Sunteți pe pagina 1din 28

Topics on Mean Value Theorems

Gen-Bin Huang
January 19, 2001
Abstract
Let f be a continuous function on [a, x], then there exists a
x
in [a, x] such
that

x
a
f(t)dt = f(
x
)(x a) .
This is the mean value theorem for integrals. We call
x
an integral mean point
of f on [a, x]. Recently, there are a number of studies on the location of
x
.
In particular, Zhang [7] showed that if f is C
r
(r N) near a point a, with
f

(a) = f

(a) = . . . = f
(r1)
(a) = 0 but f
(r)
(a) = 0, then
lim
xa

x
a
x a
=
1
(r + 1)
1
r
.
Note that there is also a dierential mean point c
x
in the classical mean value the-
orem. Using a similar method, I give a detailed estimation of the location of c
x
.
Since the classical mean value theorem implies the mean value theorem for inte-
grals, my result is more general. In fact the two theorems are roughly equivalent.
There is also some relationship between the Cauchy mean value theorem and the
generalized mean value theorem for integrals. In 1999, Schwind-Ji-Koditschek
generalized Zhangs result to consider the integral mean point of the generalized
mean value theorem, assuming f to behave like K + C
1
(x a)
r
near a (r > 1)
and g to behave like C
2
(x a)
s
near a. I also proved an analogue for its relative,
the Cauchy mean point.
On the other hand, Tong and Braza [6] studied the converse of the classical
mean value theorem. They showed that if f

(c) is not a local extremum, then c


is a dierential mean point. I obtain a parallel theorem for the converse of the
mean value theorem for integrals.
2
Contents
1 Introduction 1
1.1 The mean value theorems . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Relationships among the mean value theorems . . . . . . . . . . . 4
2 Estimation of integral mean points and dierential mean points 7
2.1 Integral mean points . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Dierential mean points . . . . . . . . . . . . . . . . . . . . . . . 13
3 Converses of mean value theorems 18
i
Chapter 1
Introduction
1.1 The mean value theorems
First, we introduce the mean value for integrals. Assume that f is continuous,
then
=
1
b a

b
a
f(t)dt =

b
a
f(t)dt

b
a
dt
.
This is called the arithmetic average or the mean value of f in the interval
[a, b]. We state the mean value theorem for integrals:
Theorem 1.1 ([1, p.258] Mean value theorem for integrals).
If f is a continuous function on [a, x], then there exists a number
x
in [a, x] such
that

x
a
f(t)dt = f(
x
)(x a) . (1.1)
This is a simple but very important mean value theorem of integral calculus.
It states that the mean value theorem of a continuous function in an interval
belongs to the range of the function. And it asserts only the existence of at least
1
one
x
in the interval for which f(
x
) is equal to the average value of f but gives
no further information about the location
x
. In this thesis, we call
x
an integral
mean point of f on [a, x].
Instead of the simple arithmetic average we can form the weighted averages:
=

x
a
f(t)g(t)dt

x
a
g(t)dt
g is the weight function and g(t) 0. We give the generalized mean value theorem
for integrals:
Theorem 1.2 ([1, p.257] Generalized mean value theorem for integrals).
If f is a continuous function on [a, x], g is integrable on [a, x] and g 0, then
there exists a number
x
in [a, x] such that

x
a
f(t)g(t)dt = f(
x
)

x
a
g(t)dt . (1.2)
Obviously, Theorem 1.1 is the special case when g(t) = 1.
Theorem 1.3 ([1, p.197] Classical mean value theorem).
Suppose f is continuous on the closed interval [a, x] and dierentiable on (a, x),
then for some c
x
between [a, x], we have
f

(c
x
) =
f(x) f(a)
x a
. (1.3)
We could estimate c
x
as
x
in the same way as in (1.2). And we call c
x
a
dierential mean point of f on [a, x].
Theorem 1.4 ([1, p.198] Cauchys mean value theorem).
Suppose f, g are continuous on [a, x], dierentiable on (a, x). If g

(t) = 0 for any


t (a, x), then there exists c
x
(a, x) such that
f(x) f(a)
g(x) g(a)
=
f

(c
x
)
g

(c
x
)
. (1.4)
2
Here we call c
x
a Cauchy dierential mean point or simply Cauchy mean point.
Recently, there are a number of studies on the location of the integral mean
point
x
, as x a. B. Zhang [7], improving B. Jacobsons result [4], showed that
if f is r (r N) times dierentiable at a, with f

(a) = f

(a) = . . . = f
(r1)
(a) = 0
but f
(r)
(a) = 0, then
lim
xa

x
a
x a
=
1
(r + 1)
1
r
.
Schwind-Ji-Koditschek [5] went on further. They allow f and g to have a singu-
larity at a. Namely they showed that if
lim
ta
f(t) K
(t a)
r
= C
1
and
lim
ta
g(t)
(t a)
s
= C
2
,
where r = 0,s > 1, r + s > 1, C
1
, C
2
= 0, then
lim
xa

x
a
x a
= (
s + 1
r + s + 1
)
1
r
.
On the other hand, Tong and Braza [6] studied the converse of the classical
mean value theorem. They showed that if f

(c) is not a local extremum, that c


is a dierential mean point.
It seems to be well-known that the four mean value theorems above are in-
terrelated. In fact, Theorem 1.3 implies Theorem 1.1. Under some additional
assumptions, Theorems 1.1 also implies Theorem 1.3. Similarly, Theorem 1.2
and Theorem 1.4 are roughly equivalent. Thus it is natural to study the esti-
mates of the dierential mean point and the converse of the mean value theorem
for integrals.
3
In the next section, we shall discuss the relationships among the mean value
theorems. In Chapter 2, we shall rst discuss the works of B. Jacobson, B. Zhang
and Schwind-Ji-Koditschek on the limiting position of the integral mean point

x
. Then we shall prove corresponding theorems for the dierential mean point
c
x
of the classical mean value theorem and Cauchy mean value theorem.
In Chapter 3, we shall discuss the converse of the classical mean value theorem
studied by Tong and Braza. Then, we prove a parallel theorem for the mean value
theorem for integrals.
1.2 Relationships among the mean value theo-
rems
It is known that Theorem 1.3 implies Theorem 1.1. There is a proof in the
Calculus book written by Campbell and Dierker [3].
Proof of Theorem 1.1 ([3, p.209]).
Suppose f : [a, x] R is continuous on [a, x]. Dene F(t) =

t
a
f(s)ds. Then F
is continuous on [a, x] and F

= f on [a, x]. Hence by Theorem 1.3, there is some


c
x
(a, x) such that
F(x) F(a) = F

(c
x
)(x a) .
That means,

x
a
f(t)dt = f(c
x
)(x a) .

Conversely, suppose F is C
1
and F

= f. So
F(x) F(a) =

x
a
f(t)dt .
4
By Theorem 1.1 and above formula, we conclude that
F(x) F(a) =

x
a
f(t)dt = f(
x
)(x a) .
Since F is a primitive of f, we could replace f(
x
) by F

(
x
) and obtain
F

(
x
) =
F(x) F(a)
x a
.
It is the same form as (1.4). Thus if F is C
1
, then Theorem 1.1 implies Theorem
1.3.

Assume g is continuous and g > 0, then Theorem 1.4 implies Theorem 1.2.
The proof is as follows:
Dene
F(t) =

t
a
f(s)g(s)ds, G(t) =

t
a
g(s)ds .
F and G are continuous on [a, x] and dierentiable on (a, x). Also
G(x) =

x
a
g(t)dt > G(a) = 0 .
Hence we apply Theorem 1.4 to obtain some c
x
(a, x) such that
F(x) F(a) =
F

(c
x
)
G

(c
x
)
[G(x) G(a)] .
Thus

x
a
f(t)g(t)dt =
f(c
x
)g(c
x
)
g(c
x
)

x
a
g(t)dt
= f(c
x
)

x
a
g(t)dt.
5

Conversely, under additional assumption on f and g, then Theorem 1.2 implies


Theorem 1.4. If f and g are both C
1
, let =
f

continuous, = g

> 0 for
simplicity, then by Theorem 1.2, there exists c
x
(a, x) such that

x
a
(t)(t)dt = (c
x
)

x
a
(t)dt .
Therefore,

x
a
f

(t)dt =
f

(c
x
)
g

(c
x
)

x
a
g

(t)dt .
So
f(x) f(a)
g(x) g(a)
=
f

(c
x
)
g

(c
x
)
.

6
Chapter 2
Estimation of integral mean
points and dierential mean
points
2.1 Integral mean points
In this section, we study the limiting position of an integral mean point
x
.
Jacobson [4] described the following theorem may fall into the category of inter-
esting facts we once knew, but have now forgotten. The theorem says that as x
approaches a, the value of
x
approaches the midpoint between a and x.
Theorem 2.1 ([4]).
If f is continuous on [a, x], dierentiable at a, f

(a) = 0 and
x
is taken as in
Theorem 1.1, then
lim
xa

x
a
x a
=
1
2
.
7
Before proving Theorem 2.1, we give Lemma 2.2:
Lemma 2.2 If (t) 0 as t a and (t)(t a)
m
is integrable, then
(a) lim
xa

x
a
(t)(ta)
m
dt
(xa)
m+1
= 0 ;
(b) lim
xa
(
x
)(
x
a)
m
(xa)
m
= 0, where m > 0 .
Proof.
(a) Given any > 0, there exist some x
0
> 0 such that for any 0 < |t a| < x
0
,
|(t)| < . Then take any x such that 0 < |x a| < x
0
. Without loss of
generality, let x > a. Then
|

x
a
(t)(t a)
m
dt
(x a)
m+1
|

x
a
(t a)
m
dt
(x a)
m+1
=

m + 1
.
(b) Since a <
x
< x, 0 <

x
a
xa
< 1. Therefore,
|
(
x
)(
x
a)
m
(x a)
m
| |(
x
)|
and so converges to 0 as x a.

Proof of Theorem 2.1.


By Taylors expansion,
f(t) = f(a) + f

(a)(t a) + (t)(t a) , (2.1)


where (t) 0 as t a. Integrate (2.1) from a to x, and obtain

x
a
f(t)dt = f(a)(x a) +
f

(a)(x a)
2
2!
+

x
a
(t)(t a)dt . (2.2)
8
Equation (2.1) evaluated at
x
(a <
x
< x) yields
f(
x
) = f(a) + f

(a)(
x
a) + (
x
)(
x
a),
where (
x
) 0 as
x
a. Thus,
f(
x
)(x a) = f(a)(x a) + f

(a)(
x
a)(x a)
+(
x
)(
x
a)(x a) . (2.3)
From (1.1), (2.2) and (2.3) we have
f

(a)(x a)
2
2!
+

x
a
(t)(t a)dt
= f

(a)(
x
a)(x a) + (
x
)(
x
a)(x a) .
Hence,
f

(a)

x
a
x a
=
f

(a)
2
+

x
a
(t)(t a)dt
(x a)
2

(
x
)(
x
a)
x a
. (2.4)
By Lemma 2.2,
lim
xa

x
a
(t)(t a)dt
(x a)
2
= lim
xa
(
x
)(
x
a)
x a
= 0 .
From (2.4) we obtain lim
xa

x
a
xa
=
1
2
.

We consider the case f

(a) = 0 in Theorem 2.1 and obtain Theorem 2.3.


Theorem 2.3 ([7]).
If f is continuous on [a, x], and is twice dierentiable at a, with f

(a) = 0,
f

(a) = 0. If
x
is taken as in Theorem 1.1, then
lim
xa

x
a
x a
=
1

3
.
9
In a similar way, we can establish the following general result.
Theorem 2.4 ([7]).
If f is continuous on [a, x], and is r times (r > 1, r N) dierentiable at a, with
f

(a) = f

(a) = . . . = f
(r1)
(a) = 0 but f
(r)
(a) = 0. If
x
is taken as in Theorem
1.1, then
lim
xa

x
a
x a
=
1
(r + 1)
1
r
.
If we consider a large function class, which need not assume f to be dieren-
tiable, to estimate
x
. And we take a transformation of f.
Theorem 2.5 ([5]).
If f is continuous on (a, x] and g is integrable on (a, b) with g(t) 0 for t (a, b).
If lim
ta
f(t)K
(ta)
r
= C
1
, lim
ta
g(t)
(ta)
s
= C
2
for some K , C
1
, C
2
= 0, r = 0 and
s > 1 such that r + s > 1, then
(a) there exists
x
(a, x] such that

x
a
f(t)g(t)dt = f(
x
)

x
a
g(t)dt , (2.5)
(b) for any choice of
x
, we have
lim
xa

x
a
x a
= (
s + 1
r + s + 1
)
1
r
. (2.6)
Proof.
(a) Since both
lim
ta
f(t) K
(t a)
r
and
lim
ta
g(t)
(t a)
s
10
exist, then given the limits
f(t) = K + C
1
(t a)
r
+
1
(t)(t a)
r
, (2.7)
and
g(t) = C
2
(t a)
s
+
2
(t)(t a)
s
, (2.8)
where
1
(t) 0,
2
(t) 0. First, from equations (2.7) and (2.8), we have
|f(t)g(t)| = |[K(C
2
+
2
(t))](t a)
s
+ [C
1
C
2
+ C
1

2
(t) + C
2

1
(t) +
1
(t)
2
(t)](t a)
r+s
|
|K(C
2
+
2
(t))|(t a)
s
+|C
1
C
2
+ C
1

2
(t) + C
2

1
(t) +
1
(t)
2
(t)|(t a)
r+s
K
1
(t a)
s
+ K
2
(t a)
r+s
where K
1
, K
2
are some constant, then

x
a
f(t)g(t)dt exists. Now,we give the
proof of (a).
Case 1: r > 0.
Assume F(a) = K, and F(t) = f(t) for t (a, b], then F is continuous on
[a, b]. Since g is integrable and g(t) 0, by applying Theorem 1.2, there
exists
x
(a, x], such that

x
a
f(t)g(t)dt =

x
a
F(t)g(t)dt = f(
x
)

x
a
g(t)dt .
Case 2: r < 0.
Without loss of generality, we take C
1
and

x
a
g(t)dt > 0. From (2.7), then
lim
ta
f(t) = , and there exists > 0 such that
f(t) >

x
a
f(t)g(t)dt

x
a
g(t)dt

for t (a, a + ). Assume that there exists
x
satisfying (2.5), but
x

(a, a + ). Take a
1
(a, a + ), f is still continuous on [a
1
, x], f(t) > on
11
(a, a
1
), then
min
at<x
f(t) = min
a
1
tx
f(t);
min
a
1
t<x
f(t) = min
at<x
f(t) < f(a
1
) max
a
1
t<x
f(t) .
Therefore, there exists
x
between a and x such that
f(
x
) = =

x
a
f(t)g(t)dt

x
a
g(t)dt
.
(b) Since

x
a
f(t)g(t)dt =

x
a
K[C
2
(t a)
s
+
2
(t)(t a)
s
]dt + C
1
C
2
(x a)
r+s+1
r + s + 1
+C
2

x
a

1
(t)(t a)
r+s
dt + C
1

x
a

2
(t)(t a)
r+s
dt
+

x
a

1
(t)
2
(t)(t a)
r+s
dt
Also,
f(
x
)

x
a
g(t)dt =

x
a
K[C
2
(t a)
s
+
2
(t)(t a)
s
]dt + C
1
C
2
(
x
a)
r
(x a)
s+1
s + 1
+C
2

1
(
x
)(
x
a)
r
(x a)
s+1
s + 1
+ C
1
(
x
a)
r

x
a

2
(t)(t a)
s
dt
+
1
(
x
)(
x
a)
r

x
a

2
(t)(t a)
s
dt
Hence by (a), Lemma 2.2, we obtain the following result
C
1
C
2
(x a)
r+s+1
r + s + 1
(1 + o(1)) = C
1
C
2
(
x
a)
r
(x a)
s+1
s + 1
(1 + o(1)) .
Then we have the result
s + 1
r + s + 1
= lim
xa
(

x
a
x a
)
r
.

Remark: Obviously, Theorem 2.4 is a special case of Theorem 2.5.


12
2.2 Dierential mean points
We have the same results as Theorem 2.4 of the classical mean value theorem.
Theorem 2.6 If f is continuous on [a, b] and dierentiable on (a, b), f

is Rie-
mann integrable and lim
ta
f

(t)l
(ta)
r
= C
1
exists in (r = 0, r > 1), then the
dierential mean point c
x
satises
lim
xa
c
x
a
x a
= (
1
1 + r
)
1
r
.
Proof.
Since lim
ta
f

(t)l
(ta)
r
= C
1
,
f

(t) = l + C
1
(t a)
r
+ (t)(t a)
r
(2.9)
where (t) 0 as t a. On the other hand, equation (2.9) evaluated at c
x
yields
f

(c
x
) = l + C
1
(c
x
a)
r
+ (c
x
)(c
x
a)
r
(2.10)
where (c
x
) 0 as c
x
a. Multiply (x a) to (2.10) we get
f

(c
x
)(x a) = l(x a) + C
1
(c
x
a)
r
(x a) + (c
x
)(c
x
a)
r
(x a) . (2.11)
Also,
f(x) f(a) =

x
a
f

(t)dt
= l(x a) + C
1
(x a)
r+1
r + 1
+

x
a
(t)(t a)
r
dt . (2.12)
Equating (2.11) and (2.12), we have
C
1
(c
x
a)
r
(x a) + (c
x
)(c
x
a)
r
(x a) = C
1
(x a)
r+1
r + 1
+

x
a
(t)(t a)
r
dt
13
Take x a, then we get the following result by Lemma 2.2
lim
xa
c
x
a
x a
= (
1
1 + r
)
1
r
.

If we consider f is in C
2
[a, b], we can derive a more detailed asymptotic
expression for c
x
.
Theorem 2.7 If f is C
2
[a, b], and
f

(t) = l + C
1
(t a)
r
+ C
2
(t a)
r+1
+ (t)(t a)
r+1
, (2.13)
where C
1
= 0, r > 0. Then l = f

(a), and
(
c
x
a
x a
)
r
=
1
r + 1

1 +
C
2
C
1
[
r + 1
r + 2
(
1
r + 1
)
1
r
](x a)

+ o(c
x
a) .
Proof.
First by Taylors theorem, there is some between a and t such that
f

(t) = f

(a) + f

()(t a) .
Compare with (2.12) and take limit t a, we obtain l = f

(a). Then multiply


(x a) to (2.13) we get
f

(t)(x a) = f

(a)(x a) + C
1
(t a)
r
(x a) + C
2
(t a)
r+1
(x a)
+(t)(t a)
r+1
(x a) (2.14)
where (t) 0, as t a. On the other hand, equation (2.14) evaluated at c
x
yields
f

(c
x
)(x a) = f

(a)(x a) + C
1
(c
x
a)
r
(x a) + (c
x
)(c
x
a)
r
(x a)
+C
2
(c
x
a)
r+1
(x a) + (c
x
)(c
x
a)
r+1
(x a) (2.15)
14
where (c
x
) 0, as c
x
a.
Also,
f(x) f(a) =

x
a
f

(t)dt
= f

(a)(x a) + C
1
(x a)
r+1
r + 1
+ C
2
(x a)
r+2
r + 2
+

x
a
(t)(t a)
r+1
dt. (2.16)
Equating (2.15) and (2.16), we have
(
c
x
a
x a
)
r
[1 +
C
2
C
1
(c
x
a) +
1
C
1
(c
x
)(c
x
a)]
=
1
r + 1
+
C
2
C
1
(
x a
r + 2
) +
1
C
1
(x a)
r+1

x
a
(t)(t a)
r+1
dt .
By Theorem 2.6 and Taylors expansion, we have
(
c
x
a
x a
)
r
= [1
C
2
C
1
(c
x
a) + o(c
x
a)][
1
r + 1
+
C
2
C
1
(
x a
r + 2
)
+
1
C
1
(x a)
r+1

x
a
(t)(t a)
r+1
dt]
= [1
C
2
C
1
x a
(1 + r)
1
r
+ o(c
x
a)][
1
r + 1
+
C
2
C
1
(
x a
r + 2
)
+o(x a)
=
1
r + 1
{1 +
C
2
C
1
[
r + 1
r + 2
(
1
r + 1
)
1
r
](x a)} + o(x a) .

Theorem 2.8 Suppose f, g are continuous on [a, b], and dierentiable on (a, b),
such that f

, g

are integrable over [a, b] and g

(x) = 0 for all x (a, b). If


lim
ta
g

(t)
(t a)
s
= C
2
;
and
lim
ta
f

(t)
g

(t)
K
(t a)
r
= C
1
15
where C
1
, C
2
= 0, and r = 0, s > 1, r + s > 1. Then the Cauchy dierential
mean point c
x
satises
lim
xa
c
x
a
x a
= (
s + 1
r + s + 1
)
1
r
.
Proof.
Since
g

(t) = C
2
(t a)
s
+
2
(t)(t a)
s
and
f

(t)
g

(t)
= K + C
1
(t a)
r
+
1
(t)(t a)
r
(2.17)
where
1
(t),
2
(t) 0 as t a, we have
f

(t) = [K + C
1
(t a)
r
+
1
(t)(t a)
r
]g

(t)
= KC
2
(t a)
s
+ C
1
C
2
(t a)
r+s
+ K
2
(t)(t a)
s
+[C
1

2
(t) + C
2

1
(t) +
1
(t)
2
(t)](t a)
r+s
.
On the other hand, equation (2.17) evaluated at c
x
yields
f

(c
x
)
g

(c
x
)
= K + C
1
(c
x
a)
r
+
1
(c
x
)(c
x
a)
r
.
Also,
f(x) f(a) =

x
a
f

(t)dt
=
KC
2
(x a)
s+1
s + 1
+
C
1
C
2
(x a)
r+s+1
r + s + 1
+ K

x
a

2
(t)(t a)
s
dt
+

x
a
[C
1

2
(t) + C
2

1
(t) +
1
(t)
2
(t)](t a)
r+s
dt
and
g(x) g(a) =

x
a
g

(t)dt
=
C
2
(x a)
s+1
s + 1
+

x
a

2
(t)(t a)
s
dt .
16
Hence by (1.4) and Lemma 2.2, we obtain the following result
C
1
C
2
(x a)
r+s+1
r + s + 1
+

x
a
[C
1

2
(t) + C
2

1
(t) +
1
(t)
2
(t)](t a)
r+s
dt
=
C
1
C
2
+ C
2

1
(c
x
)
s + 1
(c
x
a)
r
(x a)
s+1
+ [c
1
+
c
(c
x
)](c
x
a)
r

x
a

2
(t)(t a)
s
dt .
Then we have
lim
xa
c
x
a
x a
= (
s + 1
r + s + 1
)
1
r
.

17
Chapter 3
Converses of mean value
theorems
It would be interesting to study the converse of the classical mean value
theorem. That is, if c is a point in [a, b], can c be a dierential mean point of
f over some sub-interval of [a, b] ? The following example shows the converse
problem does not always hold.
Example: If f(t) = t
3
on [1, 1] and [t
1
, t
2
] (1, 1), then
f(t
2
)f(t
1
)
t
2
t
1
=
t
2
2
+ t
2
t
1
+ t
2
1
> 0 for all t
1
= t
2
, but f

(0) = 0. Therefore (1.4) does not hold.


Tong and Braza [6] give the following theorem to show the converse problem
may fail at extremum values of f

and at certain accumulation points.


Denition 3.1 We say a function f is locally linear (horizontal) about a point
c in its domain I if there is some open interval (, ) I containing c such that
f(t) |
(,)
is a linear (constant) function.
18
Theorem 3.2 ([6]).
If f is continuous on [a, b] and dierentiable on (a, b), and c is given in (a, b).
Then
(a) Weak form: If f

(c) = sup{f

(t)|t (a, b)} and f

(c) = inf{f

(t)|t (a, b)},


then there is some interval (a
1
, b
1
) (a, b) such that f

(c) = (f(b
1
)
f(a
1
))/(b
1
a
1
).
(b) Strong form: If f

(c) is not a local extremum value of f

on (a, b), and


c is not an accumulation point of A
c
= {t (a, b)|f

(t) = f

(c)}, then
there is a sub-interval (a
1
, b
1
) (a, b) such that c (a, b) and f

(c) =
(f(b
1
) f(a
1
))/(b
1
a
1
).
(c) If f

(c) is a local extremum value of f

on (a, b) (i.e, f

(c) is a total extremum


value of f

on some sub-interval (a

, b

) (a, b) containing c), then f is


either locally linear about c or f

(c) = (f() f())/( ) wherever


< c < and (, ) (a

, b

).
We consider the converse of mean value theorem for integrals. If we give a
number c is in (a, b), could we nd an interval (a
1
, b
1
) (a, b) such that c is an
integral mean point on (a
1
, b
1
), that is,
f(c) =
1
b
1
a
1

b
1
a
1
f(t)dt ? (3.1)
We give a result similar to Theorem 3.2.
Theorem 3.3 If f is continuous on [a, b], and c is given in (a, b). Then
(a) Weak form: If f(c) = sup{t (a, b)|f(t)} and f(c) = inf{t (a, b)|f(t)},
then there is some interval (a
1
, b
1
) (a, b) such that (3.1) hold.
19
(b) Strong form: If f(c) is not a local extremum value of f on (a, b), and c is
not an accumulation point of A
c
= {t (a, b)|f(t) = f(c)}, then there is a
sub-interval (a
1
, b
1
) (a, b) such that c (a, b) and (3.1) hold.
(c) If f(c) is a local extremum value of f on (a, b) (i.e. f(c) is a total extremum
value of f on some sub-interval (a

, b

) (a, b) containing c), then f is


either locally horizontal about c or f(c) =

f(t)dt/( ) wherever <


c < and (, ) (a

, b

).
Proof.
(a) If f(c) is not a total extremum value on (a, b), then there are c
1
, c
2
(a, b)
(assume c
1
< c
2
) such that f(c
1
) < f(c) < f(c
2
). Since
f(c
i
) = lim
x,yc
i
1
y x

y
x
f(t)dt
for i = 1, 2, there are sub-intervals (x
1
, y
1
), (x
2
, y
2
) (a, b) such that
1
y
1
x
1

y
1
x
1
f(t)dt < f(c) <
1
y
2
x
2

y
2
x
2
f(t)dt .
Without loss of generality, suppose x
1
< x
2
. Consider the value K =
1
y
2
x
1

y
2
x
1
f(t)dt. If K = f(c), we are done. We need consider only the cases
K > f(c) and K < f(c).
(1) K > f(c). Since g(y) =
1
yx
1

y
x
1
f(t)dt is continuous on (x
1
, b) and
g(y
1
) < f(c) < g(y
2
), there is a point y between y
1
and y
2
such that
g( y) = f(c).
(2) K < f(c). Since h(x) =
1
y
2
x

y
2
x
f(t)dt is continuous on (a, y
2
) and
h(x
1
) < f(c) < h(x
2
), there is a point x between x
1
and x
2
such that
h( x) = f(c).
20
(b) If f(c) is not a local extremum of f on (a, b), then there is a sub-interval
(a

0
, b

0
) (a, b) such that c (a

0
, b

0
) and f(c) is not a total extremum value
of f on (a

0
, b

0
). Let
a

i+1
= (a

i
+ c)/2, b

i+1
= (b

i
+ c)/2
for i = 0, 1, 2, 3... Then
a

0
< a

1
< ... < a

i
< ... < c < ... < b

i
< ... < b

1
< b

0
and lim
i
a

i
= lim
i
b

i
. By part (a), for each sub-interval (a

i
, b

i
)
(a

0
, b

0
), (i > 0), there are a
i
, b
i
such that (a
i
, b
i
) (a

i
, b

i
) and f(c) =

b
i
a
i
f(t)dt/(b
i
a
i
). If c (a
i
, b
i
) for some i, the theorem is proved. Hence
we suppose c / (a
i
, b
i
) for all i N. By Theorem 1.1 for f on [a
i
, b
i
], there
is some c
i
(a
i
, b
i
) such that f(c
i
) =
1
b
i
a
i

b
i
a
i
f(t)dt. Hence f(c) = f(c
i
).
Notice that since c
i
(a
i
, b
i
) (a

i
, b

i
), lim
i
(b

i
a

i
) = 0, and c
i
= c,
these c
i
cannot coincide innitely often. This implies there is an innite
discrete sequence c
ik
such that lim
k
c
ik
= c. This is contradiction since
it implies that c is an accumulation point of the set A
c
= {t (a, b)|f(t) =
f(c)}.
(c) Let f(c) = 0 be a local maximum of f since we can replace f by f(t)f(c).
There is a sub-interval (, ) (a

, b

) such that < c < and


f(c) =
1

f(t)dt .
So

f(t)dt = 0 and f(t) 0 for all t (, ). As f is continuous, f(t) 0


on (, ). Thus f is locally horizontal about c.

21
Alternative proof.
Dene F(x) =

x
a
f(t)dt. Then F is continuous on [a, b] and F

(x) = f(x) for any


x (a, b). Then
f(c) = sup{t (a, b)|f(t)} (= inf{t (a, b)|f(t)})
if and only if
F

(c) = sup{F

(t)|t (a, b)} (= inf{F

(t)|t (a, b)}) .


Thus, by Theorem 3.2 (a), there exists some interval (a
1
, b
1
) (a, b) such that
F

(c) =
F(b
1
) F(a
1
)
b
1
a
1
,
that is
f(c) =
1
b
1
a
1

b
1
a
1
f(t)dt .
Part (a) is proved. Similarly, part (b) and part (c) also follow from Theorem 3.2
(b) and (c) respectively.

Theorem 3.2 shows that the converse to the classical mean value theorem may
fail at extremum values of f

(x) and at certain accumulation points. In [6], Tong


and Braza also studied the number of bad points (i.e. points which are neither
strong nor weak dierential mean points). They gave two examples. The below
is one of them.
22
Example: Dene
f(x) =

x
3
, x [0, 1/4]
2(1/4)
3
+ (x 1/2)
3
, x [1/4, 5/8]
2(1/4)
3
+ 2(1/8)
3
+ (x 3/4)
3
, x [5/8, 13/16]
...
2[(1/4)
3
+ (1/8)
3
+ ... + (
1
2
k+1
)
3
] + (x
2
k
1
2
k
)
3
, x [
2
k+1
3
2
k+1
,
2
k+2
3
2
k+2
]
...
2

k=2
(
1
2
k
)
3
+ (x 1)
3
. x [1, 2]
The function f is continuous on [0, 2] and dierentiable on (0, 2). It is monoton-
ically increasing, with f

(
2
k
1
2
k
) = 0 for k = 0, 1, 2, and f

(1) = 0, but
f(x) f(y)
x y
> 0
for x, y [0, 2], with x = y. Thus f has a countable number of bad points.
Could these bad points be dense? Is it possible to have an uncountable number
of such points? In [2], Borwein and Wang answered these questions armatively.
They gave an example of a function having an uncountable number of bad points.
These points are dense in the nondegenerate interval [a, b]. However the Lebesque
measure is still zero.
23
Bibliography
[1] R.G. Bartle and D.R. Sherbert, An Introduction to Real Analysis, second
edition, Wiley, New York (1992).
[2] J.M. Borwein and Xianfu Wang, The converse of the mean value theorem
may fail generically, Amer. Math. Monthly 105 (1998), 847-848.
[3] H.E. Campbell and P.F. Dierker, Calculus, second edition, Prindle, Weber
Schmidt (1978).
[4] B. Jacobson, On the mean value theorem for integrals, Amer. Math. Monthly
89 (1982), 300-301.
[5] W.J. Schwind, J. Ji, and D.E. Koditschek, A physically motivated further
note on the mean value theorem for integrals, Amer. Math. Monthly 106
(1999), 559-564.
[6] J. Tong and P.A. Braza, A converse of the mean value theorem, Amer. Math.
Monthly 104 (1997), 939-942.
[7] B. Zhang, A note on the mean value theorem for integrals, Amer. Math.
Monthly 104 (1997), 561-562.
24

S-ar putea să vă placă și