we can use
the sequence {1/np }∞
n=1 , for any p > 5/3.
Definition 2.3. Let f : [a, b] → R be a function and let c ∈ (a, b). We say
that f 0 (c) = ∞ if, for all M > 0, there exists δ > 0 such that
f (x) − f (c)
≥M
x−c
22
For claim (1), note that, for some j ∈ Z+ , c = qj . Since fj0 (c) = ∞, for any
M > 0, there is δ > 0 such that 0 < |x − c| < δ implies
fj (x) − fj (c)
≥ M.
x−c
∞
f (x) − f (c) X fk (x) − fk (c)
=
x−c k=1
x−c
fj (c + h) − fj (c)
≥
h
≥ M.
This shows that f 0 (c) = ∞. Since each fk is increasing, fk0 (x) is positive for
all k ∈ Z+ and x ∈ [0, 1]. Additionally, recall that each ak is also positive. It
∞
ak fk0 (c) = ∞.
P
follows that
k=1
∞
ak fk0 (c) converges to ∞. Let
P
For claim (2), suppose that c ∈
/ Q and
k=1
N
ak fk0 (c) > M . Then
P
M > 0. There exists a positive integer N such that
k=1
there exists δ > 0 such that if 0 < |x − c| < δ, then
N
X fk (x) − fk (c)
ak > M.
k=1
x−c
∞
f (x) − f (c) X fk (x) − fk (c)
= ak
x−c k=1
x−c
N
X fk (x) − fk (c)
≥ > M.
k=1
x−c
23
∞
ak fk0 (c) converges to a real number rk . Note
P
For claim (3), suppose that
k=1
that c must not be a rational number. Since c is neither 0 nor 1, we may talk
about the two-sided derivative of f at c. Thus, there is some η > 0 such that
[c − η, c + η] ⊆ [0, 1]. Define the function φ : [−η, η] → R by
∞
f (c + x) − f (c) X (c + x − qk )1/3 − (c − qk )1/3
φ(x) = = ak . (2)
x k=1
x
Note that, if f 0 (c) exists, then lim φ(x) = f 0 (c). We will show that the sum
x→0
in (2) converges uniformly on its domain and apply Lemma 1.5. Note that
x = (c + x − qk ) − (c − qk )
and
a1/3 − b1/3 1
= 2/3
a−b a + a1/3 b1/3 + b2/3
∞
X (c + x − qk )1/3 − (c − qk )1/3
φ(x) = · ak
k=1
(c + x − qk ) − (c − qk )
∞
X ak
= . (3)
k=1
(c + x − qk )2/3 + (c + x − qk )1/3 (c − qk )1/3 + (c − qk )2/3
∞
X ak 1
2/3
· 2 .
k=1
(c − qk ) rk + rk + 1
24
Now, since rk is a real number, by considering rk2 + rk + 1 as a polynomial in
rk , we see that rk2 + rk + 1 has a global minimum of 3/4. Thus, for all k ∈ Z+ ,
a k 1 4 a k
(c − qk )2/3 · r2 + rk + 1 ≤ 3 (c − qk )2/3 .
k
∞ ∞
X 1X ak
ak fk0 (c) =
k=1
3 k=1 (c − qk )2/3
converges. Since the terms are positive, the sum converges absolutely, and
P∞ 4 ak
< ∞. By the Weierstrass M -test,
2/3
3 (c − qk )
k=1
∞
X (c + x − qk )1/3 − (c − qk )1/3
ak
k=1
x
∞
as desired. We have shown that f 0 (x) = ak fk0 (x) for all x ∈ [0, 1].
P
k=1
25
Recall that our goal is to create a strictly increasing function whose deriva-
tive vanishes at some point in every interval. So far we have created a function
whose derivative is infinite at some point on every interval. We turn now to
the inverse of f .
Proof. Let {wn } be any sequence in [a, b] \ {d} converging to d and let {vn } be
the corresponding sequence in [0, 1] converging to c where f (vn ) = wn . Then
h(wn ) − h(d) vn − c
lim = lim . (4)
n→∞ wn − d n→∞ f (vn ) − f (c)
Since the limit of the reciprocal in (4) converges to infinity (since f 0 (c) = ∞),
we have that h0 (d) = 0.
26
Proof. Let c ∈ (a, b) and let V be an open interval in (a, b) containing c. Since
f is continuous, the preimage of V is open. Call it U . Since the rationals are
dense in [0, 1], U contains a rational number q. Then f (q) is in S and V , so S
is dense in [a, b].
ak
ak fk0 (x) = .
3(x − qk )2/3
ak
ak fk0 (x) ≥ .
3
Since each ak is positive, we may use, for example, a1 /3 as a lower bound for
f 0 . This shows that h0 is bounded.
27
Theorem 2.8. The function h0 is not Riemann integrable on any subinterval
of its domain.
Proof. Suppose that h0 is Riemann integrable on [c, d] ⊆ [a, b]. By the Funda-
Rd
mental Theorem of Calculus, c h0 = h(d) − h(c) > 0 (recall that h is strictly
increasing). On the other hand, h0 vanishes on a dense set of [c, d]. Thus, for
any partition of [c, d], the lower Riemann sum of h0 is 0. Thus, we must have
Rd 0
c
h = 0, a contradiction. Therefore h0 is not integrable on [c, d].
As noted in Theorems 2.5 and 2.6, h0 (x) = 0 on a dense subset of [a, b].
However, the fact that h0 is not integrable on any subinterval leads to the
following observation.
Theorem 2.9. The function h0 attains the value 0 on a dense subset, attains
a positive value on another dense subset, and, where h0 is positive, it is also
discontinuous.
28
Proof. We have seen that h0 (x) = 0 on a dense subset, and we know from the
proof of Theorem 2.7 that h0 is nonnegative. Suppose that h0 (x) is identically
0 on an interval. Then h0 is integrable over that interval, which contradicts
Theorem 2.8. Thus h0 attains a positive value on every interval, so h0 (x) is
positive on a dense subset of [a, b].
29
Now define
f˜(x + 1/n) − f˜(x)
φn (x) = .
1/n
Then each φn is continuous on [a, b] since both f and f˜ are continuous on [a, b].
Since lim φn (x) = f 0 (x) for all x ∈ [a, b], f 0 is a Baire class one function.
n→∞
Theorem 2.11. If f : [a, b] → R is a Baire class one function, then the preim-
age of any open set under f is an Fσ set.
∞ [
[ ∞ \
∞
−1
f ((−∞, r)) = fp−1 ((−∞, r − 1/k]). (5)
k=1 n=1 p=n
30
Since each fp is continuous, we know fp−1 ((−∞, r − 1/k]) is closed for every
p, k ∈ Z+ . Their intersection is closed, and the countable union of these closed
sets is an Fσ set. That is, if equation (5) is true, then f −1 ((−∞, r)) is an Fσ set.
Suppose that x ∈ f −1 ((−∞, r)). Then f (x) ∈ (−∞, r), and, for some
K ∈ Z+ , we have f (x) ∈ (−∞, r − 1/K). Since {fp (x)} converges to f (x),
there is some N such that, if p ≥ N , we have fp (x) ∈ (−∞, r − 1/K). That
is,
∞
\ ∞
\
x∈ fp−1 ((−∞, r − 1/K)) ⊆ fp−1 ((−∞, r − 1/K])
p=N p=N
Conversely, begin by supposing (6) is true. Then there exist positive integers
K and N such that
∞
\
x∈ fp−1 ((−∞, r − 1/K)).
p=N
This means that fp (x) ∈ (−∞, r − 1/K) for all p ≥ N . Since {fp (x)}∞
p=N con-
verges to f (x), we have that f (x) ≤ r − 1/K < r, and thus x ∈ f −1 ((−∞, r)).
This proves equation (5). As we noted, this establishes that f −1 (−∞, r)), and
similarly that f −1 ((r, ∞)), are Fσ sets.
Now, given any open set U , we can write U as the countable union of open
intervals (ai , bi ) where i ≥ 1, and thus
∞
[
U= ((−∞, bi ) ∩ (ai , ∞)) . (7)
i=1
31
Using some basic properties of inverse set functions, note the following:
∞ ∞
!
[ [
f −1 f −1 ((−∞, bi )) ∩ f −1 ((ai , ∞)) .
(−∞, bi ) ∩ (ai , ∞) = (8)
i=1 i=1
Now, since the intersection of two Fσ sets is an Fσ set, and the countable union
of Fσ sets is an Fσ set, it follows from (7) and (8) that f −1 (U ) is an Fσ set.
Proof. By Theorem 2.10, we know that h0 is a Baire class one function. Us-
ing Theorem 2.11, we see that h0−1 ((−∞, 0) ∪ (0, ∞)) is an Fσ set, and thus
h0−1 ({0}) is a Gδ set. As we noted previously (see Theorems 2.5 and 2.6),
h0 (S) = {0}, and so S ⊆ h0−1 ({0}). If we can show that S is not a Gδ set,
then S is a proper subset of h0−1 ({0}), which guarantees some x ∈
/ S which
maps to 0 under h0 . By way of contradiction, suppose S is a Gδ set and thus
32
that [a, b] \ S is an Fσ set. There is a sequence {En } of closed sets such that
∞
[
[a, b] \ S = En .
n=1
Now, since S is dense in [a, b], none of the sets En may contain an open interval.
Thus, {En } is a sequence of nowhere dense sets. But then
∞ ∞
! !
[ [
[a, b] = ([a, b] \ S) ∪ S = En ∪ {f (qn )} .
n=1 n=1
However, the Baire Category Theorem guarantees a point x ∈ [a, b] not con-
tained in any of the nowhere dense sets En or {f (qk )}. Thus, S is not a Gδ
/ S such that h0 (x) = 0. This corresponds to a
set. There must be a point x ∈
/ [0, 1] ∩ Q such that f 0 (t) = ∞.
point t ∈
Theorem 3.1. If f 0 is bounded, then f maps any set of measure zero to a set
of measure zero.
∞ ∞
[ X
S⊆ (ak , bk ) and (bk − ak ) < .
k=1 k=1
M +1
33
Since each x ∈ S is contained in some interval (aj , bj ), which is in turn con-
tained in f ((aj , bj )),
∞
[
f (S) ⊆ f ((ak , bk )).
k=1
Using the Mean Value Theorem, if we take any two points x, y ∈ (ak , bk ), we
have
|f (y) − f (x)| ≤ |y − x|M ≤ (bk − ak )M
and thus, for each k ∈ Z+ , we can cover f ((ak , bk )) with an open interval of
length (M + 1)(bk − ak ). Since
∞
X
(M + 1)(bk − ak ) < ,
k=1
Theorem 3.2. The set E = {x : h0 (x) 6= 0} does not have measure zero.
The proof of Theorem 3.2 relies on the fact that functions with bounded
derivatives map sets of measure to sets of measure zero. We can prove a more
general result using the notion of absolute continuity.
34
Definition 3.3. A function f on an interval I is absolutely continuous if, for
each > 0, there exists δ > 0 satisfying the following condition: if the set
{[ci , di ] : 1 ≤ i ≤ n} is a finite collection of disjoint intervals in I that satisfies
n
P Pn
the inequality (di − ci ) < δ, then |f (di ) − f (ci )| < .
i=1 i=1
It is easy to see from Definition 3.3 that absolute continuity implies uniform
continuity (simply take n = 1 as it is used in the definition). What is less
obvious is that there are uniformly continuous functions that are not absolutely
continuous. The Cantor function, which we won’t explore here, is a canonical
example, but we will now give an example of a more elementary function which
is uniformly continuous, but not absolutely continuous.
π
lim+ x sin = 0,
x→0 x
and so f is continuous on [0, 1]. This also shows that f is uniformly continuous
on its domain. We note that when
π π
= + 2πn,
x 2
for n ∈ Z, then
2
f (x) = x = .
4n + 1
35
π π 2
Similarly, if = − + 2πn, then f (y) = −y = − . For the same values
y 2 4n − 1
x and y, we have
∞
1
P
Now let = 1 and let δ > 0. Since n
is the divergent harmonic series, each
n=1
∞ ∞
1 2 1
P P
tail n
is also divergent. Take N such that 4n−1
< δ. Since n
diverges,
n=N n=N
M
1
P
there is some positive integer M such that n
> 1 = . Consider the set of
n=N
disjoint intervals {[xn , yn ] : N ≤ n ≤ M } where
2 2
xn = , and yn = .
4n + 1 4n − 1
M M M
1
P P P
Then (yi − xi ) < δ and |f (yi ) − f (xi )| > n
> 1 = , so f is not
n=N n=N n=N
absolutely continuous.
Theorem 3.1 stated that a function f which has a bounded derivative maps
any set of measure zero to a set of measure zero. Here, we will prove a similar
result for absolutely continuous functions. Finally, we will give an alternate
proof of Theorem 3.1 by showing that if f has a bounded derivative on [a, b],
then f is absolutely continuous on [a, b].
Proof. Let > 0. Since f is absolutely continuous, there is some δ > 0 such
∞
P
that, for any sequence {[rk , sk ]} of disjoint intervals satisfying (sk − rk ) < δ,
k=1
∞
P
we have |f (sk ) − f (rk )| < . Since A has measure zero, there exists a
k=1
36
∞
S ∞
P
sequence of open sets {(ak , bk )} such that A ⊆ (ak , bk ) and (bk − ak ) < δ.
k=1 k=1
Since any open set is the union of disjoint open intervals, we may assume that
the (ak , bk )’s are disjoint. By the absolute continuity of f ,
∞
X
|f (bk ) − f (ak )| ≤ .
k=1
∞
[ ∞
[
B⊆ f ([ak , bk ]) = [ck , dk ],
k=1 k=1
∞
(dk − ck + /2k−1 ) ≤ + 2, it follows that
P
Because > 0 was arbitrary and
k=1
B has measure zero.
37
Proof. Let > 0 and let M > 0 be a bound for f 0 . Suppose that the set
{[ci , di ] : 1 ≤ i ≤ n} is a finite collection of disjoint intervals in [a, b] that
Pn
satisfies the inequality (di − ci ) < . By the Mean Value Theorem, for each
i=1
i ∈ {1, . . . , n}, there is some ti ∈ (ci , di ) such that |f (di ) − f (ci )| < M (di − ci ).
Then
n
X n
X
|f (di ) − f (ci )| < M (di − ci ) < M,
i=1 i=1
38
then the following analogy, which we will prove, is fitting:
• The function f : [a, b] → R is a Baire class one function if and only if the
Before we turn our attention to this main idea, we first prove another fact
which alludes back to a common fact about uniform continuity. In an analysis
course, one might learn that uniform convergence “preserves” continuity. That
is, if {fn } converges uniformly to f , and each fn is continuous, then f is
continuous. Again, in a pleasing analogy, it turns out that, if each fn is a
Baire class one function, then f is a Baire class one function. To prove this,
we will first consider the sequence {fn } in a series representation.
Now let c ∈ [a, b]. If |fk (c)| < M , then on some open interval containing c,
hk (x) is equal to fk (x), and so hk is continuous at c. Similarly, if |fk (c)| > M ,
then hk is constant on some neighborhood of c. If fk (c) = M , then for any
> 0, there is δ > 0 such that |x − c| < δ implies |fk (x) − M | < . Then,
39
if |x − c| < δ and fk (x) > M , we have |hk (x) − hk (c)| = |M − M | < .
If −M < fk (x) ≤ M , we have |hk (x) − hk (c)| = |fk (x) − M | < . The
case where fk (c) = −M is similar. Thus hk is continuous for each k. For
each x ∈ [a, b], {fk (x)} converges to f (x). If |f (x)| < M , there is some K
such that, if k > K, then |fk (x)| < M . Thus {hk (x)} converges to f (x). If
f (x) = M , then for any > 0, there is some K such that, if k > K, then
|fk (x) − K| < . Then, if fk (x) > M , |hk (x) − M | = 0, and, if fk (x) < M ,
then |hk (x) − M | = |fk (x) − M | < . We have shown that {hk } is a sequence
of continuous functions converging pointwise to f on [a, b], where each hk is
bounded by M .
As we mentioned earlier, we first prove the following result for series. Af-
terwards, we will relate our findings back to sequences.
∞
P
Theorem 3.8. Suppose that the series gj converges to h on [a, b], and,
j=1
for each j ≥ 1, there is a positive constant Nj such that |gj (x)| < Nj for all
∞
P
x ∈ [a, b], where Nj converges. If each gj is a Baire class one function,
j=1
then h is a Baire class one function.
∞
P
Proof. Suppose that gj converges to h on [a, b], where each gj is a Baire class
j=1
∞
P
one function and bounded by Nj > 0, where Nj = N . By the Weierstrass
j=1
M -test, the convergence is uniform. Each gj is bounded by Mj = Nj + 1/2j ,
∞
P
and Mj converges to M = N + 1. By Lemma 3.7, for each j ≥ 1, there
j=1
is a sequence of continuous functions {fj,i }∞
i=1 which converge pointwise to gj
on [a, b], and where each fj,i satisfies |fj,i (x)| < Mj for all x ∈ [a, b] and for
∞
P
all i ≥ 1. Note: to avoid confusion, we emphasize here that the series gj
j=1
converges uniformly to h, and, for each j, the sequence {fj,i }∞
i=1 converges
40
pointwise to gj . We claim that the sequence
( p
)∞
X
{f1,1 , f1,2 + f2,2 , f1,3 + f2,3 + f3,3 , . . .} = fj,p (9)
j=1 p=1
converges pointwise to h on [a, b]. Since each term in this sequence is a finite
sum of continuous functions, this will show that h is a Baire class one function.
∞
P
Fix x ∈ [a, b] and let > 0. Since the series Mj converges, there is a
j=1
positive integer Q such that, if n > Q, then
∞
X
Mj < . (10)
j=n
For each j ∈ {1, . . . , Q}, there is a positive integer Pj such that, if p > Pj ,
then
|fj,p (x) − gj (x)| < /n. (11)
n n ∞
X X X
fj,n (x) − h(x) = fj,n (x) − gj (x)
j=1 j=1 j=1
Q Q
∞ n
X X X X
≤ fj,n (x) − gj (x) + gj (x) + fj,n (x)
j=1 j=1 j=Q+1 j=Q+1
n
X ∞
X n
X
≤ |fj,n (x) − gj (x)| + |gj (x)| + |fj,n (x)|
j=1 j=n+1 j=Q+1
< 3.
This shows that the sequence in (9), when evaluated at a fixed value of
x, indeed converges to h(x). This suffices to show that h is a Baire class one
function.
41
We now prove a similar result for sequences of functions. Corollary 3.9 will
be instrumental in our main result (Theorem 3.13).
Since g1 is also bounded by some M > 0, it follows that h is a Baire class one
function.
We now can begin working towards the main result of this section, which
gives a necessary and sufficient condition for a function to be Baire class one.
Recall that Theorem 2.11 gave a necessary condition; it remains to show that,
if the preimage of any open set under f is an open set, then f is a Baire
class one function. To that end, we will partition the range of f , rather than
the domain, into open intervals. The preimage of these open intervals are Fσ
sets, and we will construct a sequence converging to f by using these sets.
Readers familiar with concepts like Lebesgue measure, Lebesgue integration,
and simple sets may find some of these concepts familiar. As our discussion
of Fσ sets concerns unions of closed sets, the following canonical result will be
42
useful, and we will not prove it here.
n
X
f (x) = ak XAk (x),
k=1
∞
[
Ak = Bik
i=1
where each Bik is closed. Now define the function fj on a subset of [a, b] such
j
Bik . Note that fj is a piecewise function with j
S
that fj (x) = ak if x ∈
i=1
“pieces,” and it is defined on a finite union of closed sets. Then, since the
domain of fj is closed, by the Tietze Extension Theorem (Theorem 3.10), for
43
each j, there is a continuous function gj : [a, b] → R such that gj (x) = fj (x)
for any x in the domain of fj . We claim that the sequence {gj } converges
n
S
pointwise to f . Let x ∈ [a, b]. Since [a, b] ⊆ Ak , x must be contained in
k=1
J
Bik . It follows that, for
S
some Ak . Then there is an integer J such that x ∈
i=1
any j ≥ J, gj (x) = ak . Thus f is a Baire class one function.
We now turn our attention to the main proof of this section. We will first
give a necessary and sufficient condition for bounded functions to be Baire
class one functions, and then prove a similar result for all Baire class one
functions.
Proof. For convenience, this proof uses superscript when listing both sets and
real numbers; they do not represent exponents.
Theorem 2.11 shows that, if f is a Baire class one function, then the preim-
age of any open set under f is an Fσ set. It remains to prove the converse.
Bkn = f −1 ((yk−1
n n
, yk+1 n
)) = {x ∈ [a, b] : yk−1 n
< f (x) < yk+1 }.
44
By our hypothesis, each Bkn is an Fσ set. Define An1 = B1n and, for each integer
k−1
k ∈ {2, . . . , n − 1}, define Ank = Bkn \
S n
Al−1 . Since the set difference of two
l=1
Fσ sets is an Fσ set, the set {An1 , . . . , Ann−1 } contains disjoint Fσ sets whose
union contains [a, b]. For each positive integer n, define fn : [a, b] → R where
n−1
X
fn (x) = ykn XAnk (x).
k=1
By Lemma 3.11, each fn is a Baire class one function. We will show that {fn }
converges uniformly to f on [a, b].
2M
Let > 0 and pick a positive integer N such that N
< . Then let
x ∈ [a, b]. For some k ∈ {1, . . . , n − 1}, x ∈ Ak . Note that fn (x) = ykn and
n n
f (x) ∈ (yk−1 , yk+1 ). For any n ≥ N ,
2M 2M
|fn (x) − f (x)| = |ykn − f (x)| < ≤ < .
n N
45
Suppose that f : [a, b] → R is an unbounded real valued function for which
the preimage of any open set is an Fσ set. Define h : R → (0, 1) to be a con-
tinuous, strictly increasing function mapping the real line into (0, 1). Then
h ◦ f : [a, b] → (0, 1). If U is any open set, then, since h is continuous, h−1 (U )
is open, and thus, by our hypothesis, (h ◦ f )−1 (U ) = f −1 ◦ h−1 (U ) is an Fσ
set. Since h◦f is bounded, by Theorem 3.12, h◦f is a Baire class one function.
lim h−1 (fn (c)) = h−1 lim fn (c) = (h−1 ◦ h ◦ f )(c) = f (c)
n→∞ n→∞
As we have already seen, there are some functions which are not Riemann
integrable, but perhaps feel like they “should” be integrable. The simplest
example uses the so-called improper Riemann integral, such as the limit de-
scribing the area under the curve y = x−2/3 between x = 0 and 1. In Section
2.2, we discovered a function on a closed interval which is bounded, has an
antiderivative, and yet is not Riemann integrable. In this section, we present a
46
new integral, consistent with the Riemann integral on all Riemann integrable
functions, which allows us to integrate such derivatives.
Definition 4.1. Given the closed interval [a, b] and a positive valued function
δ : [a, b] → (0, ∞), a δ-fine tagged partition of [a, b] is a set
t
Pδ = {([xi−1 , xi ], ti ) : 1 ≤ i ≤ n}
Rb
In this section, we will use the notation (R) a f to denote the Riemann
Rb
integral of f , and, if not specified, we assume that a f denotes the Riemann
integral of f .
47
It follows easily from the definition that Riemann integrable functions are
Henstock integrable.
Proof. Because f is Riemann integrable, there is some L such that, for each
> 0, there is some δ0 > 0 such that, for any tagged partition tP of [a, b] with
norm smaller then δ0 , we have
|S(f,tP ) − L| < .
Then let δ : [a, b] → (0, ∞) be defined by δ(x) = δ0 /2. If tPδ is a δ-fine tagged
partition of [a, b], then for all 1 ≤ i ≤ n, [xi−1 , xi ] ⊆ (ti − δ0 /2, ti + δ0 /2), and
thus |tPn | < δ0 . It follows that |S(f,t Pn ) − L| < , and thus
Z b Z b
(R) f = L = (H) f,
a a
Theorem 4.4. For any interval [a, b] and function δ : [a, b] → (0, ∞), there
exists a δ-fine tagged partition of [a, b].
Proof. Let D be the set of all numbers x, where a < x ≤ b, for which there
is a δ-fine tagged partition of [a, x]. By taking x = a + δ(a)/2, we see that
48
{([a, a + δ(a)/2], a)} is a δ-fine tagged partition of [a, a + δ(a)/2]. Thus, D
is non-empty. Since D is bounded above by b, it has a least upper bound —
call it β. We claim that β ∈ D and β = b, and conclude that there is a δ-fine
tagged partition of [a, b].
Proof. Let f be differentiable on [a, b] and let > 0. For each x ∈ [a, b], define
δ(x) as a positive number for which, if 0 < |y − x| < d(x), then
f (y) − f (x) 0
y−x − f (x) < .
(12)
(The existence of such a δ(x) follows from the fact that f is differentiable at
49
x.) Note that we may rewrite (12) as
Then, using (15), if we take the Riemann sum of f 0 with the partition tPδ , we
find
Xn
S(f 0 ,tPδ ) − (f (b) − f (a)) = (xi − xi−1 )f 0 (ti ) − (f (xi ) − f (xi−1 ))
i=1
n
X
< (xi − xi−1 )
i=1
= (b − a).
Rb
Thus f 0 is Henstock integrable on [a, b] and (H) a
f 0 = f (b) − f (a).
50
every subinterval of [a, b]. Furthermore, for any [c, d] ⊆ [a, b],
Z d
(H) h0 = h(d) − h(c),
c
h2 cos h12
lim =0
h→0 h
1 2 1
2x cos 2 + sin 2 , if x 6= 0,
f 0 (x) = x x x
0,
if x = 0.
51
Note that f 0 is unbounded near the origin, and thus f 0 is not Riemann
integrable on [−a, a]. However, since f is even, we know from Theorem 4.5
Ra
that (H) a f 0 = 0. But how do two reasonable definitions of the integral arrive
at different conclusions? Let’s consider the function we are trying to integrate,
shown in Figure 5.
f 0 (x)
1 1
−√ √
π π
52
can ascribe meaning to this by writing
Z −r Z a
0 0
lim+ (R) f + (R) f = 0.
r→0 −a r
Ra
However, we still cannot write (R) −a f 0 = 0. The Henstock integral solves
this problem for us by allowing δ to vary throughout the domain. Rather than
using a fixed-value δ > 0, we use a positive-valued function δ to construct a
δ-fine tagged partition of the domain. Intuitively, one can imagine that, as we
approach the origin, the values of δ get smaller so as to closely estimate the
“waves” of the curve. Then, when we take a Riemann sum using this partition,
the “evenness” of the function come into play, leaving a value which converges
to 0. The question remains: how does this approach solve the issue at the
origin, with unbounded magnitude and oscillation? Let’s consider the interval
of our δ-fine tagged partition which contains the origin. Call it ([xi−1 , xi ], ti ).
Then ti cannot be nonzero, because we constructed δ in a such a way that
[xi−1 , xi ] closely estimated a part of the “wave” of the curve containing ti
(imagine that each interval is restricted to one single “hump” of the sinusoid).
Thus, we must have ti = 0. This gives us some clearer understanding for how
the Henstock integral can solve problems that the Riemann integral cannot.
In this paper, we developed the Henstock integral with the problem of integrat-
ing derivatives in mind. In Theorem 4.5, we showed that, with the Henstock
integral, we can indeed integrate all derivatives by calculating the difference of
the endpoints in the antiderivative. In Theorem 4.3, we saw that, for Riemann
integrable functions, the Riemann and Henstock integrals give the same result.
The Henstock integral thus solves a problem that the Riemann integral has,
53
and aligns with our intuition vis a vis the Fundamental Theorem of Calculus.
Then
Z b
(H) f = F (b) − F (a).
a
54
where tPδ = {([xi−1 , xi ], ti ) : 1 ≤ i ≤ n}. We now split tPδ into two sets, based
on whether or not F is differentiable at ti . That is, define tRδ as the set
t
Rδ = {([yi−1 , yi ], ri ) : 1 ≤ i ≤ p},
t
Sδ = {([zi−1 , zi ], si ) : 1 ≤ i ≤ q},
Recall the inequality in line (15) of the proof of Theorem 4.5, which says
that, for 1 ≤ i ≤ p,
|(yi − yi−1 )f (ti ) − (F (yi ) − F (yi−1 ))| < (yi − yi−1 ). (16)
Then note that, because of the continuity of F and the way we defined δ,
together with the fact that f (ti ) = 0, for 1 ≤ i ≤ q,
|(zi − zi−1 )f (ti ) − (F (zi ) − F (zi−1 ))| = |F (zi ) − F (ti ) + F (ti ) − F (zi−1 )|
55
Incorporating (16) and (17),
X n
t
|S(f, Pδ ) − (F (b) − F (a))| = (xi − xi−1 )f (ti ) − (F (xi ) − F (xi−1 ))
i=1
p
X
≤ (yi − yi−1 )f (ti ) − (F (yi ) − F (yi−1 ))
i=1
q
X
+ (zi − zi−1 )f (ti ) − (F (zi ) − F (zi−1 ))
p i=1 q
X X
≤ (yi − yi−1 ) +
k−1
2
i=1 i=1
≤ (b − a + 2).
Rb
Therefore f is Henstock integrable on [a, b], and (H) a f = F (b) − F (a).
56
References
57