Ì ÒØÓÖ ÙÒ Ø ÓÒ

ǺÓÚÓ× ݸ ǺŠÖØÓ¸ κÊÝ Þ ÒÓÚ Ò źÎÙÓÖÒ Ò

Ì× × Ò ØØÑÔØ ØÓ Ú  ×Ý×ØÑØ ×ÙÖÚÝ Ó ÔÖÓÔ ÖØ× Ó Ø ÑÓÙ× ÒØÓÖ

ØÖÒÖÝ ÙÒØÓÒº

ÃÝ ÛÓÖ× ËÒ ÙÐÖ ÙÒØÓÒ׸ ÒØÓÖ ÙÒØÓÒ¸ ÒØÓÖ ×غ

ÅØÑØ× ËÙ Ø Ð××ØÓÒ ´¾¼¼½µ ¾¹¼¾ ¾¿¼

 Ç Æ Ì  Æ Ì Ë

ÁÒØÖÓ Ù Ø ÓÒ

¼º 2

½º Ë ÒÙÐÖ Øݸ Ñ×ÙÖ Ð ØÝ Ò ÖÔÖ×ÒØ Ð ØÝ

Ý ×ÓÐÙØÐÝ ÓÒØ ÒÙÓÙ× ÙÒ Ø ÓÒ×

 3

ËÙ Ø Ú ØݸØ Ô Ó ÒØ× Ó ÐÓ Ð ÓÒÚÜ ØÝ

¾º 7

Ö ØÖ ÞØ ÓÒ× Ý ÑÒ× Ó ÙÒ Ø ÓÒÐ ÕÙØ ÓÒ×

¿º 10

Ì ÒØÓÖ ÙÒ Ø ÓÒ ×   ×ØÖ ÙØ ÓÒ ÙÒ Ø ÓÒ

!º 16

Ð ÙÐØ ÓÒ Ó ÑÓÑÒØ× Ò Ø ÐÒØ Ó Ø ÖÔ

#º 19 ËÓÑ ØÓÔ ÓÐÓ Ð ÔÖÓÔ ÖØ ×

$º 26

& Ò ³× Ö ÚØ Ú×

%º 32

Ä ×Ù³× Ö ÚØ Ú

(º 35

À¤ÓÐÖ ÓÒØ ÒÙ Øݸ  ×ØÓÖØ ÓÒ Ó ÀÙ×ÓÖ-  ÑÒ× ÓÒ¸ ¹Ò× Ø ×

*º sc 35 ½

Ì ÒØÓÖ ÙÒ Ø ÓÒ ¾

¼º ÁÒØÖÓ Ù Ø ÓÒ

Ì ÒØÓÖ ÙÒ Ø ÓÒ × 0Ò Ò ÒØÓÖ³× ÔÔ Ö 1Ò 2 Ø ÆÓÚÑÖ ½((¿¸

G Û

Ø 0Ö×Ø 4ÒÓÛÒ ÔÔ ÖÒ  Ó Ø × ÙÒ Ø ÓÒº ÁÒ 1Ò2 5ÓÖ ÒØÓÖ Û× ÛÓÖ4 Ò ÓÒ

ÜØÒ× ÓÒ× Ó Ø 6ÙÒÑÒØÐ ÌÓÖÑ Ó Ð ÙÐÙ× ØÓ Ø × Ó  × ÓÒØ ÒÙÓÙ×

ÙÒ Ø ÓÒ× Ò × ×  ÓÙÒØÖÜÑÔÐ ØÓ ×ÓÑ ÀÖÒ 4³× 7ÖÑØ ÓÒ  ÓÙØ

G ×ÖÚ

×Ù  ÜØÒ× ÓÒ× 1ÀÛ¸ Ôº$¼2º Ì ÒØÖ×Ø Ò Ø Ð× ÖÓÑ Ø ÖÐÝ  ×ØÓÖÝ Ó Ø

ÒØÓÖ ×Ø Ò ÒØÓÖ ÙÒ Ø ÓÒ Ò   ÓÙÒ Ò 6ÐÖÓÒ³× ÒÓØ 16Ð2º Ì × ÙÒ Ø ÓÒ Û×

Ð×Ó Ù× Ý Àº Ä ×Ù Ò  × ÑÓÙ× 8Ä« ÓÒ× ×ÙÖ Ð³ ÒØ¡ÖØ ÓÒ Ø Ð Ö Ö  ×

ÓÒ Ø ÓÒ× ÔÖ Ñ Ø Ú×8¸ ´ÈÖ ×¸ 5ÙØ ֹΠÐÐÖ׸ ½*¼!µº 6ÓÖ Ø × Ö×ÓÒ

G × ×ÓÑØ Ñ×

ÖÖÖ ØÓ × Ø Ä ×Ù ÙÒ Ø ÓÒº ËÓÑ ÒØÖ×Ø Ò ÙÒ Ø ÓÒ Ð××׸ ÑÓØ ÚØ Ý

Ø ÒØÓÖ ÙÒ Ø ÓÒ¸ Ú  Ò ÒØÖÓ Ù  ØÓ ÑÓ ÖÒ ÖÐ ÒÐÝ× ×¸ ׸ ÓÖ ÜÑÔи

Ü ×Ø ÒÙÑÖÓÙ× ÒÖÐ ÞØ ÓÒ× Ó Û 

1?ÖÄ 2 ÓÖ 1ÌÓѸ &0Ò Ø ÓÒ %º¿½2º ÌÖ G

Ö ÓØ Ò × ÚÖ Ø ÓÒ× Ó ÒØÓÖ³× ÓÒ×ØÖÙ Ø ÓÒ× ÙØ Û Ó ÒÓØ ÓÒ× Ö Ø× Ò

ÓÙÖ ÛÓÖ4º Ë Ò    ×ØÖ ÙØ ÓÒ ÙÒ Ø ÓÒ ÓÖ Ø × ÑÔÐ×Ø ÒÓÒØÖ Ú Ð ×Ð¹× Ñ ÐÖ

G ×

Ñ×ÙÖ¸ Ö ØÐ ÓÑØÖÝ × ×ÓÛÒ ÒÛ ÒØÖ×Ø Ò Ø ÒØÓÖ ÙÒ Ø ÓÒº

Ï Ö ÐÐ Ø 0Ò Ø ÓÒ× Ó Ø ØÖÒÖÝ ÒØÓÖ ÙÒ Ø ÓÒ ØÓÖ ×Ø º ÄØ

G Ò Ò C

ÜÔÒ × x ∈ [0, 1] Ò x

∞ a (0.1) x = nx ,a∈{0, 1, 2}. 3n nx

n=1

Ý Ø ×ÑÐÐ×Ø Û Ø  Ø Ü ×Ø× Ò ÔÙØ ∞  ØÖ × ÒÓ

&ÒÓØ  Nx n anx =1 Nx =

×Ù  ÌÒ Ø ÒØÓÖ ÙÒ Ø ÓÒ → R Ò  0Ò × anx º G :[0, 1]

− 1 1 Nx 1 a (0.2) G(x):= + nx . 2Nx 2 2n

n=1

 ØØ Ø × ÒÔ ÒÒØ Ó Ø Ó  Ó ÜÔÒ× ÓÒ ´¼º½µ  × ØÛÓ ØÖÒÖÝ

Ç×ÖÚ x

ÖÔÖ×ÒØØ ÓÒ׺

× Ø ×Ø Ó ÐÐ Ô Ó ÒØ× ÖÓÑ Û  Ú ÜÔÒ× ÓÒ ´¼º½µ ÒØÓÖ ×Ø

Ì C [0, 1]

Ù× Ò ÓÒÐÝ Ø   Ø× ¼ Ò ¾º ÁÒ Ø × ∈{ } µ Ø ÕÙÐ ØÝ ´¼º¾µ

x ∈ C ´anx 0, 2

Ø4× Ø ÓÖÑ

∞ 1 a (0.3) G(x)= nx . 2 2n

n=1

ÓÐÐÓÛ Ò Ð×× Ð ÒÖØ Ú ÓÒ×ØÖÙ Ø ÓÒ ÓÖ Ø ØÖ  ÒØÓÖ ×Ø ×

Ì C

ÑÓÖ Ô ÓÔÙÐÖº

ØÖÚÐ 0Ò ÐÓ× ×Ù×Ø× ⊇ ⊇ ⊇ ⊇

ËØÖØ Ò Û Ø Ø Ò C0 := [0, 1] C1 C2 ... Ck

× ÓÐÐÓÛ׺ Ï ÓØ Ò Ø ×Ø Ý Ø ÖÑÓÚ Ò Ø 8Ñ Ð Ø Ö8 ÒØÖÚÐ ... Ò C0 C1

1 2

ÖÓÑ º ÌÒ Ø ×Ø × ÓØ Ò Ý ÖÑÓÚ Ò ÖÓÑ Ø ÓÔ Ò ÓÔ Ò ( 2 , 3 ) C0 C2 C1

ÒØÓÖ ÙÒ Ø ÓÒ Ì 3

1 2 7 8 k

ØÖÚÐ× Ò º ÁÒ ÒÖи ÓÒ× ×Ø× Ó  ×BÓ ÒØ ÐÓ× ÒØÖÚÐ× Ò¸

Ò ( 9 , 9 ) ( 9 , 9 ) Ck 2

¸ × ÓØ Ò Ý ÖÑÓÚ Ò Ñ Ð Ø Ö× ÖÓÑ   Ó Ø ÒØÖÚÐ× Ú Ò

 Ck Ck+1

º ÌÒ Ø × ×Ý ØÓ × ØØ Ñ4 ÙÔ ØØ Ck ∞ (0.4) C = Ck. k=0

1

Ý Ø ×Ø Ó ÐÐ ÒÔ Ó ÒØ× Ó ÓÑÔÐÑÒØÖÝ ÒØÖÚÐ× Ó Ò ×Ø &ÒÓØ C C

(0.5) C◦ := C\C1,I◦ := [0, 1]\C.

Ð×Ó I   Ñ ÐÝ Ó ÐÐ ÓÑÔ ÓÒÒØ× Ó Ø ÓÔ Ò ×Ø º ÄØ I

Figure 1: The graph of the Cantor G. This graph is sometimes called ”Devil’s staircase”.

1 Singularity, measurability and representability by absolutely continuous functions

The well-known properties of the Cantor function are collected in the following.

Proposition 1.1. 1.1.1. G is continuous and increasing but not absolutely continuous. 1.1.2. G is constant on each interval from I◦.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 4

1.1.3. G is a . 1.1.4. G maps the C onto [0, 1].

Proof. It follows directly from (0.2) that G is an increasing function, and moreover (0.2) implies that G is constant on every interval J ⊆ I◦. Observe also that if x, y ∈ [0, 1],x= y,andx tends to y, then we can take ternary representations (0.1) so that min{n : |anx − any| =0 }→∞. Thus the continuity of G follows from (0.2) as well. Since the one dimensional Lebesgue of C is zero, m1(C)=0, the monotonicity of G and constancy of G on each interval J ⊆ I◦ imply Statement 1.1.3. One easy way to check the last equality is to use the obvious recurrence relation 2 m (C )= m (C − ) 1 k 3 1 k 1 for closed subset Ck in (0.4). Really, by (0.4) 2 k m1(C) = lim m1(Ck) = lim =0. k→∞ k→∞ 3 It still remains to note that by (0.3) we have 1.1.4 and that G is not absolutely continuous, since G is singular and nonconstant. 

Remark 1.2. Recall that a monotone or bounded variation function f is called singular if f  =0a.e.

By the Radon-Nikodym theorem we obtain from 1.1.1 the following.

Proposition 1.3. The function G cannot be represented as x G(x)= ϕ(t)dt 0 where ϕ is a Lebesgue integrable function.

In general a need not map a measurable set onto a measur- able set. It is a consequence of 1.1.4 that the Cantor function is such a function.

Proposition 1.4. There is a Lebesgue measurable set A ⊆ [0, 1] such that G(A) is not Lebesgue measurable.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 5

In fact, a continuous function g :[a, b] → R transforms every measurable set onto a measurable set if and only if g satisfies Lusin’s condition (N):

(m1(E)=0)⇒ (m1(g(E)) = 0) for every E ⊆ [a, b] [Sa, p.224]. Let Lf denote the set of points of constancy of a function f, i.e., x ∈ Lf if f is constant in a neighbourhood of x.Inthecasef = G it is easy to see that ◦ LG = I =[0, 1]\C.

Proposition 1.5. Let f :[a, b] → R be a monotone continuous function. Then the following statements are equivalent 1.5.1. The inverse image f −1(A) is a Lebesgue measurable subset of [a, b] for every A ⊆ R. 1.5.2. The of the set [a, b]\Lf is zero,

(1.6) m1([a, b]\Lf )=0.

Proof. 1.5.2 ⇒ 1.5.1. Evidently, Lf is open (in the relative topology of [a, b]) and f is constant on each component of Lf .LetE be a set of endpoints of components of Lf . Let us denote by f0 and f1 the restrictions of f to Lf ∪ E and to [a, b]\(Lf ∪ E), respectively. | | f0 := f Lf ∪E,f1 := f [a,b]\(Lf ∪E). R −1 If A is an arbitrary subset of ,thenitiseasytoseethatf0 (A)isaFσ subset of [a, b]and −1 ⊆ \ f1 (A) [a, b] Lf . −1 −1 ∪ −1 \ −1 Since f (A)=f0 (A) f1 (A), the equality m1([a, b] Lf ) = 0 implies that f (A) is Lebesgue measurable as the union of two measurable sets. 1.5.1 ⇒ 1.5.2. The monotonicity of f implies that f1 is one-to-one and

(1.7) (f0(Lf ∪ E)) ∩ (f1([a, b]\(Lf ∪ E))) = ∅. ⊆ R ∗ Suppose that (1.6) does not hold. For every B with an outer measure m1(B) > 0 there exists a nonmeasurable set A ⊆ B. See, for instance, [Oxt, Chapter 5, Theorem 5.5]. Thus, there is a nonmeasurable set A ⊆ [a, b]\(Lf ∪ E). Since f1 is one-to-one, equality (1.7) implies that f −1(f(A)) = A, contrary to 1.5.1. 

Corollary 1.8. The inverse image G−1(A) is a Lebesgue measurable subset of [0, 1] for every A ⊆ R.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 6

Remark 1.9. It is interesting to observe that G(A) is a Borel set for each Borel set A ⊆ [0, 1]. Indeed, if f :[a, b] → R is a monotone function with the set of discontinuity D,then

f(A)=f(A ∩ D) ∪ f(A ∩ Lf ) ∪ f(A ∩ E) ∪ f(A\(D ∪ Lf ∪ E)),A⊆ [a, b], where E is the set of endpoints of components of Lf .Thesetsf(A ∩ D), f(A ∩ Lf ) and f(A ∩ E) are at most countable for all A ⊆ [0, 1]. If A is a Borel subset of [a, b], then f(A\(D ∪ Lf ∪ E)) is Borel, because it is the image of the Borel set under the homeomorphism f . [a,b]\(D∪Lf ∪E)

A function f :[a, b] → R is said to satisfy the Banach condition (T1)if

−1 m1({y ∈ R : card(f (y)) = ∞})=0.

◦ Since a restriction G|C◦ is one-to-one and G(I ) is a countable set, 1.1.2 implies

Proposition 1.10. G satisfies the condition (T1).

N.Bary and D.Menchoff [BaMe] showed that a continuous function f :[a, b] → R is a superposition of two absolutely continuous functions if and only if f satisfies both the conditions (T1)and(N). Moreover, if f is differentiable at every point of a set which has positive measure in each interval from [a, b], then f is a sum of two superpositions, f = f1 ◦ f2 + f3 ◦ f4 where fi, i =1, ..., 4, are absolutely continuous [Bar]. Thus, we have

Proposition 1.11. There are absolutely continuous functions f1, ..., f4 such that

G = f1 ◦ f2 + f3 ◦ f4, but G is not a superposition of any two absolutely continuous functions.

Remark 1.12. A superposition of any finite number of absolutely continuous func- tions f1 ◦ f2 ◦ ... ◦ fn is always representable as q1 ◦ q2 with two absolutely continuous q1,q2. Every continuous function is the sum of three superpositions of absolutely continuous functions [Bar]. An application of Proposition 1.10 to the nondifferen- tiability set of the Cantor function will be formulated in Proposition 7.1.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 7

2 Subadditivity, the points of local convexity

An extended Cantor’s ternary function Gˆ is defined as follows   0ifx<0 Gˆ(x)= G(x)if0≤ x ≤ 1  1ifx>1.

Proposition 2.1. The extended Cantor function Gˆ is subadditive, that is

Gˆ(x + y) ≤ Gˆ(x)+Gˆ(y) for all x, y ∈ R.

This proposition implies the following corollary.

Proposition 2.2. The Cantor function G is a first modulus of continuity of itself, i.e., sup |G(x) − G(y)| = G(δ) |x−y|≤δ x,y∈[0,1] for every δ ∈ [0, 1].

The proof of Propositions 2.1 and 2.2 can be found in Timan’s book [Tim, Section 3.2.4] or in the paper of Josef Dobo˘s[Dob]. It is well-known that a function ω :[0, ∞) → [0, ∞) is the first modulus of continuity for a continuous function f :[a, b] → R if and only if ω is increasing, continuous, subadditive and ω(0) = 0 holds. A particular way to prove the subadditivity for an increasing function ω :[0, ∞) → [0, ∞)withω(0) = 0 is to show that the function δ−1ω(δ)isde- creasing [Tim, Section 3.2.3]. The last condition holds true in the case of concave functions. The following propositions show that this approach is not applicable for the Cantor function. Let f :[a, b] → R be a continuous function. Let us say that f is locally concave- convex at point x ∈ [a, b] if there is a neighbourhood U of the point x for which f|U is either convex or concave. Similarly, a continuous function f :(a, b) → R is said to be locally monotone at x ∈ (a, b) if there is an open neighbourhood U of x such that f|U is monotone.

−1 Proposition 2.3. The function x G(x) is locally monotone at point x0 if and only ◦ if x0 ∈ I .

ÒØÓÖ ÙÒ Ø ÓÒ Ì 8

◦ Proposition 2.4. G(x) is locally concave-convex at x0 if and only if x0 ∈ I .

Propositions 2.3 and 2.4 are particular cases of some general statements.

Theorem 2.5. Let f :[a, b] → R be a monotone continuous function. Suppose that Lf is an everywhere dense subset of [a, b].Thenf is locally concave-convex at x ∈ [a, b] if and only if x ∈ Lf .

For the proof we use of the following.

Lemma 2.6. Let f :[a, b] → R be a continuous function. Then [a, b]\Lf is a compact perfect set.

Proof. By definition Lf is relatively open in [a, b]. Hence [a, b]\Lf is a compact subset of R.Ifp is an isolated point of [a, b]\Lf , then either it is a common endpoint of two intervals I,J which are components of Lf or p ∈{a, b}. In the first case it follows by continuity of f that I ∪{p}∪J ⊆ Lf . That contradicts to the maximality of the connected components I,J. The second case is similar. 

Proof of Theorem 2.5. It is obvious that f is locally concave-convex at x for x ∈ Lf . Suppose now that x ∈ [a, b]\Lf and f is concave-convex in a neighbourhood U0 of the point x. By Lemma 2.6 x is not an isolated point of [a, b]\Lf . Hence, there exist y,z in [a, b]\Lf and δ>0 such that

z

Since [a, b]\Lf is perfect, we can find z0,y0 for which

z0 ∈ (z − δ, z) ∩ ([a, b]\Lf ),y0 ∈ (y,y + δ) ∩ ([a, b]\Lf ).

Denote by l z and ly the straight lines which pass through the points (z0,f(z0)), z+y z+y z+y z+y 2 ,f 2 and (y0,f(y0)), 2 ,f 2 , respectively. Suppose that f is in- creasing. Then the point (z, f(z)) lies over lz but (y,f(y)) lies under ly (See Figure | 2). Hence, the restriction f (z0,y0) is not concave-convex. This contradiction proves the theorem since the case of a decreasing function f is similar. 

Theorem 2.7. Let f :(a, b) → [0, ∞) be an increasing continuous function and let ϕ :(a, b) → [0, ∞) be a strictly decreasing function with finite ϕ(x) at every x ∈ (a, b).If

(2.8) m1(Lf )=|b − a|, then the product f ·ϕ is locally monotone at a point x ∈ (a, b) if and only if x ∈ Lf .

ÒØÓÖ ÙÒ Ø ÓÒ Ì 9

Figure 2: The set of points of constancy is the same as the set of points of convexity.

For a function f :(a, b) → R and x ∈ (a, b)set

f(x +∆x) − f(x − ∆x) SDf(x) = lim ∆x→0 2∆x provided that the limit exists, and write

Vf := {x ∈ (a, b):SDf(x)=+∞}.

Using the techniques of differentation of Radon measures (See [EvGa, in particular Section 1.6, Lemma 1]), we can prove the following.

Lemma 2.9. Let f :(a, b) → R be continuous increasing function. If equality (2.8) holds, then Vf is a dense subset of (a, b)\Lf .

Proof of Theorem 2.7. It is obvious that ϕ · f is locally monotone at x0 for every x0 ∈ Lf . Suppose that ϕ · f is monotone on an open interval J ⊂ (a, b)and x0 ∈ ((a, b)\Lf ) ∩ J.SinceLf is an everywhere dense subset of (a, b), there exists ⊆ ∩ · | · | an interval J1 J Lf .Thenϕ f J1 is a decreasing function. Consequently, ϕ f J is decreasing too. By Lemma 2.9 we can select a point t0 ∈ Vf ∩ J. Hence, by the definition of Vf |  ∞ SDϕ(x)f(x) x=t0 = ϕ (t0)f(t0)+ϕ(t0)SDf(t0)=+ .

This is a contradiction, because ϕ · f|J is decreasing. 

ÒØÓÖ ÙÒ Ø ÓÒ Ì 10

Remark 2.10. Density of Lf is essential in Theorem 2.5. Indeed, if A is a closed subset of [a, b] with nonempty interior IntA, then it is easy to construct a continuous increasing function f such that Lf =[a, b]\A and f is linear on each interval which belongs to IntA. Theorem 2.7 remains valid if the both functions f and ϕ are negative, but if f and ϕ have different signs, then the product f · ϕ is monotone on (a, b). Functions having a dense set of constancy have been investigated by A.M.Bruckner and J.L.Leonard in [BrLe]. See also Section 6 in the present work.

3 Characterizations by means of functional equa- tions

There are several characterizations of the Cantor function G basedontheself- similarity of the Cantor ternary set. We start with an iterative definition for G. Define a sequence of functions ψn :[0, 1] → R by the rule   1 1  ψ (3x)if0≤ x ≤  2 n 3   1 1 2 (3.1) ψn+1(x)= if

Proposition 3.2. The Cantor function G is the unique element of M[0, 1] for which   1 1  G(3x) if 0 ≤ x ≤  2 3   1 1 2 (3.3) G(x)= if

ÒØÓÖ ÙÒ Ø ÓÒ Ì 11

Proof. Define a map F : M[0, 1] →M[0, 1] as   1 1  f(3x),0≤ x ≤ ,  2 3   1 1 2 F (f)(x)= ,

It should be observed here that there exist several iterative definitions for the Cantor ternary function G. The above method is a simple modification of the cor- responding one from Dobo˘s’s article [Dob]. It is interesting to compare Proposition 3.2 with the self-similarity property of the Cantor set C. Let for x ∈ R 1 1 2 (3.4) ϕ (x):= x, ϕ (x):= x + . 0 3 1 3 3

Proposition 3.5. The Cantor set C is the unique nonempty compact subset of R for which

(3.6) C = ϕ0(C) ∪ ϕ1(C) holds. Further if F is an arbitrary nonempty compact subset of R, then the iterates

k+1 k k 0 Φ (F ):=ϕ0(Φ (F )) ∪ ϕ1(Φ (F )), Φ (F ):=F, convergence to the Cantor set C in the Hausdorff metric as k →∞.

Remark 3.7. This theorem is a particular case of a general result by Hutchinson about a compact set which is invariant with respect to some finite family contraction maps on Rn [Hut].

ÒØÓÖ ÙÒ Ø ÓÒ Ì 12

In the case of a two-ratio Cantor set a system which is similar to (3.3) was found by W.A.Coppel in [Cop]. The next theorem follows from Coppel’s results.

Proposition 3.8. Every real–valued F ∈M[0, 1] that satisfies x F (x) (3.9) F = , 3 2 x F (x) (3.10) F 1 − =1− , 3 2  1 x 1 (3.11) F + = 3 3 2 is the Cantor ternary function.

The following simple characterization of the Cantor function G has been sug- gested by D.R.Chalice in [Ch].

Proposition 3.12. Every real–valued increasing function F :[0, 1] → R satisfying (3.9) and

(3.13) F (1 − x)=1− F (x) is the Cantor ternary function.

Remark 3.14. Chalice used the additional condition F (0) = 0, but it follows from (3.9).

The functional equations, given above, together with Proposition 3.5 are some- times useful in applictions. As examples, see Proposition 4.5 in Section 4 and Propo- sition 5.1 in Section 5.

The system (3.9)+(3.13) is a particular case of the system that was applied in the earlier paper of G. C. Evans [Ev] to the calculation of the moments of some Cantor functions. In this interesting paper Evans noted that (3.9) and (3.13) together with continuity do not determine G uniquely. However, they imply that  1 2 2 (3.15) F (x)= + F x − , ≤ x ≤ 1. 2 3 3 Next we show that a variational condition, together with (3.9) and (3.13), de- termines the Cantor function G.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 13

Proposition 3.16. Let F :[0, 1] → R be a continuous function satisfying (3.9) and (3.13). Then for every p ∈ (1, +∞) it holds 1 1 (3.17) |F (x)|pdx ≥ |G(x)|pdx, 0 0 and if for some p ∈ (1, ∞) an equality 1 1 (3.18) |F (x)|pdx = |G(x)|pdx 0 0 holds, then G = F .

Figure 3: The continuous function F which satisfies ”the two Cantor function equa- 1 − 2 1 1 tions” (3.9) and (3.13) but F (x)= 2 (12x 5) on 3 , 2 .

Lemma 3.19. Let f :[a, b] → R be a continuous function and let p ∈ (1, ∞ ). a+b a+b Suppose that the graph of f is symmetric with respect to the point 2 ; f 2 . Then the inequality  b p 1 | |p ≥ a + b (3.20) | − | f(x) dx f b a a 2 holds with equality only for  a + b f(x) ≡ f . 2

ÒØÓÖ ÙÒ Ø ÓÒ Ì 14

Proof. We may assume without loss of generality that a = −1, b =1.Nowfor f(0) = 0 inequality (3.20) is trivial. Hence replacing f with −f, if necessary, we may assume that f(0) < 0. Write

Ψ(x)=f(x) − f(0).

Given m>0, we decompose [−1, 1] as

A+ := {x ∈ [−1, 1] : |Ψ(x)| >m},A0 := {x ∈ [−1, 1] : |Ψ(x)| = m},

A− := {x ∈ [−1, 1] : |Ψ(x)|

g(y):=(B − y)p +(B + y)p is strictly increasing on (0,B)forp>1. Hence, + ≥ | |p | |p ≥ p (3.21) Jm 2 Ψ(x) dx = Ψ(x) dx m dx. A+∩[0,1] A+ A+

If Ψ(x) ≡ 0, then A− is nonempty and open, as Ψ(0) = 0. Thus, − p p (3.22) Jm > 2 m dx = m dx A−∩[0,1] A− for Ψ(x) ≡ 0. Moreover, it is obvious that  0 p−1 p ≥ p (3.23) Jm =2 2 m dx m dx. A0∩[0,1] A0

Choose m = −f(0). Then, in the case where f(x) ≡ f(0), from (3.21)–(3.23) we obtain the required inequality 1 1 |f(x)|p > |f(0)|p. 2 −1 

ÒØÓÖ ÙÒ Ø ÓÒ Ì 15

Remark 3.24. Inequality (3.20) holds also for p = 1, but, as simple examples show, in this case the equality in (3.20) is also attained by functions different from the constant function.

For every interval J ∈I(where I is a family of components of the open set ◦ I =[0, 1]\C) let us denote by xJ the center of J and by GJ the value of the Cantor function G at xJ .

Lemma 3.25. If F :[0, 1] → R is a continuous function satisfying (3.9) and (3.13), then the graph of the restriction F |J is symmetric with respect to the point (xJ ,GJ ) for each J ∈I.

Proof. It follows from (3.13) that   1 1 1 F = G = , 2 2 2 thus we can rewrite equation (3.13) as    1 1 1 1 F = F + x + F − x . 2 2 2 2

1 1 Consequently a graph of F | 1 2 is symmetric with respect to the point , .IfJ is ( 3 , 3 ) 2 2 I an arbitrary element of , then there is a finite sequence of contractions ϕi1, ..., ϕin 1 2 ◦ ◦ such that J is an image of the interval 3 , 3 under superposition ϕi1 ... ϕin and each ϕik, k =1, ..., n belongs to {ϕ0,ϕ1} (see formula (3.4)). Now the desired follows from (3.9) and (3.15) by induction. 

Proof of Proposition 3.16. Let J ∈I.IfF satisfies (3.9) and (3.13), then Lemma 3.25 and Lemma 3.19 imply that (3.26) |F (x)|pdx ≥ |G(x)|pdx J J for every p ∈ (1, ∞). Thus, we obtain (3.17) from the condition m1(C) = 0. Suppose now that (3.18) holds. Then we have the equality in (3.26) for every J ∈I.Thus, by Lemma 3.19 F |J = G|J ,J∈I. Since I0 = ∪ J is a dense subset of [0, 1] and F = G as required.  J∈I

In the special case p = 2 we can use an orthogonal projection to prove Proposition 3.16.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 16

2 2 ∈ 2 Define a subspace LC [0, 1] of the Hilbert space L [0, 1] by the rule: f LC [0, 1] 2 if f ∈ L [0, 1] and for every J ∈Ithere is a constant CJ such that 2 |f(x) − CJ | dx =0. J ∈ 2 It is obvious that G LC [0, 1]. 2 Let us denote by PC the operator of the orthogonal projection from L [0, 1] to 2 LC [0, 1].

Proposition 3.27. Let f be an arbitrary function in L2[0, 1] and let J ∈Ibe an interval with endpoints aJ ,bJ . 3.27.1 Suppose that f is continuous. Then the image PC (f) is continuous if and only if 1 f(aJ )=f(bJ )=| − | f(x)dx bJ aJ J for each J ∈I. 2 3.27.2. Function f ∈ L [0, 1] is a solution of the equation PC (f)=G if and only if f(x)dx = G(x)dx J J for each J ∈I.

We leave the verification of this simple proposition to the reader. Now, the conclusion of Proposition 3.16 follows directly from Lemma 3.25 by usual properties of orthogonal projections. See, for example, [BSU, Chapter VII, 9].

4 The Cantor function as a distribution function

It is well known that there exists an one-to-one correspondence between the set of the Radon measures µ in R and the set of a finite valued, increasing, right continuous functions F on R with lim F (x) = 0. Here we study the corresponding measure x→−∞ for the Cantor function. Let ϕ0,ϕ1 : R → R be similarity contractions defined by formula (3.4). Write lg 2 (4.1) s := . c lg 3

sc We let H denote the sc-dimensional Hausdorff measure in R.See[Fal1]for properties of Hausdorff measures.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 17

Proposition 4.2. There is the unique Borel regular probability measure µ such that 1 1 (4.3) µ(A)= µ(ϕ−1(A)) + µ(ϕ−1(A)) 2 0 2 1 for every Borel set A ⊆ R. Furthemore, this measure µ coincides with the restriction of the Hausdorff measure Hsc to C, i.e., (4.4) µ(A)=Hsc (A ∩ C) for every Borel set A ⊆ R.

For the proof see [Fal1, Theorem 2.8, Lemma 6.4].

Proposition 4.5. Let m1 be the Lebesgue measure on R.Then

sc (4.6) m1(G(A)) = H (A) for every Borel set A ⊆ C.

Proof. Write µ(A):=m1(G(A ∩ C)) for every Borel set A ⊆ R. If suffices to show that µ is a Borel regular probability 1 1 measure which fulfils (4.3). Since C is countable (see formula (0.5)), m1(G(C )) = 0. Hence, we have ◦ µ(A)=m1(G(C ∩ A)),A⊆ R. ◦ ◦ The restriction G|C◦ : C → G(C ) is a homeomorphism and ◦ µ(R)=m1(G(C )) = 1. Hence, µ is a Borel regular probability measure. (Note that G(A)isaBorelset for each Borel set A ⊆ C. See Remark 1.9.) To prove (4.3) we will use following functional equations (see (3.9), (3.15)).

x 1 (4.7) G = G(x), 0 ≤ x ≤ 1, 3 2  1 2 2 (4.8) G(x)= + G x − , ≤ x ≤ 1. 2 3 3 Let A be a Borel subset of R.Put     1 2 A := A ∩ 0, ,A:= A ∩ , 1 . 0 3 1 3

ÒØÓÖ ÙÒ Ø ÓÒ Ì 18

−1 −1 − 2 Since ϕ0 (x)=3x and ϕ1 (x)=3 x 3 ,wehave −1 −1 µ(ϕi (A)) = µ(ϕi (Ai)) for i =1, 2. It follows from (4.7) that

−1 2G(x)=G(ϕ0 (x)),   ∈ 1 for x 0, 3 . Hence, 1 1 1 µ(ϕ−1(A )) = m (G(ϕ−1(A ))) = m (2G(A )) = m (G(A )). 2 0 0 2 1 0 0 2 1 0 1 0 Similarly, (4.7) implies that  2 2G x − = G(ϕ−1(x)) 3 1   ∈ 2 for x 3 , 1 ,andweobtain 1 1 µ(ϕ−1(A )) = m (G(ϕ−1(A ))) = 2 1 1 2 1 1 1   1 2 2 = m 2G A − = m G A − . 2 1 1 3 1 1 3 Now using (4.8) we get   2 1 m G A − = m − + G(A ) = m(G(A )). 1 1 3 1 2 1 1 ∩ 1 The set G(A1) G(A0) is empty or contains only the point 2 . Hence we obtain

m1(G(A0)) + m1(G(A1)) = m1(G(A0 ∪ A1)) = µ(A). Therefore µ(A) satisfies (4.3). 

Corollary 4.9. The extended Cantor function Gˆ is the cumulative distribution function of the restriction of the Hausdorff measure Hsc to the Cantor set C.

This description of the Cantor function enables us to suggest a method for the proof of the following proposition. Write

Gˆh(x):=Gˆ(x + h) − Gˆ(x) ˆ for each h ∈ R. The function Gh is of bounded variation, because it equals a difference of two increasing bounded functions.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 19

Proposition 4.10. [HiTa] Let Var(Gˆh) be a total variation of Gˆh. Then we have

sup Var(Gˆh)=2 0≤h≤δ for every δ>0.

This holds because sets C ∩ (C ± 3−n) have finite numbers of elements for all positive integer n.

Remark 4.11. If F is an absolutely continuous function of bounded variation and Fh(x):=F (x + h) − F (x), then

(4.12) lim Var(Fh)=0. h→0 Really, if F is absolutely continuous on [a, b], then F  exists a.e. in [a, b], F  ∈ L1[a, b], and b Var(F )= |F (x)| dx.

a Approximating of F  by continuous functions, for which the property is obvious, we obtain (4.12).

In fact, Proposition 4.10 remains valid for an arbitrary singular function of bounded variation.

Theorem 4.13 Let F : R → R be a function of bounded variation with singular part ϕ. Then the limit relation

lim sup(Var(Fh)) = 2Var(ϕ) h→0 holds.

The theorem is an immediate adaptation of the result that was proved by Norbert Wiener and R. C. Young in [WY].

5 Calculation of moments and the length of the graph

In the article [Ev], G.C.Evans proved the recurrence relations for the moments 1 n x Gα(x)dx 0

ÒØÓÖ ÙÒ Ø ÓÒ Ì 20

where Gα is a ”Cantor function” for a α-middle Cantor set. In the classic middle- third case Evans’s results can be written in the form.

Proposition 5.1. Let n be a natural number and let Mn be a moment of the form 1 n Mn = x G(x)dx. 0 Then the following relations hold

n−1 1 1 n (5.2) 2M = + 2n−kM , n n +1 3n+1 − 1 k k k=0

n−1 1 n (5.3) (1 + (−1)n)M = + (−1)k+1 M n n +1 k k k=0

n for all positive integers n where k are the binomial coefficients.

− 1 It follows immediately from G(x)+G(1 x)=1thatM0 = 2 . Hence, (5.2) can be used to compute all moments Mn.LetµC be the restriction of the Hausdorff measure Hsc to the Cantor set C (see Corollary 4.9). Now set 1 n mn := x dµC(x). 0

Proposition 5.4. The following equality holds

n n+1 n−k k 2 mk (5.5) m = k=0 , n+1 3n+1 − 1 for every natural n and we have

− n 1 n (5.6) 2m = (−1)k m , n k k k=0

−  n 1 n +1 (5.7) (n +1)m = (−1)k m n k k k=0 for every odd n.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 21

Proof. First we prove (5.5). ∩∞ The Cantor set C can be defined as the intersection k=0Ck, see (0.4). Each Ck k i i i k i consists of 2 disjoint closed intervals Ck =[ak,bk], i =1, ..., 2 . The length of Ck is 3−k and i i − i −k µC (Ck)=G(bk) G(ak)=2 . i ⊆ Let Lk be a set of all left-hand points of intervals Ck Ck, i.e. { i k} Lk = ak : i =1, ..., 2 .

It is easy to see that L0 = {0} and   1 2 1 (5.8) L = L − ∪ + L − k 3 k 1 3 3 k 1 for k ≥ 1. Observe that 1 1 n n x dµC(x)= x dG(x) 0 0 because the integrand is continuous. Hence we have

l n mk = lim (ξi) [G(xj+1) − G(xj)] l→∞ j=1 where 0 = x1 < ... < xl

ÒØÓÖ ÙÒ Ø ÓÒ Ì 22

In order to prove (5.6) we can use (5.3), because 1 1 n+1 n+1 1 − n − (5.9) mn+1 = x dG(x)=x G(x) 0 (n+1) x G(x)dx =1 (n+1)Mn. 0 0 Suppose that n is even, then it follows from (5.3) and (5.9) that

n−1 n 2(1 − m )=2(n +1)M =1+(n +1) (−1)k+1 M n+1 n k k k=0

n−1 n +1 n =1+ (−1)k+1 (1 − m ). k +1 k k+1 k=0 n+1 n n+1 − Hence from k+1 k = k+1 and (1 m0)=0weget

−  n 1 n +1 2(1 − m )=1+ (−1)k+1 (1 − m ) n+1 k +1 k+1 k=0   n n +1 n n +1 =1+ (−1)k (1 − m )=1+ (−1)k k k k k=1 k=0  n n +1 − (−1)k m . k k k=0 Observe that   n n +1 n +1 (−1)k =(1− 1)n+1 − (−1)n+1 =1 k n +1 k=0 because n is even. Consequently,

n k 2(1 − mn+1)=2− (−1) mk. k=0

The last formula implies (5.6). The proof of (5.7) is analogous to that of (5.6). 

Remark 5.10. The measure µC is frequently referred to as the Cantor measure. In mn the proof of (5.5) we used an idea from [HiTa]. By (5.5) with τn = n!2n we obtain  1 τ τ − τ τ = n + n 1 + ... + 0 . n+1 2(3n+1 − 1) 1! 2! (n +1)!

ÒØÓÖ ÙÒ Ø ÓÒ Ì 23

The asymptotics of the moments mn was determinated in [GoWi]. Write 2πim ν(m):=1− lg 3 for m ∈ Z.

th Theorem 5.11.Let mn be the n moment of the Cantor measure µC ,then

1/2−3sc/2 −sc (5.12) mn ∼ 2 n exp(−2H(n)),n→∞, where

+∞ 1  1 H(x):= 2−ν(m)(1 − 21−ν(m))ζ(ν(m))Γ((ν(m))x1−ν(m) 2πi m m∈Z m=0 and lg 2 s = . c lg 3

Note that H(x) is a real-valued function for x>0 and periodic in the vari- able log x. See also [Fis] for an example of an absolutely continuous measure with asymptotics of moments containing oscillatory terms. The behavior of the integrals 1 1 (5.13) I(λ):= (G(x))λdx, E(λ)= exp(λG(x)) dx 0 0 was described in [GorKu]. It was noted that I extends to a function which is analytic in the half-plane Re(λ) > −sc and E extends to an entire function. The following theorem is a particular case of the results from [GorKu].

Theorem 5.14. 5.14.1 For Re(λ) > −sc, the function I obeys the formula 1 (3 · 2λ − 1)I(λ)=1+ (1 + G(x))λdx, 0 and for all λ ∈ C the function E obeys the formula

3E(2λ)=eλ +(eλ +1)E(λ)

ÒØÓÖ ÙÒ Ø ÓÒ Ì 24

5.14.2 For all natural n ≥ 2, we have  k−1 1  n 2 − 1 B(k) I(n)= − n +1 k 3 · 2k−1 − 1 n − k +1 k≤n where the primes mean that summation is over even positive k and B(k) are Bernoulli numbers 5.14.3 For λ →∞, we have

sc −sc−1 λ I(λ)=Φ(log2 λ)+O(λ ),

sc − −sc−1 λ E( λ)=Φ(log2 λ)+O(λ ) where Φ is the function analytic in the strip |Im(z)| <π/(2 lg 2),

+∞ 2 2πin 2πin Φ(z)= Γ sc + ζ sc + exp(−2πinz). −∞ 3lg2 lg 2 lg 2

Proposition 5.15. The equality ∞ 1 a a (5.16) eaxdG(x)=exp cosh 2 3k 0 k=1 holds for every a ∈ C.

Proof. Write 1 Φ(a):= eaxdG(x). 0 It follows from (3.9)–(3.11) that  x 1 2 x 1 1 G = G(x),G + = + G(x). 3 2 3 3 2 2 Hence 1/3 1 1 ax x Φ(a)= eaxdG(x)+ eaxdG(x)= e 3 dG 0 2/3 0 3  1 1 1 2 x 2 x 1 ax 1 2 x + ea( 3 + 3 )dG + = e 3 dG(x)+ ea( 3 + 3 )dG(x) 0 3 3 2 0 2 0 1 2 a a a a = (1 + ea 3 )Φ = e 3 cosh Φ . 2 3 3 3

ÒØÓÖ ÙÒ Ø ÓÒ Ì 25

Consequently, by successive iteration we obtain a a a a a a a Φ(a)=exp + + ... + cosh cosh ...cosh Φ . 3 9 3n 3 9 3k 3k Since Φ(x) → 1asx → 0, the last formula implies (5.16). 

1 Observe that the function Φ, Φ(a)= eax dG(x), is the exponential generating 0 function of the moment sequence {mn}.  The next result follows from (5.16) and shows that Fourier coefficientsµ ˆn,ˆµn := 1 2πinx →∞ 0 e dµC(x), of µC do not tend to zero as n .

Proposition 5.17. [HiTa]. For all integers n, ∞  2πn µˆ = eπin cos . n 3j j=1

Corollary 5.18. [HiTa]. Let k be a positive integer and let n =3k.Then ∞  2π µˆ = − cos =const=0 . n 3ν ν=1

∞ 2π Remark 5.19. The infinite product cos 3ν converges absolutely because ν=1 2π π π2 |1 − cos |≤2sin2 < 2 , 3ν 3ν 9ν ∞ 2π  and hence cos 3ν =0. ν=1

In the paper [WW] Wiener and Wintner proved the more general result similar to Corollary 5.18. The study of the Fourier asymptotics of Cantor type measures has been extended in [Str1, 2, 3], [LaWa], [HuLa].

Proposition 5.20. The length of the arc of the curve y = G(x) between the points (0, 0) and (1, 1) is 2.

Remark 5.21. The detailed proof of this proposition can be found in [Dar1]. In fact, this follows rather easily from 1.1.4. Probably the length of the curve y = G(x) was first calculated in [HiTa].

ÒØÓÖ ÙÒ Ø ÓÒ Ì 26

The next theorem is a refinement of Proposition 5.20.

Theorem 5.22. Let F :[0, 1] → R be a continuous, increasing function for which F (0) = 0 and F (1) = 1. Then the following two statements are equivalent. 5.22.1 The length of the arc y = F (x), 0 ≤ x ≤ 1,is2. 5.22.2 The function F is singular.

This theorem follows from the results of M.J.Pelling [Pel]. See also [DoMa].

6 Some topological properties

There exists a simple characterization of the Cantor function up to a homeomor- phism. If Φ : X → Y and F : X → Y are continuous functions from topological space X to topological space Y ,thenF is said to be (topologically) isomorphic to Φ if there exist homeomorphisms η : X → X and ψ : Y → Y such that F = ψ ◦ Φ ◦ η [RoFu, Chapter 4, §1, Section 1]. If the last equality holds with ψ equal the identity function, then F and Φ will be called the Lebesgue equivalent functions (Cf. [GNW, Definition 1.1]).

Proposition 6.1. The Cantor function G is isomorphic (as a map from [0, 1] into [0, 1]) to a continuous monotone function q :[0, 1] → [0, 1] if and only if the set of constancy Lg is everywhere dense in [0, 1] and

(6.2) g({0, 1})={0, 1}⊆[0, 1]\Lg.

Lemma 6.3. Let A, B be everywhere dense subsets of [0, 1] and let f : A → B be an increasing bijective map. Then f is a homeomorphism and, moreover, f can be extended to a self-homeomorphism of the closed interval [0, 1].

Proof. It is easy to see that

0 = lim f(x)=1− lim f(x) x→0 x→1 x∈A x∈A and lim f(x) = lim f(x) x→t x→t x∈A∩[0,t] x∈A∩[t,1] for every t ∈ (0, 1). Thus there exists the limit

f˜(t) := lim f(x),t∈ [0, 1]. x→t x∈A

ÒØÓÖ ÙÒ Ø ÓÒ Ì 27

Obviously, f˜ is strictly increasing and f˜(t)=f(t)foreacht ∈ A. Suppose that

f˜−(x ) := lim f˜(x) < lim f˜(x):=f˜+(x ) 0 x→x0 x→x0 0 x∈[0,x0) x∈(x0,1] for some x0 ∈ (0, 1). Since B is a dense subset of [0, 1], there is b0 ∈ B such that ˜− ˜+ f(x0) = b0 ∈ (f (x0), f (x0)).

Let b0 = f(a0), where a ∈ A. Now we have the following: ˜ (i) a0 = x0, because b0 = f(x0), ˜− (ii) a0 ∈/ [0,x0), because b0 > f (x0), ˜+ (iii) a0 ∈/ (x0, 1], because b0 < f (x0), and this yields a contradiction. Reasoning similary, we can prove the continuity of f˜ for the points 0 and 1. Hence, f˜ is a continuous bijection of the compact set [0, 1] onto itself and every such bijection is a homeomorphism. 

Proof of Proposition 6.1. It follows directly from the definition of isomorphic functions that (6.2) holds and

(6.4) Clo(Lg)=[0, 1], if G is isomorphic to a continuous function g :[0, 1] → [0, 1]. Suppose now that g :[0, 1] → [0, 1] is a continuous monotone function for which (6.2) and (6.4) hold. We may assume, without loss of generality, that g is increasing. It follows from Lemma 2.6, that a set [0, 1]\Lg is compact and perfect. Moreover, (6.2) and (6.4) imply that [0, 1]\Lg is a nonempty nowhere dense subset of [0, 1]. Hence, there exists an increasing homeomorphism η :[0, 1] → [0, 1] such that

η(C)=[0, 1]\Lg.

0 In fact, there is an order preserving homeomorphism η : C → [0, 1]\Lg [Al, Chapter 4, §6, Theorem 25]. It can be extended on each complementary interval J ∈I as a linear function. The resulting extension η is strictly increasing and maps [0, 1] onto [0, 1]. Hence, by Lemma 6.3, η is a homeomorphism. It is easy to see that a set of constancy of g ◦ η coincides with I◦,thesetof constancy of G.Set gI◦ := g(J),GI◦ := G(J) J∈I J∈I (g(J)andG(J) are one-point sets for every J ∈I). The maps IJ → g(J) ∈ gI◦ and IJ → G(J) ∈ GI◦ are one-to-one and onto. Hence the map

◦ ◦ Ψ : gI◦ → GI◦ , Ψ (g(J)) = G(J),J∈I

ÒØÓÖ ÙÒ Ø ÓÒ Ì 28 is a bijection. It follows from the definition of G and (6.4) that

Clo(gI◦ )=Clo(GI◦ )=[0, 1].

Moreover, since g and G are increasing functions, the function Ψ◦ is strictly increas- ing. Hence, by Lemma 6.3, Ψ◦ can be extended to a homeomorphism ψ :[0, 1] → [0, 1]. It is easy to see that G(x)=ψ(g(η(x))) for every x ∈ I◦.SinceI◦ is a everywhere dense subset of [0, 1], the last equality implies that G = ψ ◦ g ◦ η. Thus g and G are topologically isomorphic. 

Let f :[0, 1] → R be a continuous function whose total variation is finite. Then, by the theorem of Bruckner and Goffman [BrGo], f is Lebesgue equivalent to some function with bounded derivative. As a corollary we obtain

Proposition 6.5. There exists a differentiable function f with an uniformly bounded derivative f  such that G and f are Lebesgue equivalent.

We now recall a notion of set of varying monotonicity ([GNW, Definition 3.7]). Let f :[a, b] → R.Apointx ∈ (a, b) is called a point of varying monotonicity of f if there is no neighborhood of x on which f is either strictly monotonic or constant. We also make the convention that both a and b are points of varying monotonicity for every f :[a, b] → R. Let us denote by Kf the set of points of varying monotonicity of f. Then, as was shown by Bruckner and Goffman in [BrGo], m1(f(Kf )) = 0 if and only if f is Lebesgue equivalent to some continuously differentiable function. For every homeomorphism ψ :[0, 1] :→ [0, 1] we evidently have

Kψ◦G = C, ψ(G(C)) = [0, 1].

Thus, we obtain.

Proposition 6.6. If F is topologically isomorphic to the Cantor function G,then F is not continuously differentiable.

Let GC be the restriction of G to the Cantor set C. Asitiseasytosee,GC is continuous, closed but not open. For example we have        1 1 1 G 0, ∩ C = G 0, = 0, . C 2 2 2

ÒØÓÖ ÙÒ Ø ÓÒ Ì 29

However, GC is weakly open in the following sence.

Proposition 6.7. Let A be a nonempty subset of C. If the interior of A is nonempty in C, then the interior of GC (A) is nonempty in [0, 1],orinotherwords

(6.8) (IntC A = ∅) ⇒ Int[0,1]G(A) = ∅) .

◦ Proof. Suppose that A ⊆ C and IntC A = ∅.SinceC is a dense subset of C,there is an interval (a, b) such that both a and b are in C◦ and

A ⊇ (a, b) ∩ C.

It is easy to see that

(6.9) (x

(6.10) (G(a),G(b)) = G((a, b) ∩ C) ⊆ G(A)=GC (A). 

Recall that a subset B of topological space X is residual in X if X\B is of the first category in X.

−1 Proposition 6.11. If B is a residual (first category) subset of [0, 1],thenGC (B) is residual (first category) in C.

Proof. Let B be a residual subset of [0, 1]. Write

◦ −1 K1 := [0, 1]\C ,W:= G (B). It is sufficient to show that W ∩ C◦ is residual in C. Since G is one–to–one on C◦,

◦ ◦ (6.12) K1 ∪ C =[0, 1] = G(K1) ∪ G(C ), and

◦ ◦ (6.13) C ∩ K1 = ∅ = G(C ) ∩ G(K1). Moreover, we have ∩ ◦ −1 ∩ ◦ −1 ∩ ◦ (6.14) W C = G (G(W ) G(C )) = GC (G(W ) G(C )).

ÒØÓÖ ÙÒ Ø ÓÒ Ì 30

It follows immediately from (6.12) and (6.13) that

◦ G(W ) ∩ G(C )=G(W ) ∩ ([0, 1]\G(K1)).

◦ The set G(K1) is countable. Thus [0, 1]\G(K1) is residual and G(W ) ∩ G(C )is { }∞ residual too, as an intersection of residual sets. Hence, there is a sequence On n=1 for which ∞ ◦ (6.15) G(W ) ∩ G(C ) ⊇ On n=1 ◦ and each On is a dense open subset of [0, 1]. Since GC is continuous on C , (6.14) and (6.15) imply that ∞ ∩ ◦ ⊇ −1 W C GC (On), n=1 −1 where each GC (On)isopeninC. Suppose that we have \ −1  ∅ IntC (C GC (On0 )) = for some positive integer n0. Then by Proposition 6.7 we obtain \  ∅ Int[0,1]([0, 1] On0 ) = { }∞ −1 contrary to the properties of On n=1.ConsequentlyGC (On)isdenseinC for each positive integer n. It follows that W ∩ C◦ is residual in C. For the case where B is of the first category in [0, 1] the conclusion can be obtained by passing to the complement. 

Remark 6.16. In a special case this proposition was mentioned without proof in the work by J.A.Eidswick [Eid].

Proposition 6.17. Let B ⊆ C.ThenB is a everywhere dense subset of C if and only if GC (B) is a everywhere dense subset of [0, 1].

Proof. Since GC is continuous, the image GC (B) is a dense subset of [0, 1] for every dense B. Suppose that Clo[0,1](GC (B)) = [0, 1] but C\CloC (B) = ∅.

GC is a closed map, hence

(6.18) GC (CloC (B)) ⊇ Clo[0,1](GC (B)) = [0, 1].

ÒØÓÖ ÙÒ Ø ÓÒ Ì 31

As in the proof of Proposition 6.7 we can find a two-point set {a, b}⊆C◦ such that (6.10) holds with A = C\CloC (B). Taking into account implication (6.9) we obtain that G(x) ∈/ (G(a),G(b)) for every x ∈ CloC (B), contrary to (6.18). 

Remark 6.19. If (X, τ1,τ2) is a space with two topological structures τ1 and τ2, then one can prove that the condition  ∅ ⇔  ∅ ∀ ⊆ (Intτ1 (A) = ) (Intτ2 (A) = ) A X is equivalent to ⇔ ∀ ⊆ (Cloτ1 (A)=X) (Cloτ2 (A)=X) A X.

The formulations and proofs of Propositions 6.7, 6.10, 6.17 can be easily carried over to the general case of functions which are topologically isomorphic to G.Inthe rest of this section we discuss a new characterization for such functions. We say that a subset A of R has the Baire property if there is an open set U ⊆ R such that the symmetric difference A∆U, A∆U =(A\U) ∪ (U\A), is of first category in R.

Theorem 6.20. The Cantor function G is isomorphic (as a map from [0, 1] into R) to a continuous monotone function f :[0, 1] → R with {0, 1}⊆([0, 1]\Lf ) if and only if the inverse image f −1(A) has the Baire property for every A ⊆ R.

Proof. Let f :[0, 1] → R be a function which is topologically isomorphic to G and let A ⊆ R. Then it follows from Proposition 6.1 that the set of constancy Lf is everywhere dense in [0, 1]. Reasoning as in the proof of Proposition 1.5 we can −1 prove that f (A) is the union of a Fσ subset of [0, 1] with some nowhere dense set. Since the collection of all subsets of [0, 1] having the Baire property forms σ-algebra [Oxt, Theorem 4.3] we obtain the Baire property for f −1(A). Suppose now that f :[0, 1] → R is a continuous monotone function, {0, 1}⊆ ([0, 1]\Lf ), but f is not isomorphic to G. Then using Proposition 6.1 we see that thereisanopeninterval(a, b) ⊆ ([0, 1]\Lf ). Let B be subset of (a, b) which does not have the Baire property. It is easy to see that B = f −1(f(B)). Thus, there exists a set A such that f −1(A) does not have a Baire property. 

Remark 6.21. The existence of a subset of the reals not having the Baire property depends on the axiom of choice. In fact, from the axiom of determinateness it follows that every A ⊆ R is Lebesgue measurable (Cf. Propositions 1.4, 1.5) and has the Baire property. See, for example, [Kan] and [Kl].

ÒØÓÖ ÙÒ Ø ÓÒ Ì 32

7 Dini’s

We recall the definition of the Dini derivatives. Let a real–valued function F be defined on a set A ⊆ R and let x0 be a point of A. Suppose that A contains some + half-open interval [x0,a). The upper right Dini derivative D F of F at x0 is defined by − + F (x) F (x0) D F (x0) = lim sup . x→x0 x − x0 x∈(x0,a) − We define the other three extreme unilateral derivatives D+F , D F and D−F sim- ilarly. (See [Br] for the properties of the Dini derivatives). Let C be the Cantor ternary set and G the Cantor ternary function. We first note that the upper Dini derivatives D+G and D−G are always +∞ or 0. We denote by M the set of points where G fails to have a finite or infinite derivative. It is clear that M ⊆ C. Since G satisfies Banach’s (T1) condition (see Proposition 1.10), the following conclusion holds.

Proposition 7.1. The set G(M) has a (linear Lebesgue) measure zero.

Proof. See Theorem 6.2 in Chapter IX of [Sa]. 

Hsc lg 2 Corollary 7.2. Let be a Hausdorff measure with sc = lg 3 . Then we have

Hsc (M)=0.

Proof. This follows from Propositions 7.1 and 4.5. 

Write W := {x ∈ (0, 1) : D−G(x)=D+G(x)=0}. In the next theorem we have collected some properties of Dini derivatives of G which follow from [Mor].

Proposition 7.3. 7.3.1. The set G(W ) is residual in [0, 1]. 7.3.2. For each λ ≥ 0 the set {x ∈ [0, 1) : D+G(x)=λ} has the power of the continuum.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 33

Remark 7.4. However, the set {x ∈ [0, 1) : D+G(x)=λ} is void whenever 0 <λ<∞.

Corollary 7.5. The set W ∩ C is residual in C.

Proof. The proof follows from 7.3.1 and Proposition 6.11. 

Remark 7.6. This corollary was formulated without a proof in [Eid].

Let x be a point of C. Let us denote by zx(n) the position of the nth zero in the ternary representation of x and by tx(n) the position of the nth digit two in this representation. The next theorem was proved by J.A.Eidswick in [Eid].

Theorem 7.7. Let x ∈ C◦ and let 3zx(n) (7.8) λx := lim inf . n→∞ 2zx(n+1) Then

(7.9) λx ≤ D+G(x) ≤ 2λx.

Furthermore, if lim zx(n +1)− zx(n)=∞,thenD+G(x)=λx. n→∞ Similarly, if 3tx(n) (7.10) µx := lim inf , n→∞ 2tx(n+1) then

(7.11) µx ≤ D−G(x) ≤ 2µx, with the equality on the right if lim tx(n +1)− tx(n)=∞. n→∞

◦  Corollary 7.12. If x ∈ C ,thenG (x)=∞ if and only if λx = µx = ∞.

Remark 7.13. This Corollary implies that the four Dini derivatives of G agree at ◦ x0 ∈ (0, 1) if and only if either x ∈ I or λx = µx = ∞.

◦ Corollary 7.14. If x ∈ C ,thenD−G(x)=D+G(x)=0if and only if

λx = µx =0.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 34

The following result is an improvement of 7.3.2.

Theorem 7.15. [Eid] For each µ ∈ [0, ∞] and λ ∈ [0, ∞],theset

{z ∈ C : D−G(z)=λ and D+G(z)=µ} has the power of the continuum.

Corollary 7.16. For each µ ∈ [0, ∞] and λ ∈ [0, ∞] there exists a nonempty, compact perfect subset in {z ∈ C : D−G(z)=λ and D+G(z)=µ}.

Proof. Since G is continuous on [0, 1], each of the Dini derivatives of G is in a Baire class two [Br, Chapter IV, Theorem 2.2]. Hence, the set {z ∈ C : D−G(z)=λ and D+G(z)=µ} is a Borel set. It is well-known that, if A is uncountable Borel set in a complete separable topological space X,thenA has a nonempty compact perfect subset [Kur, §39, I, Theorem 0]. 

Recall that M is the set of points at which G fails to have a finite or infinite lg 2 derivative and sc = lg 3 . The next result was proved by Richard Darst in [Dar2].

2 Theorem 7.17. The Hausdorff dimension of M is sc,orinotherwords

2 (7.18) dimH M = sc .

Remark 7.19. It is interesting to observe that the packing and box counting dimension are equal for M and C

(7.20) dimP M =dimB M =dimP C =dimB C.

The proof can be found in [DeLi], see also [Fal2] and [LXD] for the similar question.

The following proposition was published in [Dar2] with a short sketch of the proof.

Proposition 7.21. The equality

dimH(G(M)) = sc holds.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 35

An equality which is similar to (7.18) was established in [Dar3] for more general Cantor sets. Some interesting results about differentiability of the Cantor functions in the case of fat symmetric Cantor sets can be found in [Dar1] and [Den]. The Hausdorff dimension of M was found by J.Morris [Mo] in the case of nonsymmetric Cantor functions.

8 Lebesgue’s derivative

The so-called Lebesgue’s derivative or first symmetric derivative of a function f is defined as f(x +∆x) − f(x − ∆x) SDf(x) = lim . ∆x→0 2∆x It was noted by J.Uher that the Cantor function G has the following curious property.

Proposition 8.1. [Uh] G has infinite Lebesgue’s derivative, SDG(x)=∞, at every point x ∈ C\{0, 1},andSDG(x)=0for all x ∈ I◦.

Remark 8.2. It was shown by Buczolich and Laczkovich [BuLa] that there is no symmetrically differentiable function whose Lebesgue’s derivative assumes just two finite values. Recall also that if f is continuous and has a derivative everywhere (finite or infinite), then the range of f  must be a connected set.

From the results in [BuLa] we obtain

Theorem 8.3. For all sufficiently large b the inequality G(x + b∆x) − G(x − b∆x) (8.4) lim sup

Remark 8.5. It was shown in [BuLa] that (8.4) is true for all b ≥ 81 and false if b = 4. Moreover, as it follows from [BuLa] (see, also [Thom, Theorem 7.32]) that Theorem 8.3 implies Proposition 8.1.

9H¨older continuity, distortion of Hausdorff di- mension, sc-densities

lg 2 Proposition 9.1. The function G satisfies a H¨older condition of order sc = lg 3 , the H¨older coefficient being not greater than 1. In other words (9.2) |G(x) − G(y)|≤|x − y|sc

ÒØÓÖ ÙÒ Ø ÓÒ Ì 36

for all x, y ∈ [0, 1]. The constants sc and 1 are the best possible in the sense that the inequality |G(x) − G(y)|≤a|x − y|β does not hold for all x, y in [0, 1] if either β>sc or β = sc and a<1.

Proof. Using Lemma from [GorKu] we obtain the inequality 1 sc (9.3) t ≤ G(t) ≤ tsc 2 for all t ∈ [0, 1]. Since G is its first modulus of continuity (see Proposition 2.2), the last inequality implies (9.2). Moreover it follows from (9.3) that sc is a sharp H¨older exponent in (9.2). By a simple calculation we have from formula (0.2) that G(3−n)=3−nsc for all positive integers n.Thus,theH¨older coefficient 1 is the best possible in (9.2). 

Remark 9.4. It was first proved in [HiTa] that G satisfies the H¨older condition with the exponent sc and coefficient 2. R.E.Gilman erroneously claimed (without proof and for a more general case) that both constants sc and 2 are the best possible [Gil]. In [GorKu] a similar to (9.3) inequality does not link with the H¨older condition for 1 sc G. Observe also that 2 is the sharp constant for (9.3) as well as the constant 1. See Figure 4.

Proposition 9.5. Let x be an arbitrary point of C. Then we have lg |G(x) − G(y)| lim inf = s . y→x | − | c y∈C lg x y

Proof. It suffices to show that lg |G(x) − G(y)| lim inf ≤ s . y→x | − | c y∈C lg x y

| − | 2 →∞  Choosing x y = 3n for n we obtain the result.

Let x be a point of the Cantor ternary set C.Thenx has a triadic representation ∞ α (9.6) x = m 3m m=1

ÒØÓÖ ÙÒ Ø ÓÒ Ì 37

∈{ } {R }∞ where αm 0, 2 . Define a sequence x(n) n=1 by the rule inf{m − n : αm = αn,m>n} if ∃m>n: αm = αn Rx(n):= 0, if ∀m>n: αm = αn,

i.e., Rx(n)=1 ⇐⇒ (αn = αn+1);

Rx(n)=2 ⇐⇒ (αn = αn+1)&(αn+1 = αn+2) and so on.

sc 1 sc Figure 4: The graph of G lies between two curves y1(x)=x and y2(x)= 2 x .

Theorem 9.7. [DMRV] For x ∈ C we have

lg |G(x) − G(y)| lim = s y→x | − | c y∈C lg x y if and only if R (n) lim x =0. n→∞ n

This theorem and a general criterion of constancy of linear distortion of Hausdorff dimension (see [DMRV]) imply the following result.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 38

Theorem 9.8. There exists a set M ⊆ C such that Hsc (M)=1and 1 dimH(G(A)) = dimH(A) sc for every A ⊆ M.

Remark 9.9. In the last theorem we can take M as

−1 M = G (SN2) where SN2 is the set of all numbers from [0, 1] which are simply normal to base 2. See, for example, [Ni] for the definition and properties of simply normal numbers.

∗sc sc Let Θ (t)andΘ∗ (t) denote the upper and lower sc-densities of µC at t ∈ R, i.e.,

|Gˆ(t + x) − Gˆ(t − x)| Θ∗sc (t) := lim sup x→0 |2x|sc x=0 | ˆ − ˆ − | sc G(t + x) G(t x) Θ∗ (t) := lim inf . x→0 | |sc x=0 2x It can be shown that ∗sc sc Θ (t) < Θ∗ (t) for every t ∈ C, i.e., sc-density for measure µC does not exist at any t at the Cantor set C. However, the logarithmic averages sc-density does exist for many fracfals including C.

Theorem 9.10. For Hsc x ∈ C the logarithmic average density 1 T Gˆ(x + e−t) − Gˆ(x − e−t) Asc (x) := lim dt →∞ | −t|sc T T 0 2e exists with 1 Asc (x)= |x − y|−scdG(x)dG(y)=0, 6234... 2sc log 2 | − |≥ 1 x y 3

For the proof see [Fal1, Theorem 6.6] and also [BeFi].

ÒØÓÖ ÙÒ Ø ÓÒ Ì 39

The interesting results about upper and lower sc-densities of µC can be found in [FHW]. For x with representation (9.6) we defineτ ˆ(x)andτ(x)as ∞ −m τˆ(x) := lim inf αm+k3 , k→∞ m=1 τ(x):=min(ˆτ(x), τˆ(1 − x)).

Theorem 9.11. [FHW] 9.11.1 For all x ∈ C,

sc −sc Θ∗ (x)=(4− 6τ(x)) . 9.11.2 For all x ∈ C,  −  2 sc if x ∈ C1 ∗ − Θ sc = 2 sc  (2 + τ(x))−sc if x ∈ C◦. 3 9.11.3 For all x ∈ C◦,

∗sc −1/sc sc −1/sc 9(Θ (x)) +(Θ∗ (x)) =16.

∗sc sc −sc 9.11.4 sup{Θ (x) − Θ∗ (x):x ∈ C} =4 , where the supremum can be attained at {x ∈ C◦ : τ(x)=0},and

− − 3 sc 5 sc inf{Θ∗sc (x) − Θsc : x ∈ C} = − , ∗ 2 2 ∈ ◦ 1 wheretheinfimumcanbeattainedat x C : τ(x)= 4 . 9.11.5 For Hsc -almost all x ∈ C,

sc −sc ∗sc −sc Θ∗ (x)=4 , Θ (x)=2· 4 .

Remark 9.12. The general result similar to 9.11.5 can be found in the work Salli [Sal].

Acknowledgements. The research was partially supported by grants from the University of Helsinki and by grant 01.07/00241 of Scientific Fund of Fundamental Investigations of Ukraine.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 40

References

IJKLMNO¸ þ           ¹

[Al] CºDºüFGH

PNIHOJ¸ 8QJRHJ8¸ ½*%%º  º [P.S.Aleksandrov, ”Introdaction to Set The- ory and General Topology”. Izdat. ”Nauka”, Moskow, 1977.]

[Bar] N.Bary, Memoire sur la representation finie des fonctions continues. Math.Ann., 103, (1930), 185–248 and 598–653.

[BaMe] N.Bary et D.Menchoff, Sur l’integrale de Lebesgue-Stieltjes et les fonctions absolumant continues. Ann. Math. Pura e Appl., 5, (1928), 19–54.

[BeFi] T.Bedford and A.M.Fisher, Analogues of the Lebesgue density theorem for sets of reals and integer. Proc. London Math. Soc., 64, (1992), 95– 124.

[Br] A.M.Bruckner, ”Differentiation of Real Functions”. Lecture Notes in Math- ematics 659, Springer–Verlag, Berlin, 1978.

[BrGo] A.M.Bruckner and J.L.Goffman, Differentiability through change of vari- ables, Proc. Amer. Math. Soc., 61, (1976), 235–241.

[BrLe] A.M.Bruckner and J.L.Leonard, On differentiable functions having an ev- erywhere dense set of intervals of constancy, Cand. Math. Bull., 8 (1965), 73–76.

[BSU] Yu.M. Berezansky, Z.G.Sheftel, G.F.Us, ”Functional analysis. Volume II”. Birkh¨auser Veglar, Basel, 1996.

[BuLa] Z.Buczolich and M.Laczkovich, Concentrated Borel measures. Acta. Math. Hungar., 57, (349–362), 1991.

[Can] G.Cantor, De la puissance des ansembles parfaits de points.Acta Math. 4, (1884), 381–392. Reprinted in G.Cantor, Gezammelte Abhandlungen math- ematischen und philosophischen Inhalts. E.Zermelo (ed.), Springer-Verlag, New York, 1980.

[Ch] D.R.Chailice, A characterization of the Cantor function. Amer. Math. Monthly, 98, (1991), N3, 255–258.

[Cop] W.A.Coppel, An interesting Cantor set. Amer. Math. Monthly, 90, (1983), N7, 456–460.

[Dar1] R.B.Darst, Some Cantor sets and Cantor functions. Mathematics Magazine, 45, (1972), 2–7.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 41

[Dar2] R.B.Darst, The Hausdorff dimension of the nondifferentiability set of the Cantor function is [ln 2/ ln 3]2. Proc. Amer. Math. Soc., 119, N1, (1993), 105–108.

[Dar3] R.B.Darst, Hausdorff dimension of sets of non-differentiability points of Cantor functions. Math. Proc. Camb. Phil. Soc., 117, (1995), 185–191.

[DeLi] F. M. Dekking and Wenxia Li, How smoth is a devil’s staircase? , 11, 1, (2003), 101–107.

[Den] T.P.Dence, Differentiable points of the generalized Cantor function. Rocky Mountain J. Math., 9, 2, (1979), 239–249.

[Dob] J.Dobo˘s, The standard Cantor function is subadditive. Proc. Amer. Math. Soc., 124, (1996), N11, 3425–3426.

[DMRV] O.Dovgoshey, O.Martio, V.Ryazanov, M.Vuorinen, Linear distortion of Hausdorff dimension and Cantor’s function. to appear in Real Anal. Ex- change.

[DoMa] O.Dovgoshey and O.Martio, The evaluation and Lebesgue decom- position. Preprint 379, (2004), University of Helsinki.

[Eid] J.A.Eidswick, A characterization of the nondifferentiability set of the Cantor function. Proc. Amer. Math. Soc., 42, 1, (1974), 214–217.

[Ev] G.C.Evans, Calculation of moments for a Cantor–Vitali function. Amer. Math. Monthly, 64, 8, (1957), 22–27.

[EvGa] Lawrence C. Evans and Ronald F. Gariepy ”Measure Theory and Fine Prop- erties of Functions”. CRC Press, Boca Raton, 1992.

[Fal1] Kenneth Falconer, ”Techniques in Fractal Geometry”. John Wiley, New York, 1997.

[Fal2] Kenneth J. Falconer, One-side miltifractal analysis and points of non- differentiability of devil’s staircases. Math. Proc. Cambridge Philos.Soc., 136, 1, (2004), 167–174.

[FHW] De-Jun Feng, Su Hua, Zhi-Ying Wen, The Pointwise Densites of the Cantor Measure. J.Math. Anal. Appl. 250, (2000), 692–705.

[Fis] H.-J. Fisher, On the paper: ”Asymptotics for the moments of singular dis- tributions”. J. Approx. Theory, 82, 3, (1995), 362–374.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 42

[Fle] Julian F.Fleron, A note on the history of the Cantor set and Cantor func- tion. Mathematics Magazine, 67, (1994), 136–140.

[Gil] R.E.Gilman, A class of functions continuous but not absolutely continuous. Annals of Math. (2), 33, (1932), 433–442.

[GNW] Casper Goffman, Togo Nishiura, Daniel Waterman, ”Homeomorphisms in Analysis”. Mathematical Survey and Monographs, Vol. 54. American Math- ematical Society, Providence, R.I., 1998.

[GoWi] William Goh and Jet Wimp, Asymptotics for the moments of singular dis- tributions. J. Approx. Theory, 74, 3, (1993), 301–334.

[GorKu]E.A.GorinandB.N.Kukushkin,Integrals related to the Cantor function. St. Petersburg Math. J., 15, 3, (2004), 449–468.

[Haw] T.Hawkins, ”Lebesgue’s Theory of Integration”. Chelsea Publishing Co., Madison, WI, 1975.

[HiTa] E.Hille and J.D.Tamarkin, Remarks on known example of a monotone con- tinuous function. Amer. Math. Monthly, 36, (1929), 255–264.

[HuLa] Tian-You Hu and Ka-Sing Lau, Fourier asymptotics of Cantor type measures at infinity. Proc. Amer. Math. Soc., 130, 9, (2002), 2711–2717.

[Hut] J.E.Hutchinson, Fractals and self–similarity. Indiana Univ. Math. J., 30, 5,

(1981), 713–747.

1ÃÒ2 þºÿº VJKNOGW¸ ü  þ   ü     º

8QJRHJ8¸ ½*(!º PNIHOJ¸ [V.G. Kanoveˇı, ”The Axiom of Coice and the Ax- iom of Determinateness”. Izdat. ”Nauka”, Moskow, 1984.]

[Kl] Eugene M. Kleinberg, ”Infinitary Combinatorics and the Axiom of Determi- nateness”. Lecture Notes in Mathematics, 612, Springer-Verlag, Berlin–New York, 1977.

[Kur] K.Kuratowski, ”Topology, Volume 1”. Academic Press, New York and Lon- don, 1966.

[LaWa] K. S. Lau and J. Wang, Mean quadratic variations and Fourier asymptotics of self-similar measures. Monat. Math., 115, (1993), 99–132.

[LXD] Wenxia Li, Dongmei Xiao and F. M. Dekking, Non-differentiability of devil’s staircaces and dimentios of subsets of Moran sets. Math. Proc. Cambridge Philos. Soc., 133, 2, (2002), 345–355.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 43

[Mo] J.Morris, The Hausdorff dimension of the nondifferentiability set of a non- symmetric Cantor function. Rocky Mountain J. Math., 32, 1 (2002), 357– 370.

[Mor] A.P.Morse, Dini derivatives of continuous functions. Proc. Amer. Math. Soc., 5, N1, (1954), 126–130.

[Ni] Ivan Niven, ”Irrational numbers”. Wiley and Sons, Inc., New York, 1956.

[Oxt] J.Oxtoby, ”Measure and Category”. Springer-Verlag, New York, 1971.

[Pel] M.J.Pelling, Formular for the arc-length of curve in Rn. Amer. Math.

Montly, 84, 6 (1977), 465–467.

NYFZK¸ [ºýº]RHI¸   !" #     º ÿ   %

[RoFu] þºüºX

PNIHOJ¸ 8QJRHJ8¸ ½*%%º º [V.A.Rohlin, D.B.Fuks, ”A Beginnig Course in Topology: Geometrical Chapters”. Izdat. ”Nauka”, Moskow, 1977.]

[Sa] S.Saks, ”Theory of the Integral”. Second revised edition, Dover Publications, INC., 1964.

[Sal] A.Salli , Upper density properties of Hausdorff measures on fractals. Ann. Acad. Sci. Fenn. Ser AI Math. Dissertationes, 55, 1985.

[Str1] R. Strichartz, Fourier asymptotics of fractal measures. J. Functional Anal., 89, (1990), 154–181.

[Str2] R. Strichartz, Self-similar measure and their Fourier transform I. Indiana University Math. J., 39, (1990), 797–817.

[Str3] R. Strichartz, Self-similar measure and their Fourier transform II. Trans. Amer. Math. Soc., 336, N 1, (1990), 335–361.

[Thom] B.C.Thomson, ”Symmetric Properties of Real Function”. Marcel Dekker, New York, 1994.

[Tim] A.F.Timan, ”Theory of Approximation of Functions of a Real Variable”. In- ternational Series of Monographs in Pure and Applied Mathematics, Vol.34. Pergamon Press, Oxford, 1963.

[Uh] J.Uher, Symmetric continuity implies continuity. Rel. Anal. Exchange, 13, (1987/88), 36–38.

[WW] N. Wiener and A. Wintner, On singular distributions. J. Math. Phy., 17, (1939), 233–246.

ÒØÓÖ ÙÒ Ø ÓÒ Ì 44

[WY] Norbert Wiener and R. C. Young, The total variation of g(x + h) − g(x). Trans. Amer. Math. Soc., 35, N 1, (1933), 327–340.

Oleksiy Dovgoshey and Vladimir Ryazanov Institute of Applied Mathematics and Mechanics, NAS of Ukraine, 74 Roze Luxemburg str., Donetsk, 83114 UKRAINE email: [email protected] [email protected]

Olli Martio Department of Mathematics and Statistics, P.O. Box 68 (Gustaf H¨allstr¨omin katu 2b) FIN – 00014 University of Helsinki FINLAND email: [email protected]

Matti Vuorinen Department of Mathematics, FIN – 20014 University of Turku FINLAND email: [email protected]