The Cantor Function
The Cantor Function
2
" G / 0 ' 12 3 ((
0 4/ 12 5 / /4
6 "
G + 4' 7
1+/ $2 "
6' 162 " /
+ ) 8)9 ' :
8 ;< 5 .= *!> 6 G
)
1?)2 1" &0 % 2 " G /
' /
/4 G .
/ /
@ 0 G C )
x ∈ [0, 1] x
∞
anx
(0.1) x= , anx ∈ {0, 1, 2}.
n=1
3n
1
Nx −1
1 anx
(0.2) G(x) := + .
2 Nx 2 n=1 2n
A ; > x /
" C [0, 1] / ; >
x ∈ C ;anx ∈ {0, 2}> ;>
4
1 anx
∞
(0.3) G(x) = .
2 n=1 2n
& C 1 C
Figure 1: The graph of the Cantor function G. This graph is sometimes called
”Devil’s staircase”.
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 ◦ .
4
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, and x 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 measure 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
m1 (Ck ) = m1 (Ck−1)
3
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.
In general a continuous function 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.
5
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 . Let E 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) .
If A is an arbitrary subset of R, then it is easy to see that f0−1 (A) is a Fσ subset of
[a, b] and
f1−1 (A) ⊆ [a, b]\Lf .
Since f −1 (A) = f0−1 (A)∪f1−1 (A), the equality m1 ([a, b]\Lf ) = 0 implies that f −1 (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
Suppose that (1.6) does not hold. For every B ⊆ R with an outer measure m∗1 (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.
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
Since a restriction G|C ◦ is one-to-one and G(I ◦ ) is a countable set, 1.1.2 implies
Proposition 1.11. There are absolutely continuous functions f1 , ..., f4 such that
G = f1 ◦ f2 + f3 ◦ f4 ,
for all x, y ∈ R.
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 Dobos̆ [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, ∞) with ω(0) = 0 is to show that the function δ −1 ω(δ) is de-
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.
Proposition 2.3. The function x−1 G(x) is locally monotone at point x0 if and only
if x0 ∈ I ◦ .
8
Propositions 2.3 and 2.4 are particular cases of some general statements.
Proof. By definition Lf is relatively open in [a, b]. Hence [a, b]\Lf is a compact subset
of R. If p 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.
Denote by lz and ly the straight lines which pass through the points (z0 , f (z0 )),
z+y
2
, f z+y
2
and (y0 , f (y0 )), z+y 2
, f z+y
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 derivative ϕ (x) at
every x ∈ (a, b). If
Figure 2: The set of points of constancy is the same as the set of points of convexity.
f (x + ∆x) − f (x − ∆x)
SDf (x) = lim
∆x→0 2∆x
provided that the limit exists, and write
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 <x<
2 3 3
1 + 1 G(3x − 2) if
2
≤ x ≤ 1.
2 2 3
If ψ0 ∈ M[0, 1], then the sequence {ψn }∞
n=0 converges uniformly to G.
11
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 Dobos̆’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) ϕ0 (x) := x, ϕ1 (x) := x + .
3 3 3
Proposition 3.5. The Cantor set C is the unique nonempty compact subset of R
for which
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.
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].
(3.13) F (1 − x) = 1 − F (x)
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.
13
holds, then G = F .
Figure 3: The continuous function F which satisfies ”the two Cantor function equa-
tions” (3.9) and (3.13) but F (x) = 12 (12x − 5)2 on 13 , 12 .
Lemma 3.19. Let f : [a, b] → R be a continuous function and let p ∈ (1, ∞).
Suppose that the graph of f is symmetric with respect to the point a+b
2
; f a+b
2
.
Then the inequality
b p
1 a + b
(3.20) |f (x)| dx ≥ f
p
|b − a| a 2
Proof. We may assume without loss of generality that a = −1, b = 1. Now for
f (0) = 0 inequality (3.20) is trivial. Hence replacing f with −f , if necessary, we
may assume that f (0) < 0. Write
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
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 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 I 0 = ∪ J is a dense subset of [0, 1] and F = G as required.
J∈I
Define a subspace L2C [0, 1] of the Hilbert space L2 [0, 1] by the rule: f ∈ L2C [0, 1]
if f ∈ L2 [0, 1] and for every J ∈ I there is a constant CJ such that
|f (x) − CJ |2 dx = 0.
J
lg 2
(4.1) sc := .
lg 3
We let Hsc denote the sc -dimensional Hausdorff measure in R. See [Fal1] for
properties of Hausdorff measures.
17
Proposition 4.2. There is the unique Borel regular probability measure µ such that
1 1
(4.3) µ(A) = µ(ϕ−1 −1
0 (A)) + µ(ϕ1 (A))
2 2
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.
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
measure which fulfils (4.3). Since C 1 is countable (see formula (0.5)), m1 (G(C 1 )) =
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) is a Borel set
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
A0 := A ∩ 0, , A1 := A ∩ , 1 .
3 3
18
Since ϕ−1 −1
0 (x) = 3x and ϕ1 (x) = 3 x −
2
3
, we have
µ(ϕ−1 −1
i (A)) = µ(ϕi (Ai ))
2G(x) = G(ϕ−1
0 (x)),
for x ∈ 0, 13 . Hence,
1 1 1
µ(ϕ−1 −1
0 (A0 )) = m1 (G(ϕ0 (A0 ))) = m1 (2G(A0 )) = m1 (G(A0 )).
2 2 2
Similarly, (4.7) implies that
2
2G x − = G(ϕ−1
1 (x))
3
2
for x ∈ 3
, 1 , and we obtain
1 1
µ(ϕ−1 −1
1 (A1 )) = m1 (G(ϕ1 (A1 ))) =
2 2
1 2 2
= m1 2G A1 − = m1 G A1 − .
2 3 3
Now using (4.8) we get
2 1
m1 G A1 − = m1 − + G(A1 ) = m(G(A1 )).
3 2
The set G(A1 ) ∩ G(A0 ) is empty or contains only the point 12 . Hence we obtain
This description of the Cantor function enables us to suggest a method for the
proof of the following proposition. Write
Proposition 4.10. [HiTa] Let V ar(Ĝh ) be a total variation of Ĝh . Then we have
sup V ar(Ĝh ) = 2
0≤h≤δ
This holds because sets C ∩ (C ± 3−n ) have finite numbers of elements for all
positive integer n.
Really, if F is absolutely continuous on [a, b], then F exists a.e. in [a, b], F ∈ L1 [a, b],
and
b
V ar(F ) = |F (x)| dx.
a
Approximating of F by continuous functions, for which the property is obvious, we
obtain (4.12).
holds.
The theorem is an immediate adaptation of the result that was proved by Norbert
Wiener and R. C. Young in [WY].
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
Mn = xn G(x)dx.
0
1 1 n n−1
(5.2) 2Mn = + n+1 2n−k Mk ,
n+1 3 − 1 k=0 k
1 n−1
n k+1 n
(5.3) (1 + (−1) )Mn = + (−1) Mk
n + 1 k=0 k
n
for all positive integers n where k
are the binomial coefficients.
n−1 n
(5.6) 2mn = (−1)k mk ,
k
k=0
n−1
k n+1
(5.7) (n + 1)mn = (−1) mk
k=0
k
Lk = {aik : i = 1, ..., 2k }.
l
mk = lim (ξi)n [G(xj+1 ) − G(xj )]
l→∞
j=1
where 0 = x1 < ... < xl < xl+1 = 1 is any subdivision of [0, 1] with maxj (xj+1 −xj ) →
0 as l → ∞ and ξj ∈ [xj , xj+1 ]. Subdivide [0, 1] into l = 3k equal intervals [xj , xj+1]
and take ξj = xj . Then
1 n
mn = lim k x .
k→∞ 2
x∈L k
1 1 1 n 1 1 1
= · n lim k−1 x + · n lim k−1 (x + 2)n
2 3 k→∞ 2 x∈L
2 3 k→∞ 2 x∈L
k−1 k−1
1 1
n
1 1 n
= · n mn + · n mp 2n−p ,
2 3 2 3 p=0 p
Suppose that n is even, then it follows from (5.3) and (5.9) that
n−1 n
2(1 − mn+1 ) = 2(n + 1)Mn = 1 + (n + 1) (−1)k+1 Mk
k=0
k
n + 1 n
n−1
=1+ (−1)k+1 (1 − mk+1 ).
k=0
k+1 k
n+1 n n+1
Hence from k+1 k
= k+1
and (1 − m0 ) = 0 we get
n−1
k+1 n+1
2(1 − mn+1 ) = 1 + (−1) (1 − mk+1 )
k=0
k+1
n
n
k n+1 k n+1
=1+ (−1) (1 − mk ) = 1 + (−1)
k k
k=1 k=0
n
n+1
− (−1)k mk .
k=0
k
Observe that
n
k n+1 n+1 n+1
(−1) = (1 − 1) − (−1)n+1 = 1
k=0
k n+1
The last formula implies (5.6). The proof of (5.7) is analogous to that of (5.6).
1 τn τn−1 τ0
τn+1 = + + ... + .
2(3n+1 − 1) 1! 2! (n + 1)!
23
where
1 1 −ν(m)
+∞
H(x) := 2 (1 − 21−ν(m) )ζ(ν(m))Γ((ν(m))x1−ν(m)
2πi m∈Z m
m=0
and
lg 2
sc = .
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
λ
(5.13) I(λ) := (G(x)) dx, 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
λ
(3 · 2 − 1)I(λ) = 1 + (1 + G(x))λ dx,
0
1 n 2k−1 − 1 B(k)
I(n) = −
n + 1 k≤n k 3 · 2 k−1 −1 n−k+1
where the primes mean that summation is over even positive k and B(k) are Bernoulli
numbers
5.14.3 For λ → ∞, we have
+∞
2 2πin 2πin
Φ(z) = Γ sc + ζ sc + exp(−2πinz).
−∞
3 lg 2 lg 2 lg 2
Proof. Write 1
Φ(a) := eax dG(x).
0
1
Observe that the function Φ, Φ(a) = eax dG(x), is the exponential generating
0
function of the moment sequence {mn }.
1 The
2πinx
next result follows from (5.16) and shows that Fourier coefficients µ̂n , µ̂n :=
0
e dµC (x), of µC do not tend to zero as n → ∞.
∞
2π
Remark 5.19. The infinite product cos 3ν
converges absolutely because
ν=1
2π π π2
2
|1 − cos | ≤ 2 sin < 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].
26
This theorem follows from the results of M.J.Pelling [Pel]. See also [DoMa].
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
0 = lim
x→0
f (x) = 1 − lim
x→1
f (x)
x∈A x∈A
and
lim f (x) =
x→t
lim f (x)
x→t
x∈A∩[0,t] x∈A∩[t,1]
f˜(t) := lim
x→t
f (x), t ∈ [0, 1].
x∈A
27
Obviously, f˜ is strictly increasing and f˜(t) = f (t) for each t ∈ A. Suppose that
for some x0 ∈ (0, 1). Since B is a dense subset of [0, 1], there is b0 ∈ B such that
Ψ◦ : gI ◦ → GI ◦ , Ψ◦ (g(J)) = G(J), J ∈ I
28
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 ◦ . Since I ◦ is a everywhere dense subset of [0, 1], the last equality
implies that
G = ψ ◦ g ◦ η.
Thus g and G are topologically isomorphic.
Thus, we obtain.
A ⊇ (a, b) ∩ C.
Proposition 6.11. If B is a residual (first category) subset of [0, 1], then G−1
C (B)
is residual (first category) in C.
and
Moreover, we have
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
∞
◦
W ∩C ⊇ G−1
C (On ),
n=1
IntC (C\G−1
C (On0 )) = ∅
Remark 6.16. In a special case this proposition was mentioned without proof in
the work by J.A.Eidswick [Eid].
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
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).
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. In the
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.
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].
32
7 Dini’s derivatives
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)
Proposition 7.1. The set G(M) has a (linear Lebesgue) measure zero.
Hsc (M) = 0.
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.
33
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].
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 = ∞.
Theorem 7.15. [Eid] For each µ ∈ [0, ∞] and λ ∈ [0, ∞], the set
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, then A 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
derivative and sc = lg 2
lg 3
. The next result was proved by Richard Darst in [Dar2].
Remark 7.19. It is interesting to observe that the packing and box counting
dimension are equal for M and 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.
dimH (G(M)) = sc
holds.
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.
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.
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.
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.
Remark 9.4. It was first proved in [HiTa] that G satisfies the Hölder 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 Hölder condition for
s
G. Observe also that 12 c is the sharp constant for (9.3) as well as the constant 1.
See Figure 4.
Let x be a point of the Cantor ternary set C. Then x has a triadic representation
∞
αm
(9.6) x=
m=1
3m
37
1 sc
Figure 4: The graph of G lies between two curves y1 (x) = xsc and y2 (x) = 2
x .
lg |G(x) − G(y)|
lim = sc
y→x
y∈C
lg |x − y|
if and only if
Rx (n)
lim = 0.
n→∞ n
Theorem 9.8. There exists a set M ⊆ C such that Hsc (M) = 1 and
1
dimH (G(A)) = dimH (A)
sc
for every A ⊆ M.
M = G−1 (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.
Let Θ∗sc (t) and Θs∗c (t) denote the upper and lower sc -densities of µC at t ∈ R,
i.e.,
Theorem 9.10. For Hsc almost all x ∈ C the logarithmic average density
T
sc 1 Ĝ(x + e−t ) − Ĝ(x − e−t )
A (x) := lim dt
T →∞ T 0 |2e−t |sc
exists with
sc 1
A (x) = sc |x − y|−sc dG(x)dG(y) = 0, 6234...
2 log 2
|x−y|≥ 31
For the proof see [Fal1, Theorem 6.6] and also [BeFi].
39
The interesting results about upper and lower sc -densities of µC can be found in
[FHW]. For x with representation (9.6) we define τ̂ (x) and τ (x) as
∞
τ̂ (x) := lim inf αm+k 3−m ,
k→∞
m=1
Remark 9.12. The general result similar to 9.11.5 can be found in the work Salli
[Sal].
References
[Al] CDEFGHIJKLMNO
PNIHOJ 8QJRHJ8 *%% [P.S.Aleksandrov, ”Introdaction to Set The-
ory and General Topology”. Izdat. ”Nauka”, Moskow, 1977.]
[BeFi] T.Bedford and A.M.Fisher, Analogues of the Lebesgue density theorem for
fractal sets of reals and integer. Proc. London Math. Soc., 64, (1992), 95–
124.
[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.
41
[DeLi] F. M. Dekking and Wenxia Li, How smoth is a devil’s staircase? Fractals,
11, 1, (2003), 101–107.
[Dob] J.Dobos̆, The standard Cantor function is subadditive. Proc. Amer. Math.
Soc., 124, (1996), N11, 3425–3426.
[DoMa] O.Dovgoshey and O.Martio, The arc length evaluation and Lebesgue decom-
position. Preprint 379, (2004), University of Helsinki.
[EvGa] Lawrence C. Evans and Ronald F. Gariepy ”Measure Theory and Fine Prop-
erties of Functions”. CRC Press, Boca Raton, 1992.
[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.
42
[Fle] Julian F.Fleron, A note on the history of the Cantor set and Cantor func-
tion. Mathematics Magazine, 67, (1994), 136–140.
[GoWi] William Goh and Jet Wimp, Asymptotics for the moments of singular dis-
tributions. J. Approx. Theory, 74, 3, (1993), 301–334.
[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.
1S2 TU VJKNOGW
PNIHOJ 8QJRHJ8 *(! [V.G. Kanoveı̌, ”The Axiom of Coice and the Ax-
iom of Determinateness”. Izdat. ”Nauka”, Moskow, 1984.]
[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.
[Ni] Ivan Niven, ”Irrational numbers”. Wiley and Sons, Inc., New York, 1956.
[Sa] S.Saks, ”Theory of the Integral”. Second revised edition, Dover Publications,
INC., 1964.
[Str3] R. Strichartz, Self-similar measure and their Fourier transform II. Trans.
Amer. Math. Soc., 336, N 1, (1990), 335–361.
[Uh] J.Uher, Symmetric continuity implies continuity. Rel. Anal. Exchange, 13,
(1987/88), 36–38.
[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.
Olli Martio
Department of Mathematics and Statistics,
P.O. Box 68 (Gustaf Hällströmin katu 2b)
FIN – 00014 University of Helsinki
FINLAND
email: martio@cc.helsinki.fi
Matti Vuorinen
Department of Mathematics,
FIN – 20014 University of Turku
FINLAND
email: vuorinen@csc.fi