Skip to main content

Topics on Measurable Functions of Real Variables

  • Chapter
  • First Online:
Real Analysis

Part of the book series: Birkhäuser Advanced Texts Basler Lehrbücher ((BAT))

  • 4472 Accesses

Abstract

Let f be a real-valued function defined and bounded in some interval \([a,b]\subset \mathbb {R}\).

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 49.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 49.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Also in [122]. A proof independent of measure theory is in [134], pp. 5–9.

  2. 2.

    Also in [161].

  3. 3.

    If the modulus of continuity is of Lipschitz type, the theorem is in M.D. Kirzbraun [83]. The proof of Kirzbraun is rather general as it does include vector-valued functions. A simpler proof for scalar functions, is in McShane [106]. Pucci observed that the concavity of the modulus of continuity is sufficient to construct the extension. The proof has been taken from the 1974 lectures on Real Analysis by C. Pucci, at the Univ. of Florence, Italy.

  4. 4.

    First proved by Ascoli in [6] for equi-Lipschitz functions, and extended by Arzelà in [5] to a general family of equi-continuous functions.

  5. 5.

    This construction was suggested by V. Vespri and U. Gianazza.

  6. 6.

    [70], Chap. II, § 5 contains an alternate proof of this inequality that does not use Jensen’s inequality.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emmanuele DiBenedetto .

Appendices

Problems and Complements

1c Functions of Bounded Variations

1.1. :

The continuous function on [0, 1]

$$\begin{aligned} f(x)=\left\{ \begin{array}{ll} {\displaystyle x^2\cos \frac{\pi }{x}}\quad &{}\text { for }\> x\in (0,1]\\ 0&{} \text { for }\> x=0 \end{array} \right. \end{aligned}$$

is of bounded variation in [0, 1].

1.2. :

Let f be the continuous function defined in [0, 1] by

$$\begin{aligned} \begin{array}{cl} f(0)=0,\quad f\left( \frac{1}{2n+1}\right) =0,\quad f\left( \frac{1}{2n}\right) =\frac{1}{2n}\quad &{}\text { for all }\>n\in \mathbb {N}\\ f\>\text { is affine on the intervals }\> \left[ \frac{1}{m+1},\frac{1}{m}\right] \quad &{}\text { for all }\> m\in \mathbb {N}. \end{array} \end{aligned}$$

Such an f is not of bounded variation in [0, 1].

1.3. :

Prove Propositions 1.11.3.

1.4. :

Let \(\{f_n\}\) be a sequence of functions in [ab] converging pointwise in [ab] to f. Then \(\mathcal {V}_f[a,b]\le \liminf \mathcal {V}_{f_n}[a,b]\), and strict inequality may occur as shown by the sequence

$$\begin{aligned} f_n(x)=\left\{ \begin{array}{ll} 0\quad &{}\text { for }\> x=0\\ \frac{1}{2}\quad &{}\text { for }\> x\in (0,\frac{1}{n}]\\ 0\quad &{}\text { for }\> x\in (\frac{1}{n},1]. \end{array} \right. \end{aligned}$$
1.5. :

Let f be continuous and of bounded variation in [ab]. Then the functions \(x\rightarrow \mathcal {V}_f[a,x]\), \(\mathcal {V}_f^+[a,x]\), \(\mathcal {V}_f^-[a,x]\) are continuous in [ab].

1.6. :

Prove that the distribution function \(f_*(\cdot )\) of a measurable function f on a measure space \(\{X,\mathcal {A},\mu \} \), as defined by (15.1c) of Chap. 4, is of bounded variation in every interval \([a,b]\subset \mathbb {R}\).

1.1 1.1c The Function of the Jumps

Let f be of bounded variation in [ab] and regard it as defined in \(\mathbb {R}\), by extending f to be equal to f(a) on \((-\infty ,a]\) and equal to f(b) in \([b,\infty )\). Prove that the limits

$$\begin{aligned} f(x^{\pm })\buildrel \mathrm { def}\over {=}\lim _{h\rightarrow 0} f(x\pm h)\quad \text { for }\> h>0 \end{aligned}$$

exist for all \(x\in [a,b]\). The function of the jumps of f is defined by

$$\begin{aligned} J_f(x)=\mathop {\textstyle {\sum }}\limits _{a\le c_j\le x} \big [f(c_j^+)-f(c_j^-)\big ]. \end{aligned}$$

The difference \(f-J_f\) is continuous in [ab]. Also \(J_f\) is of bounded variation in [ab] and

$$\begin{aligned} \mathcal {V}_f[a,b]=\mathcal {V}_{f-J_f}[a,b]+\mathcal {V}_{J_f}[a,b]. \end{aligned}$$

Therefore a function f of bounded variation in [ab] can be decomposed into the continuous function \(f-J_f\) and \(J_f\). The latter bears the possible discontinuities of f in [ab].

1.7. :

Construct a non-decreasing function in [0, 1] which is discontinuous at all the rational points of [0, 1].

1.2 1.2c The Space BV[ab]

Let \([a,b]\subset \mathbb {R}\) be a finite interval and denote by BV[ab] the collection of all functions \(f:[a,b]\rightarrow \mathbb {R}\) of bounded variation in [ab]. One verifies that BV[ab] is a linear vector space. Also setting

$$\begin{aligned} d(f,g)=|f(a)-g(a)|+\mathcal {V}_{f-g}[a,b]\quad \text { for }\> f,g\in BV[a,b] \end{aligned}$$
(1.1c)

defines a distance in BV[ab] by which \(\{BV[a,b];d\}\) is a metric space.

For any two functions \(f,g\in BV[a,b]\)

$$\begin{aligned} \sup _{[a,b]}|f-g|\le |f(a)-g(a)|+\mathcal {V}_{f-g}[a,b]. \end{aligned}$$
(1.2c)
Fig. 1
figure 1

Cauchy sequence in the sup-norm and not in the BV-norm

Therefore a Cauchy sequence in BV[ab] is also Cauchy in the sup-norm. The converse is false as illustrated in Fig. 1c. The sequence \(\{f_n\}\) generated as in Fig. 1c is not a Cauchy sequence in the topology of BV[0, 1], while it is a Cauchy sequence in the topology of C[0, 1].

1.2.1 1.2.1.c Completeness of BV[ab]

Let \(\{f_n\}\) be a Cauchy sequence in BV[ab]. There exists \(f\in BV[a,b]\) such that \(\{f_n\}\rightarrow f\) in the topology of BV[ab].

The proof is in two steps. First one uses (1.2c) to identify the limit f. Then one proves that such an f is actually in BV[ab] by using that \(\{f_n\}\) is Cauchy in BV[ab]. As a consequence BV[ab] is a complete metric space.

2c Dini Derivatives

2.1. :

Compute \(D^+f(0)\) and \(D_+f(0)\) for the function in (1.1).

2.2. :

Let f have a maximum at some \(c\in (a,b)\). Then \(D^-f(c)\ge 0\).

2.3. :

Let f be continuous in [ab]. If \(D^+f\ge 0\) in [ab], then f is non-decreasing in [ab]. The assumption that f be continuous cannot be removed.

Hint: Fix \([\alpha ,\beta ]\subset (a,b)\), and by continuity extend the function f to be equal to \(f(\beta )\) for \(x\ge \beta \) and to be equal to \(f(\alpha )\) for \(x\le \alpha \). Having fixed \(\varepsilon >0\), the assumption implies that for each \(x\in [\alpha ,\beta ]\) there exists

$$\begin{aligned} 0<h_x=h(x,\varepsilon )<\textstyle {\frac{1}{2}}(b-\beta ) \end{aligned}$$

such that

$$\begin{aligned} f(x)<f(x+h_x)+ h\varepsilon . \end{aligned}$$

By continuity, such inequality continues to hold for all \(y\in (x-\delta _x,x+\delta _x)\) for some \(\delta _x>0\), which without loss of generality we may assume not to exceed \(\frac{1}{4} h_x\). The collection of these open intervals covers \([\alpha ,\beta ]\). From this select a finite sub-collection, say

$$\begin{aligned} (x_j-\delta _j,x_j+\delta _j)\quad \text { for }\>j=1,\dots ,n. \end{aligned}$$

To these are associated positive numbers \(h_j<\frac{1}{2}(\beta -b)\) such that

$$\begin{aligned} f(y)<f(y+h_j)+ h_j\varepsilon \quad \text { for all }\> y\in (x_j-\delta _j,x_j+\delta _j). \end{aligned}$$

Starting with \(y=\alpha \) and iterating this inequality \(m\le n\) times gives

$$\begin{aligned} f(\alpha )<f\Big (\alpha +\mathop {\textstyle {\sum }}\limits _{j=1}^m h_j\Big )+ \mathop {\textstyle {\sum }}\limits _{j=1}^mh_j\varepsilon . \end{aligned}$$

By construction, there exists \(m\le n\) such that

$$\begin{aligned} \alpha +\mathop {\textstyle {\sum }}\limits _{j=1}^m h_j\ge \beta \quad \text { and } \quad \mathop {\textstyle {\sum }}\limits _{j=1}^m h_j \le b-a. \end{aligned}$$

Therefore

$$\begin{aligned} f(\alpha )<f(\beta )+\varepsilon (b-a) \quad \text { for all }\>\varepsilon >0. \end{aligned}$$
2.4. :

Let \(f\in BV[a,b]\) and let \(J_f\) the function of the jumps of f introduced in § 1.1c. Prove that \(J_f^\prime =0\) a.e. in [ab].

Hint: Assume that f is non-decreasing, so that \(J_f\) is non decreasing. Denoting by \(\{c_n\}\) the sequence of the jumps of f

$$\begin{aligned} J_f(b)-J_f(a) =\mathop {\textstyle {\sum }}\limits _{a<c_j\le b}\big [f(c_j^+)-f(c_j^-)\big ]. \end{aligned}$$
(2.1c)

Fix \(m\in \mathbb {N}\) and let \(\varepsilon >0\) be so small that the intervals

$$\begin{aligned} (\alpha _j,\beta _j]=\Big (c_j-\frac{\varepsilon }{2m}\,,\, c_j+\frac{\varepsilon }{2m}\Big ]\mathop {\textstyle {\bigcap }}\limits (a,b] \quad \text { for }\> j=1,\dots ,m \end{aligned}$$

are disjoint. The complement of their union is the finite union of disjoint intervals \([\alpha _j^\prime ,\beta _j^\prime )\). For \(t>0\) assume that \(\mu ([DJ_f>t])>0\), and choose

$$\begin{aligned} \varepsilon <\textstyle {\frac{1}{2}}\mu ([DJ_f>t]). \end{aligned}$$

For such a choice estimate

$$\begin{aligned} \mu \Big ([DJ_f>t]\mathop {\textstyle {\bigcap }}\limits \mathop {\textstyle {\bigcup }}\limits _{j=1}^{m^\prime } [\alpha _j^\prime ,\beta _j^\prime ]\big )>\frac{1}{2}\mu ([DJ_f>t])>0 \end{aligned}$$

for a positive integer \(m^\prime \). By Proposition 2.2

$$\begin{aligned} J_f(\beta _j^\prime )-J_f(\alpha _j^\prime )\ge t\mu \big (\big [DJ_f>t\big ]\cap \big [\alpha _j^\prime ,\beta _j^\prime \big ]\big ). \end{aligned}$$

Therefore

$$\begin{aligned} \begin{aligned} J_f(b)-J_f(a)&=\mathop {\textstyle {\sum }}\limits _{j=1}^m\big [J_f(\beta _j)-J_f(\alpha _j)\big ] +\mathop {\textstyle {\sum }}\limits _{j=1}^{m+1} \big [J_f(\beta _j^\prime )-J_f(\alpha _j^\prime )\big ]\\&\ge \mathop {\textstyle {\sum }}\limits _{j=1}^m \big [f(c_j^+)-f(c_j^-)\big ] +\textstyle {\frac{1}{2}}\mu ([DJ_f>t]). \end{aligned} \end{aligned}$$

Letting \(m\rightarrow \infty \) and taking into account (2.1c) implies \(\mu ([DJ_f>t])=0\) for all \(t>0\). For a different approach use Theorem 4.1.

1.1 2.1c A Continuous, Nowhere Differentiable Function ([167])

For a real number x, denote by \(\{x\}\) the distance from x to its nearest integer and set

$$\begin{aligned} f(x)=\mathop {\textstyle {\sum }}\limits _{n=0}^\infty \>\frac{\{10^n x\}}{10^n}. \end{aligned}$$
(2.2c)

Each term of the series is continuous. Moreover the series is uniformly convergent being majorized by the geometric series \({\sum } 10^{-n}\). Therefore f is continuous. Since \(f(x)=f(x+j)\) for every integer j and all \(x\in \mathbb {R}\), it suffices to consider \(x\in [0,1)\). Any such x has a decimal expansion of the form \(x=0.a_1a_2\dots a_n\dots \), where \(a_i\) are integers from 0 to 9. By excluding the case when \(a_i=9\) for all i larger than some m, such a representation is unique.

For \(n\in \mathbb {N}\) fixed compute

$$\begin{aligned} \begin{array}{ll} \{10^n\,x\}=0.a_{n+1}a_{n+2}\dots &{}\text { if }\> 0.a_{n+1}a_{n+2}\dots \le \frac{1}{2}\\ \{10^n\,x\}=1- 0.a_{n+1}a_{n+2}\dots &{}\text { if }\> 0.a_{n+1}a_{n+2}\dots >\frac{1}{2}. \end{array} \end{aligned}$$

Having fixed \(x\in [0,1)\) choose increments

$$\begin{aligned} h_m=\left\{ \begin{array}{ll} -10^{-m}\qquad &{}\text { if either } a_m=4 \text { or } a_m=9\\ +10^{-m}\qquad &{}\text { otherwise}. \end{array} \right. \end{aligned}$$

Then form the difference quotients of f at x

$$\begin{aligned} \frac{f(x+h_m)-f(x)}{h_m}=10^m\mathop {\textstyle {\sum }}\limits _{n=0}^\infty \,\pm \frac{\{10^n(x\pm 10^{-m})\}-\{10^n\,x\}}{10^n}. \end{aligned}$$

The numerators of the terms of this last series, all vanish for \(n\ge m\), whereas for \(n=0,1,\dots ,(m-1)\) they are equal to \(\pm 10^{n-m}\). Therefore the difference quotient reduces to the sum of m terms each of the form \(\pm 1\). Such a sum is an integer, positive or negative, which has the same parity of m. Thus the limit as \(h_m\rightarrow 0\) of the difference ratios does not exists.

Remark 2.1c

The function in (2.2c) is not of bounded variation in any interval \([a,b]\subset \mathbb {R}\).

1.2 2.2c An Application of the Baire Category Theorem

The existence of a continuous and nowhere differentiable function, can be established indirectly, by a category-type argument. The Dini derivatives \(D_+f\) and \(D^+f\) introduced in (2.1), are called the right Dini numbers.

Proposition 2.1c

(Banach [12]) There exists a real-valued, continuous function in [0, 1] such that its Dini’s numbers \(|D_+f|\) and \(|D^+f|\) are infinity at every point of (0, 1).

Proof

For \(n\in \mathbb {N}\) let \(E_n\) denote the collection of all functions \(f\in C[0,1]\), for which there exists at least one point \(t\in [0,1-\frac{1}{n}]\) for which

$$\begin{aligned} \Big |\frac{f(t+h)-f(t)}{h}\Big |\le n \quad \text { for all }\>h\in (0,1-t). \end{aligned}$$

Each \(E_n\) is closed and nowhere dense in C[0, 1]. Both statements are meant with respect to the topology of the uniform convergence in C[0, 1]. To prove that \(E_n\) is nowhere dense in C[0, 1] observe that any continuous function in [0, 1] can be approximated in the sup-norm by continuous functions with polygonal graph of arbitrarily large Lipschitz constant. Then the complement \(C[0,1]-\cup E_n\) is non-empty.\(\blacksquare \)

4c Differentiating Series of Monotone Functions

4.1. :

Let \(\{f_n\}\) be a sequence of functions of bounded variation in [ab] such that the series \(\mathop {\textstyle {\sum }}\limits f_n(x)\) and \(\mathop {\textstyle {\sum }}\limits \mathcal {V}_{f_n}[a,x]\) are both convergent in [ab]. Then the sum f of the first series is of bounded variation in [ab] and the derivative can be computed term by term, a.e. in [ab].

 

5c Absolutely Continuous Functions

5.1. :

Let f be absolutely continuous in [ab]. Then f is Lipschitz continuous in [ab] if and only \(f^\prime \) is a.e. bounded in [ab].

5.2. :

The function

$$\begin{aligned} f(x)=\left\{ \begin{array}{ll} {\displaystyle x^{1+\varepsilon }\sin \frac{1}{x}}\quad &{}\text { for }\>x\in (0,1]\\ 0&{}\text { for }\>x=0 \end{array} \right. \end{aligned}$$

is absolutely continuous in [0, 1] for all \(\varepsilon >0\).

1.1 5.1c The Cantor Ternary Function ([23])

Set \(f(0)=0\) and \(f(1)=1\). Divide the interval [0, 1] into 3 equal subintervals and on the central interval \([\frac{1}{3},\frac{2}{3}]\) set \(f=\frac{1}{2}\), i.e., f is defined to be the average of its values at the extremes of the parent interval [0, 1]. Next divide the interval \([0,\frac{1}{3}]\) into 3 equal subintervals, and on the central interval \([\frac{1}{3^2},\frac{2}{3^2}]\) set \(f=\frac{1}{4}\), i.e., f is defined to be the average of its values at the extremes of the parent interval \([0,\frac{1}{3}]\). Likewise divide the interval \([\frac{2}{3},1]\) into 3 equal subintervals, and on the central interval \([\frac{7}{3^2},\frac{8}{3^2}]\) set \(f=\frac{3}{4}\), i.e., f is defined to be the average of its values at the extremes of the parent interval \([\frac{2}{3},1]\).

Proceeding in this fashion we define f in the whole [0, 1], by successive averages. By construction f is non-constant, non-decreasing, and continuous in [0, 1]. Since it is constant on each of the intervals making up the complement of the Cantor set \(\mathcal {C}\), its derivative vanishes in [0, 1] except on \(\mathcal {C}\). Thus \(f^\prime =0\) a.e. on [0, 1].

1.1.1 5.1.1.c Another Construction of the Cantor Ternary Function

The same function can be defined by an alternate procedure that uses the ternary expansion of the elements of the Cantor set. For \(x\in \mathcal {C}\) let \(\{\epsilon _j\}\) be sequence, with entries only 0 or 1, corresponding to the ternary expansion of x, as in (2.1) of Chap. 1. Then define

$$\begin{aligned} f(x)=f\Big (\mathop {\textstyle {\sum }}\limits _{j=1}^\infty \, \frac{2}{3^j}\epsilon _i{x,j}\Big )\buildrel \mathrm { def}\over {=} \mathop {\textstyle {\sum }}\limits _{j=1}^\infty \,\frac{1}{2^j}\epsilon _{x,j}. \end{aligned}$$
(5.1c)

Let \((\alpha _n,\beta _n)\) be an interval removed in the \(n^{\text {th}}\) step of the construction of the Cantor set. The extremes \(\alpha _n\) and \(\beta _n\) belong to \(\mathcal {C}\) and their ternary expansion is described in 2.2 of the Complements of Chap. 1. From the form of such expansion compute

$$\begin{aligned} f(\alpha _n)-f(\beta _n)=\frac{1}{2^n}-\mathop {\textstyle {\sum }}\limits _{j=n+1}^\infty \,\frac{1}{2^j}=0. \end{aligned}$$

If \((\alpha _n,\beta _n)\) is an interval in \([0,1]-\mathcal {C}\) we set

$$\begin{aligned} f(x)=f(\alpha _n)\qquad \text { for all } \> x\in [\alpha _n,\beta _n]. \end{aligned}$$

In such a way f is defined in the whole [0, 1] and \(f^\prime =0\) a.e. in [0, 1]. The right-hand side of (5.1c) is the decimal representation of the number in [0, 1] whose binary representation is the sequence \(\epsilon _x\). As x ranges over \(\mathcal {C}\), the sequences \(\epsilon _x\), with only entries 0 or 1, range over all such sequences. Therefore f maps \(\mathcal {C}\) onto [0, 1]. To show that f is continuous, observe that f is monotone and finite. Therefore its possible discontinuity points are discrete jumps. If f were not continuous then it would not be a surjection over [0, 1].

The Cantor ternary function is continuous, of bounded variation, but not absolutely continuous. This can be established indirectly by means of Corollary 5.2. Give a direct proof.

1.2 5.2c A Continuous Strictly Monotone Function with a.e. Zero Derivative

The Cantor ternary function is piecewise constant on the complement of the Cantor set. This accounts for \(f^\prime =0\) a.e. in [0, 1]. We next exhibit a continuous strictly increasing function in [0, 1], whose derivative vanishes a.e. in [0, 1].

Let \(t\in (0,1)\) be fixed and define \(f_o(x)=x\) and

$$\begin{aligned} f_1(x)=\left\{ \begin{array}{ll} (1+t)x\quad &{}\text { for }\> 0\le x\le \frac{1}{2}\\ (1-t)x+t\quad &{}\text { for }\> \frac{1}{2}\le x\le 1. \end{array} \right. \end{aligned}$$

The function \(f_1\) is constructed by dividing [0, 1] into two equal subintervals, by setting \(f_1=f_o\) at the end points of [0, 1], by setting

$$\begin{aligned} f_1\Big (\frac{1}{2}\Big )=\frac{1-t}{2}f_o(0)+ \frac{1+t}{2}f_o(1)=\frac{1+t}{2} \end{aligned}$$

and by defining \(f_1\) to be affine in the intervals \([0,\frac{1}{2}]\) and \([\frac{1}{2},1]\).

This procedure permits one to construct an increasing sequence \(\{f_n\}\) of strictly increasing functions in [0, 1]. Precisely if \(f_n\) has been defined, it must be affine in each of the subintervals

$$\begin{aligned} \Big [\frac{j}{2^n}\,,\,\frac{j+1}{2^n}\Big ] \qquad j=0,1,\dots ,2^n-1. \end{aligned}$$

Subdivide each of these into two equal subintervals, and define \(f_{n+1}\) to be affine on each of these with values at the end points, given by

$$\begin{aligned} f_{n+1}\Big (\frac{j}{2^n}\Big )&=f_n\Big (\frac{j}{2^n}\Big )\\ f_{n+1}\Big (\frac{j+1}{2^n}\Big )&=f_n\Big (\frac{j+1}{2^n}\Big )\\ f_{n+1}\Big (\frac{2j+1}{2^{n+1}}\Big )&= \frac{1-t}{2} f_n\Big (\frac{j}{2^n}\Big ) +\frac{1+t}{2}f_n\Big (\frac{j+1}{2^n}\Big ). \end{aligned}$$

By construction \(\{f_n\}\) is increasing and

$$\begin{aligned} f_m\Big (\frac{j}{2^n}\Big )=f_n\Big (\frac{j}{2^n}\Big ) \quad \text { for all }\> m\ge n,\quad j=0,1,\dots ,2^n-1. \end{aligned}$$
(5.2c)

The limit function f is non-decreasing. We show next that it is continuous and strictly increasing in [0, 1]. Every fixed \(x\in [0,1]\) is included into a sequence of nested and shrinking intervals \([\alpha _n,\beta _n]\) of the type

$$\begin{aligned} \alpha _n=\frac{m_{n,x}}{2^n}\quad \beta _n= \frac{m_{n,x}+1}{2^n}\quad \text { for some }\> m_{n,x}\in \mathbb {N}\cup \{0\}. \end{aligned}$$

By the construction of \(f_{n+1}\), if the parent interval of \([\alpha _n,\beta _n]\) is \([\alpha _n,\beta _{n-1}]\)

$$\begin{aligned} f_{n+1}(\beta _n)-f_{n+1}(\alpha _n)=\frac{1+t}{2} [f_n(\beta _{n-1})-f_n(\alpha _n)]. \end{aligned}$$

Likewise if the parent interval of \([\alpha _n,\beta _n]\) is \([\alpha _{n-1},\beta _n]\)

$$\begin{aligned} f_{n+1}(\beta _n)-f_{n+1}(\alpha _n)=\frac{1-t}{2} [f_n(\beta _n)-f_n(\alpha _{n-1})]. \end{aligned}$$

Therefore by (5.2c) either

$$\begin{aligned} f_{n+1}(\beta _n)-f_{n+1}(\alpha _n)=\frac{1+t}{2} [f_n(\beta _{n-1})-f_n(\alpha _{n-1})] \qquad {(5.2\mathrm{c})_+} \end{aligned}$$

or

$$\begin{aligned} f_{n+1}(\beta _n)-f_{n+1}(\alpha _n)=\frac{1-t}{2} [f_n(\beta _{n-1})-f_n(\alpha _{n-1})].\qquad {(5.2\mathrm{c})_-} \end{aligned}$$

From this by iteration

$$\begin{aligned} f_{n+1}(\beta _n)-f_{n+1}(\alpha _n) =\mathop {\textstyle {\prod }}\limits _{i=1}^n\,\frac{1+\varepsilon _i t}{2} \quad \text { where }\quad \varepsilon _i=\pm 1. \end{aligned}$$

Since for all \(m\ge n+1\)

$$\begin{aligned} f_m(\beta _n)=f_{n+1}(\beta _n)\quad \text { and }\quad f_m(\alpha _n)=f_{n+1}(\alpha _n) \end{aligned}$$

the previous equality implies

$$\begin{aligned} f(\beta _n)-f(\alpha _n)=\mathop {\textstyle {\prod }}\limits _{i=1}^n\,\frac{1+\varepsilon _i t}{2} \quad \text { where }\quad \varepsilon _i=\pm 1. \end{aligned}$$

For each fixed n the right-hand side is strictly positive. Thus \(f(\beta _n)>f(\alpha _n)\), i.e., f is strictly monotone. On the other hand (5.2c)\({}_{\pm }\) imply also

$$\begin{aligned} f(\beta _n)-f(\alpha _n)\le \Big (\frac{1+t}{2}\Big )^n\longrightarrow 0 \quad \text { as }\quad n\rightarrow \infty . \end{aligned}$$

Thus f is continuous in [0, 1]. Still from (5.2c)\({}_{\pm }\) we compute

$$\begin{aligned} \frac{f(\beta _n)-f(\alpha _n)}{\beta _n-\alpha _n}= \mathop {\textstyle {\prod }}\limits _{i=1}^n\,(1+\varepsilon _i t). \end{aligned}$$

As \(n\rightarrow \infty \) the right-hand side either converges to zero, or diverges to infinity or the limit does not exists. However, since f is monotone it is a.e. differentiable. Therefore the limit exists for a.e. \(x\in [0,1]\) and is zero. By Corollary 5.2 such a function is not absolutely continuous. Give a direct proof.

5.3. :

The function f constructed in § 14 of Chap. 3 is not absolutely continuous.

5.4. :

Let \(\mu \) be a Radon measure on \(\mathbb {R}\) defined on the same \(\sigma \)-algebra of the Lebesgue measurable sets in \(\mathbb {R}\), and absolutely continuous with respect to the Lebesgue measure on \(\mathbb {R}\). Then set

$$\begin{aligned} f(x)=\left\{ \begin{array}{ll} +\mu ([\alpha ,x])\quad &{}\text { for }\> x\in [\alpha ,\infty )\\ -\mu ([x,\alpha ])\quad &{}\text { for }\> x\in (-\infty ,\alpha ]. \end{array} \right. \end{aligned}$$

The function f is locally absolutely continuous, i.e., its restriction to any bounded interval is absolutely continuous. The function f can be used to generate the Lebesgue-Stieltjes measure \(\mu _f\). The measure \(\mu _f\) coincides with \(\mu \) on the Lebesgue measurable sets.

1.3 5.3c Absolute Continuity of the Distribution Function of a Measurable Function

For a measurable function f on a measure space \(\{X,\mathcal {A},\mu \} \), let \(\mu _f\) and \(f_*\) be respectively, the distribution measure and the distribution function of f, as defined (9.1) and (15.1c of the Complements of Chap. 4).

Proposition 5.1c

The distribution function \(f_*\) is absolutely continuous, in every closed subinterval of \(\mathbb {R}\), if and only if the distribution measure \(\mu _f\) is absolutely continuous with respect to the Lebesgue measure on \(\mathbb {R}\). In such a case

$$\begin{aligned} f_*(t)=\int _{-\infty }^t \varphi (s)ds \end{aligned}$$

where \(\varphi \) is the Radon-Nikodým derivative of \(\mu \) with respect to the Lebesgue measure ds.

5.5. :

Give an example of \(\{X,\mathcal {A},\mu \} \) and f for which \(f_*\) is continuous and not absolutely continuous.

7c Derivatives of Integrals

7.1. :

Construct a measurable set \(E\subset (-1,1)\) such that \(d_E^{\,\prime }(0)=\frac{1}{2}\). For \(n\in \mathbb {N}\) setFootnote 5

$$\begin{aligned} I_n=\Big [\frac{1}{2^n},\frac{1}{2^{n-1}}\Big ) \quad \text { so that }\quad (0,1)=\mathop {\textstyle {\bigcup }}\limits _{n\in \mathbb {N}} I_n. \end{aligned}$$

Divide each \(I_n\) into \(2^n\) equal subintervals, each of length \(2^{-2n}\) and retain only those \(\frac{1}{2}\)-closed intervals of even parity, to obtain sets

$$\begin{aligned} E_n=\mathop {\textstyle {\bigcup }}\limits _{j=0}^{2^{n-1}-1} \Big [\frac{2^n+2j}{2^{2n}}\,,\, \frac{2^n+2j+1}{2^{2n}}\Big ),\quad \text { and }\quad E^+=\mathop {\textstyle {\bigcup }}\limits E_n. \end{aligned}$$

Verify that \(\mu (E_n)=\frac{1}{2}\mu (I_n)\) and \(\mu (E^+)=\frac{1}{2}\). having fixed \(h\in (0,1)\) there exists \(n_h, i_h\in \mathbb {N}\), such that

$$\begin{aligned} h\in \Big [\frac{1}{2^{n_h}},\frac{1}{2^{n_h-1}}\Big ]\quad \text { and }\quad h\in \Big [\frac{2^{n_h}+i_h}{2^{2n_h}}\,,\, \frac{2^{n_h}+(i_h+1)}{2^{2n_h}}\Big ]. \end{aligned}$$

The first of these implies that \(h=O(2^{-{n_h}})\). Next compute

$$\begin{aligned} (0,h]\cap E^+=\Big (\mathop {\textstyle {\bigcup }}\limits _{\ell =n_h+1}^\infty E_{\ell }\Big )\mathop {\textstyle {\bigcup }}\limits \Big (\mathop {\textstyle {\bigcup }}\limits _{j=0}^{2^{n_h-1}-1}(0,h]\mathop {\textstyle {\bigcap }}\limits \Big (\frac{2^{n_h}+2j}{2^{2n_n}}\,,\, \frac{2^{n_h}+2j+1}{2^{2n_h}}\Big )\Big ). \end{aligned}$$

From this

$$\begin{aligned} \mu \big [(0,h]\cap E\big ]={\textstyle \frac{1}{2}}h+ O(2^{-{2n_h}})\quad \Longrightarrow \quad \mu \big [(0,h]\cap E^+\big ]={\textstyle \frac{1}{2}}h+ O(h^2). \end{aligned}$$

Thus

$$\begin{aligned} \lim _{h\rightarrow 0}\frac{1}{h} \int _0^h \chi _{E^+} dx=\frac{1}{2}. \end{aligned}$$

Let \(E^-\) be the symmetric of \(E^+\) in \((-1,0]\) and set \(E=E^-\cup E^+\).

7.2. :

Let f be absolutely continuous in [ab]. Then the function \(x\rightarrow \mathcal {V}_f[a,x]\) is also absolutely continuous in [ab]. Moreover

$$\begin{aligned} \mathcal {V}_f[a,x]=\int _a^x|f^\prime (t)|dt\quad \text { for all }\> x\in [a,b]. \end{aligned}$$
7.3. :

Let f be of bounded variation in [ab]. The singular part of f is the function ([90])

$$\begin{aligned} \sigma _f(x)=f(x)-f(a)-\int _a^xf^\prime (t)dt. \end{aligned}$$
(7.1c)

The singular part of f is of bounded variation and \(\sigma ^\prime =0\) a.e. in [ab]. It has the same singularities as f, and \((f-\sigma )\) is absolutely continuous.

Thus every function f of bounded variation in [ab], can be decomposed into the sum of an absolutely continuous function in [ab] and a singular function. Compare the \(\sigma _f\) with the functions of the jumps \(J_f\) given in § 1.1c of the Complements.

7.4. :

Let f be Lebesgue integrable in the interval [ab] and let F denote a primitive of f. Then for every absolutely continuous function g defined in [ab]

$$\begin{aligned} \int _a^b fgdx=F(b)g(b)-F(a)g(a)- \int _a^b Fg^\prime dx. \end{aligned}$$
7.5. :

Let \(f,g:[a,b]\rightarrow \mathbb {R}\) be absolutely continuous. Then

$$\begin{aligned} \int _a^b fg^\prime dx+\int _a^b f^\prime gdx= f(a)g(a)- f(b)g(b). \end{aligned}$$
7.6. :

Let \(h:[a,b]\rightarrow [c,d]\) be absolutely continuous, increasing and such that \(h(a)=c\) and \(h(b)=d\). Then for every nonnegative, Lebesgue measurable function \(f:[c,d]\rightarrow \mathbb {R}\), the composition f(h) is measurable and

$$\begin{aligned} \int _c^d f(s)ds=\int _a^b f\big (h(t)\big ) h^\prime (t)dt. \end{aligned}$$

This is established sequentially for f the characteristic function of an interval, the characteristic function of an open set, the characteristic function of a measurable set, for a simple function.

Proposition 7.1c

Let f be absolutely continuous in [ab]. Then for every measurable \(E\subset [a,b]\) of measure zero, \(f(E)\subset \mathbb {R}\) is a set of measure zero.

Proof

Combine Proposition 7.2 with and Vitali’s absolute continuity of the integral (Theorem 11.1 of Chap. 4).\(\blacksquare \)

The converse of Proposition 7.1c is in false. The characteristic function of the rationals maps any set into a set of measure zero. Such a function however is not continuous.

7.7. :

Prove by a counterexample that the converse of Proposition 7.1c is false, even if f is assumed to be continuous. Hint: The function f in (1.1) is continous in [0, 1] and not BV[0, 1]. To show that it maps measurable sets of measure zero, into sets of measure zero, observe that \(f\in AC[\varepsilon ,1]\) for all \(\varepsilon >0\).

Proposition 7.2c

Let \(f:[a,b]\rightarrow \mathbb {R}\) be continuous and monotone and such that for every measurable \(E\subset [a,b]\) of measure zero, \(f(E)\subset \mathbb {R}\) is a set of measure zero. Then f is absolutely continuous in [ab].

Proof

Assuming f non-decreasing the function \(h(x)=x+f(x)\) is strictly increasing. The set function \(\nu (E)=\mu (h(E))\) for all Lebesgue measurable sets in [ab] is a measure on the same \(\sigma \)-algebra, satisfying \(\nu \ll \mu \). Apply the Radon-Nykodým theorem.\(\blacksquare \)

1.1 7.1c Characterizing BV[ab] Functions

Denote by \(C_o^1[a,b]\) the collection of continuously differentiable functions \(\varphi \) of compact support in [ab].

Proposition 7.3c

Let \(f\in BV[a,b]\). Then

$$\begin{aligned} \sup _{\genfrac{}{}{0.0pt}{}{\varphi \in C_o^1(\mathbb {R})}{|\varphi |\le 1}} \int _a^b f\varphi ^\prime dx \le \mathcal {V}_f[a,b]. \end{aligned}$$
(7.2c)

Proof

One may assume that f is monotone increasing and nonnegative. There exists an increasing sequence of simple functions such that \(\{f_n\}\rightarrow f\) everywhere in [ab] (Proposition 3.1 of Chap. 4). By the monotonicity of f, the construction of \(\{f_n\}\) identifies a partition

$$\begin{aligned} P=\{a=x_o<x_1<\cdots <x_n=b\} \end{aligned}$$
(7.3c)

of [ab] such that

$$\begin{aligned} f_n(x) =f(x_{j-1}) \quad \text { in the interval }\> [x_{j-1},x_j]\quad \text { for } \> j=1,\dots ,n. \end{aligned}$$

For \(0<\delta \ll 1\) and n fixed construct the Lipschitz continuous functions

$$\begin{aligned} f_{n,\delta }(x)=\mathop {\textstyle {\sum }}\limits _{j=1}^n \Big \{f(x_{j-1})\chi _{[x_{j-1},x_j]} +[f(x_j)-f(x_{j-1})] \chi _{[x_j-\delta ,x_j]} \frac{x-x_j +\delta }{\delta }\Big \}. \end{aligned}$$

One verifies that

$$\begin{aligned} \int _a^b f\varphi ^\prime dx =\lim _{n\rightarrow \infty }\lim _{\delta \rightarrow 0} \int _a^b f_{n,\delta }\varphi ^\prime dx. \end{aligned}$$

Since \(f_{n,\delta }\) are absolutely continuous in [ab], integration by parts is justified. Hence for every \(\varphi \in C_o^1[a,b]\),

$$\begin{aligned} \Big |\int _a^b f_{n,\delta }\varphi ^\prime dx\Big |=\Big | \mathop {\textstyle {\sum }}\limits _{j=1}^n\int _{x_{j-1}}^{x_j}f_{n,\delta } \varphi ^\prime dx\Big |\le \mathcal {V}_f[a,b]. \end{aligned}$$

\(\blacksquare \)

The converse of (7.2c) is in general false. For example \(\chi _{\{\frac{1}{2}\}}\in BV[0,1]\) with variation 2, whereas the left-hand side of (7.2c) is zero. The latter only requires that f be integrable, whereas the notion of bounded variation requires that f be defined at every point of [ab]. However if f is integrable in [ab] then it is unambigously defined by (11.1), or (11.1)\({}_{N=1}^\prime \), everywhere in [ab] except for a set of measure zero. Given f integrable in [ab] introduce partitions \(P_f\) as in (7.3c), where however \(x_j\) are differentiability points of f. Then define the essential variation of f in (ab) as

$$\begin{aligned} {\text {ess}}-\mathcal {V}_f(a,b)=\sup _{P_f}\mathop {\textstyle {\sum }}\limits _{j=1}^n|f(x_{j})-f(x_{j-1})|. \end{aligned}$$

Proposition 7.4c

Let f be integrable in [ab]. Then

$$\begin{aligned} {\text {ess}}-\mathcal {V}_f(a,b)\le \sup _{\genfrac{}{}{0.0pt}{}{\varphi \in C_o^1[a,b]}{|\varphi |\le 1}} \int _a^b f\varphi ^\prime dx. \end{aligned}$$
(7.4c)

Proof

Assume the right-hand side of (7.4c) is finite, and fix a partition \(P_f\). Without loss of generality may assume that a and b are points of \(P_f\), and construct the polygonal of vertices \(\big (x_j, f(x_j)\big )\). Assume momentarily that such a polygonal changes its monotonicity at each of its veritice, i.e.,

$$\begin{aligned} \begin{aligned}&\text {if }\> f(x_{j-1})< f(x_j)\>\text { then }\> f(x_j)> f(x_{j+1});\\&\text {if }\> f(x_{j-1})> f(x_j)\>\text { then }\> f(x_j)< f(x_{j+1}). \end{aligned} \end{aligned}$$
(7.5c)

Assume \(f(a)<f(x_1)\) and set

$$\begin{aligned} \varphi _\delta (x)=\left\{ \begin{array}{ll} {\displaystyle \frac{a-x}{\delta }} \>&{}\text { for }\> a\le x<a+\delta ,\\ -1\>&{}\text { for }\> a+\delta \le x\le x_1-\delta ,\\ {\displaystyle \frac{x-x_1+\delta }{\delta }-1}\>&{}\text { for }\> x_1-\delta \le x<x_1.\\ \end{array}\right. \end{aligned}$$

Then, for \(j=1,\dots ,n\), define \(\varphi \) recursively as

$$\begin{aligned} \begin{array}{ll} {\varphi _\delta (x)=\left\{ \begin{array}{ll} {\displaystyle \frac{x_{j-1}-x}{\delta }}\>&{}\text { for }\> x_{j-1}\le x<x_{j-1}+\delta ,\\ -1\>&{}\text { for }\> x_{j-1}+\delta \le x\le x_j-\delta ,\\ {\displaystyle \frac{x-x_j+\delta }{\delta }-1}\>&{}\text { for }\> x_j-\delta \le x<x_j, \end{array}\right. } \quad &{}\text { if }\> f(x_{j-1})<f(x_j);\\ {}\\ {}\\ {\varphi (x)_\delta =\left\{ \begin{array}{ll} {\displaystyle \frac{x-x_{j-1}}{\delta }}\>&{}\text { for }\> x_{j-1}\le x<x_{j-1}+\delta ,\\ 1\>&{}\text { for }\> x_{j-1}+\delta \le x\le x_j-\delta ,\\ {\displaystyle \frac{x_j-\delta -x}{\delta }+1}\>&{}\text { for }\> x_j-\delta \le x<x_j, \end{array}\right. } \quad&\text { if }\> f(x_{j-1})>f(x_j). \end{array} \end{aligned}$$

Assuming momentarily that such a choice is admissible, compute

$$\begin{aligned} \sup _{\genfrac{}{}{0.0pt}{}{\varphi \in C_o^1[a,b]}{|\varphi |\le 1}}\int _a^b f\varphi ^\prime dx&\ge \limsup _{\delta \rightarrow 0}\Big |\int _a^b f\varphi _\delta ^\prime dx\Big |\\&=\limsup _{\delta \rightarrow 0}\Big |\frac{\text { sign }{\varphi _\delta ^\prime }}{\delta } \int _a^{a+\delta } fdx\\&\qquad \qquad + \mathop {\textstyle {\sum }}\limits _{j=1}^{n-1} \frac{\text { sign }{\varphi _\delta ^\prime }}{\delta }\int _{x_j-\delta }^{x_j+\delta } fdx +\frac{\text { sign }{\varphi _\delta ^\prime }}{\delta }\int _{b-\delta }^b fdx\Big |\\&\ge \mathop {\textstyle {\sum }}\limits _{j=1}^n |f(x_j)-f(x_{j-1})|, \end{aligned}$$

since \(x_j\) are differentiability points of f.\(\blacksquare \)

Complete the proof by the following steps:

i. :

Prove that in (7.4c) the supremum can be taken over all Lipschitz continuous functions \(\varphi \in C_o[a,b]\).

ii. :

Remove the assumptions and that \(P_f\) satisfies (7.5c), to establish the inequality for all partitions \(P_f\) of [ab]. \(\blacksquare \)

Corollary 7.1c

An integrable function f in [ab] is of essentially bounded variation in [ab] if and only if the right-hand side of (7.4c) is finite.

1.2 7.2c Functions of Bounded Variation in N Dimensions [55]

The characterization of Corollary 7.1c suggests a notion bounded variations in several dimensions. A function f locally integrable in \(\mathbb {R}^N\) is of bounded variation in a Lebesgue measurable set \(E\subset \mathbb {R}^N\) if there exists a constant C such that

$$\begin{aligned} \Big |\int _Ef{\text {div}}{\varvec{\varphi }}dx\Big |\le C \end{aligned}$$
(7.6c)

for all vector valued functions

$$\begin{aligned} {\varvec{\varphi }}=(\varphi _1,\dots ,\varphi _N)\in [C_o^1(\mathbb {R}^N)]^N \quad \text { such that }\> |{\varvec{\varphi }}|\le 1. \end{aligned}$$

The smallest constant C for which (7.6c) holds is the variation of f in E and is denoted by \(\Vert Df\Vert (E)\).

1.2.1 7.2.1c Perimeter of a Set

Let E be a bounded measurable set in \(\mathbb {R}^N\) with smooth boundary \({\partial }{E}\). Prove that \(\chi _E\in BV(\mathbb {R}^N)\) and that \(\mathcal {V}_{\chi _E}(\mathbb {R}^N)=\mathcal {H}^{N-1}({\partial }{E})\), where \(\mathcal {H}^{N-1}\) is the \((N-1)\)-dimensional Hausdorff measure of sets in \(\mathbb {R}^N\).

If \({\partial }{E}\) is not smooth, (7.6c) suggests defining the “measure of \({\partial }{E}\)” or, roughly speaking, the perimeter of E, by

$$\begin{aligned} {\text {Per}}(E)=\sup _{{\varvec{\varphi }}\in [C_o^1]^N,\,|{\varvec{\varphi }}|\le 1} \int _{\mathbb {R}^N} \chi _E{\text {div}}{\varvec{\varphi }}dx, \end{aligned}$$
(7.7c)

provided the right-hand side is finite. Sets \(E\subset \mathbb {R}^N\) for which \({\text {Per}}(E)<\infty \) are called of finite perimeter. They correspond, roughly speaking, to sets for which the Gauss-Green theorem holds.

13c Convex Functions

13.1. :

Give an example of a bounded, discontinuous, convex function in [ab]. Give an example of a convex function unbounded in (ab).

13.2. :

A continuous function f in (ab) is convex if and only if

$$\begin{aligned} f\Big (\frac{x+y}{2}\Big )\le \frac{f(x)+f(y)}{2}\qquad \text { for all }\> x,y\in (a,b). \end{aligned}$$
13.3. :

Let f be convex, non-decreasing and non-constant in \((0,\infty )\). Then \(f(x)\rightarrow \infty \) as \(x\rightarrow \infty \).

13.4. :

Let f be convex in \([0,\infty )\). Then the limit of \(x^{-1}f(x)\) as \(x\rightarrow \infty \), exists finite or infinite.

13.5. :

Let \(\{f_n\}\) be a sequence of convex functions in (ab) converging to some real-valued function f. Then the convergence is uniform within any closed subinterval of (ab). The conclusion is false if f is permitted to take values in \(\mathbb {R}^*\), as shown by the sequence \(\{x^n\}\) for \(x\in (0,2)\).

13.6. :

Let \(f\in C^2(a,b)\). Then f is convex in (ab) if and only if \(f^{\prime \prime }(x)\ge 0\) for each \(x\in (a,b)\). Proof: Having fixed \(x<y\) it suffices to prove that

$$\begin{aligned}{}[0,1]\ni t\rightarrow \varphi (t) = f(tx+(1-t)y)-t f(x)-(1-t) f(y) \end{aligned}$$
(13.1c)

is nonpositive in [0, 1]. Such a function vanishes at the end points of [0, 1] and its extrema are minima since

$$\begin{aligned} \varphi ^{\prime \prime }(t)=(x-y)f^{\prime \prime }(tx+(1-t)y)\le 0. \end{aligned}$$

Proposition 13.1c

A continuous function f in (ab) is convex if and only if either one of the two one-sided derivatives \(D_\pm f\) is non-decreasing.

Proof

Assume for example that \(D_+f\) is non-decreasing. If the function \(\varphi \) in (13.1c) has a positive maximum \(\varphi (t_o)>0\), at some \(t_o\in (0,1)\), then

$$\begin{aligned} D_+\varphi (t_o)=(x-y)D_+f(t_ox+(1-t_o)y)+f(y)-f(x)\le 0. \end{aligned}$$

Therefore, since \(D_+f\) is non-decreasing, \(D_+\varphi (t)\) is non-positive in \([0,t_o]\). Thus \(\varphi \) is non-increasing in \([0,t_o]\) and \(\varphi (t_o)\le 0\).\(\blacksquare \)

13.7. :

The function \(f(x)=|x|^p\) is convex for \(p\ge 1\) and concave for \(p\in (0,1)\).

1.1 13.8c Convex Functions in \(\mathbb {R}^N\)

Let E be a convex subset of \(\mathbb {R}^N\). A function \(f:E\rightarrow \mathbb {R}^*\) is convex if for every pair of points x and y in E and every \(t\in [0,1]\)

$$\begin{aligned} f(t x+(1-t)y)\le tf(x)+(1-t)f(y). \end{aligned}$$

The \((N+1)\)-dimensional set

$$\begin{aligned} \mathcal {G}_f=\left\{ (x,x_{N+1})\in \mathbb {R}^{N+1}\bigm | x\in E,\quad x_{N+1}\ge f(x)\right\} \end{aligned}$$

is the epigraph of f. The function f is convex if and only if its epigraph is convex.

13.9. :

Let \(E\subset \mathbb {R}^N\) be open and convex. A function \(f\in C^2(E)\) is convex if and only if \({\mathop {\textstyle {\sum }}\limits }_{i,j=1}^N f_{x_ix_j}\xi _i\xi _j\ge 0\) for all \(\xi \in \mathbb {R}^N\).

Hint: Fix \(B_\rho (x)\subset E\) and \(\xi \) in the unit sphere of \(\mathbb {R}^N\). The function \((-\rho ,\rho )\ni t\rightarrow \varphi (t)=f(x+t\xi )\) is convex.

13.10. :

Construct a non convex function \(f\in C^2(\mathbb {R}^2)\) such that \(f_{xx}\) and \(f_{yy}\) are both nonnegative.

13.11. :

Let \(E\subset \mathbb {R}^N\) be open and convex, and let f be convex and real-valued in E. Then f is continuous on E. Moreover for every \(x\in E\) there exist the left and right directional derivatives

$$\begin{aligned} D^\pm _{\mathbf {u}}f(x)= D^\pm _t f(x+t\mathbf {u})\bigm |_{t=0} \qquad \text { for all }\>|\mathbf {u}|=1. \end{aligned}$$

Moreover \(D^-_{\mathbf {u}}f\le D^+_{\mathbf {u}}f\). In particular for each \(x\in E\), there exist the left and right derivatives \(D^\pm _{x_j}f\), along the coordinate axes and \(D^-_{x_j}f\le D^+_{x_j}f\).

13.12. :

Let \(E\subset \mathbb {R}^N\) be convex. A function f defined in E is convex if and only if \(f(x)=\sup \pi (x)\), where \(\pi \le f\) is affine.

13.13. :

Let \(f:\mathbb {R}^N\rightarrow \mathbb {R}\) be convex. There exist a positive number k such that \(\liminf _{|x|\rightarrow \infty }|x|^{-1}f(x)\ge -k\).

1.2 13.14c The Legendre Transform ([92])

The Legendre transform \(f^*\) of a convex function \(f:\mathbb {R}^N\rightarrow \mathbb {R}^* \) is defined by

$$\begin{aligned} f^*(x)=\sup _{y\in \mathbb {R}^N}\{x\cdot y-f(y)\}. \end{aligned}$$
(13.2c)

Proposition 13.2c

\(f^*\) is convex in \(\mathbb {R}^N\) and \(f^{**}=f\).

Proof

The convexity of f follows from 13.12. From the definition (13.2c), \(f(y)+f^*(x)\ge y\cdot x\), for all \(x,y\in \mathbb {R}^N\). Therefore

$$\begin{aligned} f(y)\ge \sup _{x\in \mathbb {R}^N}\{y\cdot x-f^*(x)\}= f^{**}(y). \end{aligned}$$

Also, still from (13.2c)

$$\begin{aligned} f^{**}(x)&=\sup _{y\in \mathbb {R}^N}\big \{x\cdot y- \sup _{z\in \mathbb {R}^N}\{y\cdot z-f(z)\}\big \}\\&=\sup _{y\in \mathbb {R}^N}\inf _{z\in \mathbb {R}^N}\{y\cdot (x-z)+f(z)\}. \end{aligned}$$

Since f is convex, for a fixed \(x\in \mathbb {R}^N\), there exists a vector \(\mathbf {m}\) such that

$$\begin{aligned} f(z)-f(x)\ge \mathbf {m}\cdot (z-x)\qquad \text { for all }\> z\in \mathbb {R}^N. \end{aligned}$$

Combining these inequalities yields

$$\begin{aligned} f^{**}(x)\ge f(x)+\sup _{y\in \mathbb {R}^N}\inf _{z\in \mathbb {R}^N} (z-x)\cdot (\mathbf {m}-y)=f(x). \end{aligned}$$

\(\blacksquare \)

1.3 13.15c Finiteness and Coercivity

The Legendre transform \(f^*\), as defined by (13.2c), could be infinite even if f is finite in \(\mathbb {R}^N\). For example in \(\mathbb {R}\)

$$\begin{aligned} |x|^*=\left\{ \begin{array}{ll} 0\quad &{}\text { if }\> |x|\le 1\\ \infty \quad &{}\text { if }\> |x|>1. \end{array} \right. \end{aligned}$$

A convex function \(f:\mathbb {R}^N\rightarrow \mathbb {R}\) is coercive at infinity if

$$\begin{aligned} \lim _{|x|\rightarrow \infty }\frac{f(x)}{|x|}=\infty . \end{aligned}$$

Proposition 13.3c

If f is coercive at infinity, then \(f^*\) is finite in \(\mathbb {R}^N\). If f is finite, then \(f^*\) is coercive at infinity.

Proof

Assume f is coercive at infinity. If the \(\sup \) in (13.2c) is achieved for \(y=0\) the assertion is obvious. Otherwise

$$\begin{aligned} f^*(x)=\sup _{y\in \mathbb {R}^N-\{0\}}|y| \Big \{x\cdot \frac{y}{|y|}-\frac{f(y)}{|y|}\Big \}. \end{aligned}$$

Therefore the supremum is achieved for some finite y and \(f^*(x)\) is finite.

To prove the converse statement, fix \(\lambda >0\) and write

$$\begin{aligned} f^*(x)&=\sup _{y\in \mathbb {R}^N}\{x\cdot y-f(y)\}\ge \{x\cdot y-f(y)\}\bigm |_{y=\lambda x/|x|}\\&=\lambda |x|-f\big (\lambda \frac{x}{|x|}\big )\ge \lambda |x|-\sup _{|\mathbf {u}|=\lambda }|f(\mathbf {u})|. \end{aligned}$$

Therefore, since \(x\in \mathbb {R}^N-\{0\}\) is arbitrary

$$\begin{aligned} \lim _{|x|\rightarrow \infty }\frac{f^*(x)}{|x|}\ge \lambda \qquad \text { for all }\>\lambda >0. \end{aligned}$$

\(\blacksquare \)

1.4 13.16c The Young’s Inequality

Prove that the Legendre transform of the convex function

$$\begin{aligned} \mathbb {R}\ni a\rightarrow f(a)=\frac{1}{p} |a|^p\quad \text { for }\> 1< p<\infty \end{aligned}$$

is

$$\begin{aligned} \mathbb {R}\ni b\rightarrow f^*(b)=\frac{1}{q} |b|^q\quad \text { for }\> 1<q<\infty \quad \text { and }\>\frac{1}{p}+\frac{1}{q}=1. \end{aligned}$$

Then the definition (13.2c) of the Legendre transform implies the Young’s inequality

$$\begin{aligned} |ab|\le \frac{1}{p}|q|^p+\frac{1}{q}|b|^q\quad \text { for all }\> a,b\in \mathbb {R}. \end{aligned}$$
(13.3c)

The inequality continues to holds for the limiting case \(p=1\) and \(q=\infty \). For a different proof see Proposition 2.1 of Chap. 6.

14c Jensen’s Inequality

Proposition 14.1c

(Hölder [75]) Let \(\{\alpha _i\}\) be a sequence of nonnegative numbers such that \(\sum \alpha _i=1\), and let \(\{\xi _i\}\) be a sequence in \(\mathbb {R}\). Then

$$\begin{aligned} \exp \left( \mathop {\textstyle {\sum }}\limits \,\alpha _i\xi _i\right) \le \mathop {\textstyle {\sum }}\limits \, \alpha _i\exp (\xi _i). \end{aligned}$$
(14.1c)

Proof

Apply (14.1) to \(e^x\) with \(\eta =\xi _j\) and \(\alpha =\mathop {\textstyle {\sum }}\limits \alpha _i\xi _i\), to get

$$\begin{aligned} \exp \left( \mathop {\textstyle {\sum }}\limits \,\alpha _i\xi _i\right) + m_j\left( \xi _j- \mathop {\textstyle {\sum }}\limits \,\alpha _i\xi _i\right) \le \exp (\xi _j). \end{aligned}$$

Multiply by \(\alpha _j\) and add over j.\(\blacksquare \)

Corollary 14.1c

Let \(\{\alpha _i\}\) be a sequence of nonnegative numbers such that \(\sum \alpha _i=1\), and let \(\{\xi _i\}\) be a sequence of positive numbers. Then

$$\begin{aligned} \mathop {\textstyle {\prod }}\limits \,\xi _i^{\alpha _i}\le \mathop {\textstyle {\sum }}\limits \,\alpha _i\xi _i. \end{aligned}$$
(14.2c)

1.1 14.1c The Inequality of the Geometric and Arithmetic Mean

In the case where \(\alpha _i=0\) for \(i>n\) and \(\alpha _i=1/n\) for \(i=1,2,\dots ,n\), inequality (14.2c) reduces to the inequality between the geometric and arithmetic mean of n positive numbersFootnote 6

$$\begin{aligned} (\xi _1\xi _2\cdots \xi _n)^{1/n}\le \frac{\xi _1+\xi _2+\cdots +\xi _n}{n}. \end{aligned}$$
(14.3c)

1.2 14.2c Integrals and Their Reciprocals

Proposition 14.2c

Let E be a measurable set of finite measure and let \(f:E\rightarrow \mathbb {R}^+\) be measurable. Then

$$\begin{aligned} \frac{1}{\displaystyle \left( \frac{1}{\mu (E)}\int _Ef d\mu \right) ^p}\le \frac{1}{\mu (E)} \int _E\frac{1}{f^p} d\mu ,\quad \text { for all }\> p>0. \end{aligned}$$
(14.4c)

Proof

Assume first that f is integrable and that \(f\ge \varepsilon \), and apply Jensen’s inequality with \(\varphi (t)=t^{-p}\). \(\blacksquare \)

15c Extending Continuous Functions

15.1. :

Let f be convex in a closed interval [ab] and assume that \(D_+f(a)\) and \(D_-f(b)\) are both finite. There exists a convex function \(\widetilde{f}\) defined in \(\mathbb {R}\) such that \(f=\widetilde{f}\) in [ab].

16c The Weierstrass Approximation Theorem

16.1. :

Let \(E\subset \mathbb {R}^N\) be open, bounded and with smooth boundary \({\partial }{E}\). Let \(f\in C^1({\bar{E}})\) vanish on \({\partial }{E}\), and let \(P_j\) denote the jth Stieltjes polynomial relative to f . Then

$$\begin{aligned} \lim _{j\rightarrow \infty } \frac{\partial P_j}{\partial x_i}=\frac{\partial f}{\partial x_i} \qquad j=1,\dots ,N\quad \text { in }\> E. \end{aligned}$$
16.2. :

Let \(E\subset \mathbb {R}^N\) be bounded and open, and fet \(f:{\bar{E}}\rightarrow \mathbb {R}\) and be Lipschitz continuous in \({\bar{E}}\). with Lipschitz constant L. Then the Stieltjes polynomials \(P_j\) relative to f are equi-Lipschitz continuous in E, with the same constant L.

16.3. :

A continuous function \(f:[0,1]\rightarrow \mathbb {R}\), can be approximated by the Bernstein polynomials \(B_j\), relative to f

$$\begin{aligned} B_j(x)=\mathop {\textstyle {\sum }}\limits _{i=1}^j\,\Big ( \begin{array}{c} j\\ i \end{array} \Big ) f\Big (\frac{i}{j}\Big ) x^i(1-x)^{j-i}. \end{aligned}$$

State and prove a N-dimensional version of such an approximation ([98]).

16.4. :

Let f be uniformly continuous on a bounded, open set \(E\subset \mathbb {R}^N\) and denote by \(\mathcal {P}_n\) the set of all polynomials of degree n in the coordinate variables. Then

$$\begin{aligned} \int _Efp_ndx=0\>\text { for all } p_n\in \mathcal {P}_n \text { and all } n\in \mathbb {N}\>\Longrightarrow \> f=0. \end{aligned}$$

17c The Stone-Weierstrass Theorem

17.1. :

The Stone-Weierstrass theorem fails for complex valued functions.

Let D be the closed, unit disc in the complex plane \(\mathbb {C}\) and denote by \(C(D;\mathbb {C})\) the linear space of all the continuous complex valued functions defined in D endowed with the topology generated by the metric in (17.1).

Consider also the subset \(\mathcal {H}(D)\) of \(C(D;\mathbb {C})\), consisting of all holomorphic functions defined in D. One verifies that \(\mathcal {H}(D)\) is an algebra. Moreover uniform limits of holomorphic functions in D are holomorphic ([24], Chap. V, Théoréme 1, page 145). Thus \(\mathcal {H}(D)\) is closed under the metric in (17.1). The algebra \(\mathcal {H}(D)\) is called the disc algebra.

Such an algebra separates points since it contains the holomorphic function \(f(z)=z\). Moreover \(\mathcal {H}(D)\) contains the constants. However \(\mathcal {H}(D)\ne C(D;\mathbb {C})\). Indeed the function \(f(z)=\overline{z}\) is continuous but not holomorphic in D.

17.2. :

Let \(f:\mathbb {R}\rightarrow \mathbb {R}\) be continuous and \(2\pi \)-periodic. For every \(\varepsilon >0\) there exists a function of the type

$$\begin{aligned} \varphi (x)= a_o+\mathop {\textstyle {\sum }}\limits _{n=1}^m (b_n\cos {nx}+ c_n\sin {nx}) \end{aligned}$$

such that \(\sup _{\mathbb {R}}|f-\varphi |\le \varepsilon \) (Hint: Use Stone’s Theorem).

19c A General Version of the Ascoli-Arzelà Theorem

The proof of Theorem 19.1 uses only the separability of \(\mathbb {R}^N\) and the metric structure of \(\mathbb {R}\). Thus it can be extended into any abstract framework with these two properties.

Let \(\{f_n\}\) be a countable collection of continuous functions from a separable topological space \(\{X;\mathcal {U}\} \) into a metric space \(\{Y;d_Y\}\). The functions \(f_n\) are equibounded at x if the closure in \(\{Y;d_Y\}\) of the set \(\{f_n(x)\}\) is compact.

The functions \(f_n\) are equi-continuous at a point \(x\in X\) if for every \(\varepsilon >0\), there exists an open set \(\mathcal {O}\in \mathcal {U}\) containing x and such that

$$\begin{aligned} d_Y(f_n(x),f_n(y))\le \varepsilon \quad \text { for all }\> y\in \mathcal {O}\quad \text { and all }\> n\in \mathbb {N}. \end{aligned}$$

Theorem 19.1c

Let \(\{f_n\}\) be a sequence of continuous functions from a separable space \(\{X;\mathcal {U}\} \) into a metric space \(\{Y;d_Y\}\). Assume that the functions \(f_n\) are equibounded and equi-continuous at each \(x\in X\). Then, there exists a subsequence \(\{f_{n^\prime }\}\subset \{f_n\}\) and a continuous function \(f:X\rightarrow Y\) such that \(\{f_{n^\prime }\}\rightarrow f\) pointwise in X. Moreover the convergence is uniform on compact subsets of X.

State and prove an analog of Proposition 19.1.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer Science+Business Media New York

About this chapter

Cite this chapter

DiBenedetto, E. (2016). Topics on Measurable Functions of Real Variables. In: Real Analysis. Birkhäuser Advanced Texts Basler Lehrbücher. Birkhäuser, New York, NY. https://doi.org/10.1007/978-1-4939-4005-9_5

Download citation

Publish with us

Policies and ethics