LOCAL GENERALIZATIONS OF THE ON THE REAL LINE

DIMITER PRODANOV

Abstract. From physical perspective, derivatives can be viewed as mathe- matical idealizations of the linear growth. The linear growth condition has special properties, which make it preferred. The manuscript investigates the general properties of the local generalizations of derivatives assuming the usual topology of the real line. The concept of is generalized in terms of the class of the modulus of continuity of the primitive . This definition is suitable for applications involving continuous but possibly non-absolutely continuous functions of a real variable. The main application of the approach is the generalization of the Lebesgue monotone differentiation theorem. On the second place, the conditions of continuity of generalized derivative are also demonstrated.

Keywords: non-differentiable functions ; singular functions ; moduli of continu- ity; H¨older classes 1) Environment, Health and Safety, IMEC, Belgium 2) MMSDP, IICT, BAS, Bulgaria Correspondence Address: IMEC vzw, Kapeldreef 75, 3001 Leuven, Belgium

1. Introduction Since the time of Newton, it is accepted that celestial mechanics and physical phenomena are, by and large, described by smooth and continuous functions. The second law of Newton demands that the velocity is a differentiable function of time. This ensures mathematical modelling in terms of differential equations, and hence (almost everywhere) differentiable functions. Ampere even tried to prove that all functions are almost everywhere differentiable. Now we know that this attempt was doomed to fail. Various non-differentiable functions have been constructed in the XIXth century and regarded with a mixture of wonder and horror. The interest in fractal and non-differentiable functions was rekindled with the works of Mandelbrot in fractals arXiv:1909.02728v2 [math.CA] 23 Sep 2020 [17]. For example, Fonf et al. (1999) have established that there is a closed, infinite- dimensional subspace of C[0, 1] consisting of nowhere differentiable functions[11]. This existence result was extended with constructive proofs in [12, 4]. Scientific developments in the last 50 years indicate that the use of non-differentiable functions can not be avoided when modelling nature. For instance, it is easy to es- tablish that stochastic paths of the classical Wiener process are non-differentiable. Applications of this process are ubiquitous in physics, biology and economy. On a second place, some physical theories also consider non-differentiable functions. The stochastic interpretation of Quantum Mechanics, introduced by Nelson [19], as- sumes a reversible sub-quantum Brownian motion (i.e. reversible Wiener process), having non-differentiable trajectories. In a closely related manner, almost all, in 1 2 DIMITER PRODANOV the measure sense, paths in the formulation of the Feynman path-integral are non- differentiable [9]. The deterministic approach of scale relativity theory, introduced by Nottale [21] also assumes non-differentiability of the fundamental space-time manifold. The Ornstein–Uhlenbeck (OU) process was introduced in the kinetic theory of gasses [27]. In this process, the particle velocities are non-differentiable. The OU process arises as the scaling limit of the Ehrenfest urn model which de- scribes the diffusion of particles through a permeable membrane. Purely mathematically, the derivatives can be generalized in several ways. Deriva- tives can be defined in the usual way as limits of difference quotients on the accu- mulation sets of points [18, ch 3, p. 105]. This approach can be applied also to functions defined on fractal sets [29]. On the other hand, the question of continuity on intervals of so-defined functions requires further specification. If continuity is perceived as an essential property such generalization leads to var- ious integro-differential operators. The best known examples here are the Riemann- Liouville and Caputo operators. However, such operators lead to non-local (inter- val) functions. Application of a subsequent limiting localization operation can lead to a local operator. An example of this is the local fractional derivative introduced by Kolwankar and Gangal [15]: 1 d Z x f(t) DKG f(x) := lim dt x a − β → Γ(1 β) dx a (x − t) However, such localization can only lead to trivially continuous functions – that is – the result of the localization is zero where the derivative function is continuous [23]. Point-wise, the derivatives can be generalized by formal ”fractionalization” – i.e. by the substitution  → β leading to

β f(x + ) − f(x) υ+f (x) := lim  0 β →  The quantity in this expression is called fractional velocity. Such an approach has been considered for the first time by the mathematicians Paul du Bois-Reymond and Georg Faber in their studies of the point-wise differentiability of functions [7, 8]. In the late XXth century, the physicist Guy Cherbit introduced the same quantity under the name fractional velocity by analogy with the Hausdorff dimension as a tool to study the fractal phenomena and physical processes for which instantaneous velocity was not well-defined [6]. The properties of fractional velocity have been extensively studied in [1] and [22]. The special choice of the function β can be justified from the theory of the fractional calculus as the limit of the regularized Riemann-Liouville differ-integral (see above). As can be expected, the overlap of the definitions of the Cherebit’s fractional velocity and the Kolwankar-Gangal local fractional derivative is not complete. The precise equivalence conditions have been established elsewhere [2, 23]. Both defi- nitions are closely linked with conditions for the growth of the functions. Notably, Kolwankar-Gangal fractional derivatives are sensitive to the critical local H¨olderex- ponents, while the fractional velocities are sensitive to the critical point-wise H¨older exponents and there is no complete equivalence between those quantities [16]. In contrast to physical applications, mathematically, there is no reason to limit the choice of the function in the denominator of the difference quotient to a power function. In such way, more diverse limit objects generalizing derivatives can be LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 3 studied. Such is the objective of the present paper. Here, derivatives are generalized in terms of the class of the modulus of continuity of the primitive function. Such definition focuses on applications involving continuous but possibly non-absolutely continuous functions of a real variable. The paper is structured as follows: Section 2 discusses some properties of totally disconnected sets. Section 3 introduces point-wise oscillation of functions. Section 4 characterizes some of the properties of the moduli of continuity. Section 5 introduces the concept of generalized ω-derivatives, defined from the maximal modulus of continuity. Section 6 discusses the continuity sets of derivatives form the perspective of the theory developed in Sec. 3 and 4. Section 7 introduces the concept of modular derivatives. Notational conventions (see Appendix A) follow previously published works and are repeated here for convenience [22, 23, 24]. The notation |I| for an interval I will mean its length.

2. Totally disconnected sets The following definition is given in Bartle (2001)[3, Part 1, Ch. 2] and Silva (2007) [25, Ch. 2]: Definition 1 (Null sets). A null set Z ⊂ R (or a set of measure 0) is called a set, such that for every 0 <  < 1 there is a countable collection of sub-intervals

{Ik}k∞=1, such that [∞ X∞ Z ⊆ Ik, |Ik| ≤  k=1 k=1 where |.| is the interval length. Then we write |Z| = 0. Remark 1. The next statement is a generalization of a well-known property of countable sets (see for example [25, Ch. 2]). The result is given here for complete- ness of the subsequent presentation. Definition 2 (Totally disconnected space). A M is totally discon- nected if every non-empty connected subset of M is a singleton [28, p. 210]. That is, for every S ⊂ M, S non-empty and connected implies ∃p ∈ M with S = {p}. Theorem 1 (Null set disconnectedness). Suppose that E is a null set. Then E is totally disconnected. Conversely, suppose that E is totally disconnected and count- able. Then E is a null set. Proof. Forward statement: Suppose that Z ⊂ E is connected and open. Then there exist 3 numbers x1 < z < x2, such that [x1, x2] ⊂ Z. Then |[x1, x2]| = x2 − x1 > 0. Therefore, ∃, such that 0 <  ≤ z − x1 < x2 − x1; so that  < |Z| ≤ |E|, which is a contradiction. Therefore, x2 = x1 and hence Z is singleton. Therefore, by induction E is totally disconnected. Converse statement: The countability requirement in the statement of the theorem comes from the fact that there are sets that are totally discon- nected, uncountable and non-null [14, 10] (see example 1). Since E is totally disconnected for every z, w ∈ E, trivially, there is a number h, such that [z − h/2, z + h/2] ∩ [w − h/2, w + h/2] = ∅. Therefore, there is a

collection of such intervals, {Ik}k∞=1 k+1 k+1 Ik = [zk − h/2 , zk + h/2 ] 4 DIMITER PRODANOV

k of length |Ik| = 1/2 . Therefore,

X∞ |Ik| = h k=1 for any such a number h. Since h can be chosen arbitrarily small the claim follows. 

It should be noted that not all totally disconnected sets are null. There are totally disconnected uncountable sets of positive measure. For example, the Smith– Volterra– is of Lebesgue measure 1/2.

Example 1. The construction of the Smith–Volterra–Cantor set is given as follows [14, p.15]: The set is constructed by iteratively removing certain intervals from the unit interval I0 = [0, 1]. At each step k, the length that is removed pk+1 = pk/4 from the middle of each of the remaining intervals. That is, starting from I0 and p0 = 1/4 on every step

l r Ik = [u, v] −→ Ik+1 = [u, (u + v)/2 − pk/2] ,Ik+1 = [(u + v)/2 + pk/2, v]

pk −→ pk+1 = pk/4

For example,

 3   5  k = 1 : I1 = 0, 8 ,I2 = 8 , 1

 5   7 3   5 25   27  k = 2 : I21 = 0, 32 ,I22 = 32 , 8 ,I23 = 8 , 32 ,I24 = 32 , 1

 9   11 5   7 37   39 3  k = 3 : I31 = 0, 128 ,I32 = 128 , 32 ,I33 = 32 , 128 ,I34 = 128 , 8 ,

 5 89   91 25   27 117   119  I35 = 8 , 128 ,I36 = 128 , 32 ,I37 = 32 , 128 ,I38 = 128 , 1 During the process, disjoint intervals of total length

X∞ 1 1 L = = 4 . 2k 2 k=0 are removed so that the resulting set is of measure 1/2. The Smith–Volterra–Cantor set is closed as it is an intersection of closed sets. Furthermore, at step n the length 1 of each closed subinterval is ln = (ln 1 − pn 1). Starting from l0 = 1 one gets 2 − − 1  1 1  l = + n 2 2n 4n

Therefore, by the Nested Interval theorem the SVC set is totally disconnected and contains no intervals. The SVC set was used as an example in [24].

The set presented in the above example can be used to construct a singular func- tion, resembling by some of its properties the famous ”Devil’s staircase” function (see [24]). LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 5

3. Interval and Point-wise Oscillation of Functions The concept of point-wise oscillation can be used to characterize the set of con- tinuity of a function. This can be done in a way similar to the approach presented as theorem 3.5.2 in Trench [26][p. 173]. This is the so-called Oscillation lemma [22, 24]. Since it was published before, the statement of the lemma is relegated to an Appendix. Definition 3 (Oscillation). Define the oscillation of the function f in the interval J = [a, b] as osc f := sup f − inf f J J J Definition 4 (Directed Oscillation). Define the directed oscillations as (i) the forward oscillation: + osc [f](x) := sup f − inf f, I = [x, x + ] [x,x+] [x,x+] and the backward oscillation:

osc−[f](x) := sup f − inf f, I = [x − , x] [x ,x] [x ,x] − − Finally, define the limits, if such exist as finite numbers, as   osc+[f](x) := lim sup f − inf f ,I = [x, x + ] I 0 I I | |→   osc−[f](x) := lim sup f − inf f ,I = [x − , x] I 0 I | |→ I according to previously introduced notation [22, 23]. The Oscillation lemma is of a fundamental importance for it opens up the pos- sibility to characterize the discontinuity of functions in terms of their oscillation at a given point. The oscillation of a function can be viewed in two ways: as a functional having the interval of study fixed; or, alternatively, as a function of the interval having the function under study fixed. There is no ambiguity as in fact both aspects are complementary as will be demonstrated. Definition 5 (super/sub-additivity on an interval). A function f is called sub- additive on the interval I = [x, x + ] if f(x + a) + f(x + b) ≥ f(x + a + b), a, b ∈ [0, ] The converse holds for super-additivity f(x + a) + f(x + b) ≤ f(x + a + b) Example 2. f(x) = x3 is sub-additive in (−∞, 0) and super-additive in (0, ∞). Let a, b > 0. Then a3 + b3 ≤ (a + b)3 since 0 ≤ 3 ab(a + b) and hence super-additivity follows on the positive real axis. The above definition allows one to establish some properties of the oscillation. The subsequent lemma is useful for that purpose: 6 DIMITER PRODANOV

Lemma 1. Let f be a non-decreasing function on I = [x, x + ]. If f is also super-additive on I then sup f + sup f ≤ sup f A B I for A = [x, x + a] ⊂ I, B = [x, x + b] ⊂ I, A ∪ B = I. Conversely, if f is increasing and sub-additive on I then sup f + sup f ≥ sup f A B I Proof. Let I = [x, x + a + b], A = [x, x + a], B = [x, x + b]. Let f be super-additive and non-decreasing on I. Consider the following table of values

a0 ≤ a ≤ b0 < b < c0 < a + b supA f = f(a0) ≤ f(a) ≤ supB f = f(b0) ≤ f(b) ≤ supI f = f(c0) ≤ f(a + b) By the non-decreasing property

M := f(a0) + f(b0) − f(c0) ≤ f(c0) Then also

M ≤ f(a) + f(b) − f(c0) ≤ f(a + b) − f(c0) ≤ 0 by super-additivity. Therefore, sup f + sup f ≤ sup f A B I Let f be sub-additive and increasing on I. It follows also that

M ≥ f(a0 + b0) − f(c0)

However, since the function is bounded a = a0, b = b0, c0 = a + b. Therefore, M ≥ 0 and sup f + sup f ≥ sup f A B I 

2 2 Example 3. Let f(x) = x and I = [0, a + b], 0 < a < b. Then supA f = a , 2 2 supB = b , supI f = (a + b) and a2 + b2 ≤ (a + b)2 which is true. √ √ √ √ Let f(x) = x. Then supA f = a, supB = b, supI f = a + b and √ √ √ a + b ≥ a + b which is true. Lemma 2. Consider a function, which is sub-additive on I = [x, x + ]. Then f is concave on I. That is, for 0 ≤ λ ≤ 1 f(λa + (1 − λ) b) ≥ λ f(a) + (1 − λ) f(b) holds for any a + b ≤ , a, b > 0. Conversely, if f is super-additive on I then it is convex: f(λa + (1 − λ) b) ≤ λ f(a) + (1 − λ) f(b) LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 7

Proof. We prove first the sub-additive case. Consider the integer k ≥ 1. Then for some real a it follows by induction that f(ka) ≤ k f(a). Further, suppose that a = b/k for some b. Then f(b)/k ≤ f(b/k) so that combining with the previous inequality it follows that for a q = p/k ≥ 1 : f(qa) ≤ q f(a). Let r = 1/q and b = a/r; then f(qa) ≤ q f(a) =⇒ r f(r) ≤ f(b r) for r ≤ 1. Since a is arbitrary then the inequality is valid for any b > 0. Letting a = λ/q, λ ∈ R it follows that f(λ) ≤ λ/a f(a) for λ ≥ 1. Since now both variables are real-valued the entire domain becomes real. The concavity of f is established as follows: Let λ = a/(a + b) ≤ 1 and the opposite be assumed true.  a   b  f (a) + f (b) = f (a + b) + f (a + b) = a + b a + b f (λ(a + b)) + f ((1 − λ)(a + b)) So that f (λ(a + b)) + f ((1 − λ)(a + b)) ≤ λ f (a + b) + (1 − λ) f (a + b) = f (a + b) Therefore, f (a)+f (b) ≤ f (a + b), which is a contradiction to the initial hypothesis. Therefore, f is concave on I. The super-additive case can be proven in the same way and holds by duality.  The next point is to establish the properties of oscillation. The reasoning is symmetric with regard to sub-additive and super-additive functions. Lemma 3. Suppose that f is non-negative and super-additive on I = [x, x + ] and infI = 0. Then osc f(x) + osc f(x) ≤ osc f a b a+b  ≤ Proof. Let A ⊂ B ⊂ I. Suppose that f is non-negative and super-additive. osc f+osc f−osc f = sup f+sup f−sup f−inf f−inf f+inf f ≤ inf f−inf f−inf f ≤ 0 A B I A B I A B I I A B Therefore, osc f + osc f ≤ osc f A B I  Lemma 4. Suppose that f is increasing and sub-additive on I = [x, x + ] and infI = 0. Then osc f(x) + osc f(x) ≥ osc f a b a+b  ≤ Proof. Let A ⊂ B ⊂ I. Suppose that f is increasing and sub-additive.

M := osc f + osc f − osc f = sup f + sup f − sup f − inf f − inf f + inf f ≥ A B I A B I A B I inf f − inf f − inf f =: N I A B

However, N ≤ 0, since infI f ≤ infA f +infB f by hypothesis. Therefore, if infI = 0 osc f + osc f − osc f ≥ 0 A B I  8 DIMITER PRODANOV

From this last result it is clear that the condition infI f = 0 is not attainable in general. Moreover, the use of the infimum brings also another point in the interval into consideration, for trivially + osc [f](x) = sup |f(u) − f(v)| ; u,v I,u=v ∈ 6 therefore, the oscillation, defined as above, includes information about two points u and v in some relation to the point of interest x. Therefore, another function will be introduced that maintains the desirable properties established in Lemma 1. 3.1. The Point Oscillation Function. Definition 6. Consider a bounded function f defined on a compact interval I. Define the left (resp. right ) point oscillation functions as  supI |f(x + ) − f(x)| I = [x, x + ] ωx±() := supI |f(x) − f(x − )| I = [x − , x] In such way the directionality information is preserved. It can be established that these quantities are majorized by the oscillation. Proposition 1 (Majorization of the point oscillation).

osc f ≥ ωx±() ≥ |∆± [f](x) |,I = [x, x ± ] I

ωx is a non-decreasing non-negative function. Proof. Suppose that f is positive in I.

osc f = sup f − f(x) + f(x) − inf f ≥ sup f − f(x) = ωx±() I I I I | {z } 0 ≥ The second inequality is trivial. The third assertion follows from the properties of the supremum: For A ⊂ B implies supA f ≤ supB f.  Lemma 5 (Second Oscillation Lemma). Consider a function f continuous in the ∼ ∼ compact interval I of length h. Then f = C [I] ⇐⇒ ωx(h) = C [0, h] and ωx(0) = 0. Proof. Forward implication: The proof follows directly from Prop 1 by application of the First Oscillation Lemma (6). Furthermore, since f(x) = a is fixed then if f is continuous so is f − a. Reverse implication: Consider the right-continuous case. By hypothesis lim ω+() = lim sup |f(x + ) − f(x)| = 0  0  0 → → However, lim sup |f(x + ) − f(x)| ≥ lim inf |f(x + ) − f(x)| =⇒ lim inf |f(x + ) − f(x)| = 0  0  0  0 → → → by majorization. Therefore, both limits coincide and the function f is right- continuous about x. Let + + |ωx () − ωx (0)| ≤ µ where µ is arbitrary but fixed. sup f = f(x ), sup f = f(x ), Then for I = I 0 I0 00 0 [x, x + 0] + + |ωx () − ωx (0)| = | sup f − sup f| = |f(x0) − f(x00)| ≤ µ I I0 LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 9

However, |x0 −x00| ≤ min(, 0) := δ. Since  and 0 can be made arbitrary small then f is continuous in x. The left- continuous case can be proven by the substitution  → −.  Corollary 1. The following two statements are equivalent

lim ωx±() > 0 ⇐⇒ lim f(x ± ) =6 f(x)  0  0 → → Theorem 2 (Properties of ωx() for sub-additive functions). Consider a sub- additive bounded function f on I = [x, x + ]. Consider any two compact nested sub-intervals Ia = [x, x + a) ⊆ Ib = [x, x + b) ⊆ I, such that a + b ≤  and assume that ωx(a) =6 0 and ωx(b) =6 0. Then the triangle inequality holds:

ωx(a + b) ≤ ωx(a) + ωx(b) (1) Under the same hypothesis, for a real λ ≥ 1

ωx(λa) ≤ λ ωx(a) (2)

ωx(a + λb) ≤ ωx(a) + λ ωx(b) (3) For a real λ, such that 0 ≤ λ ≤ 1,

ωx(λa) ≥ λ ωx(a) (4)

ωx(a + λb) ≥ ωx(a) + λ ωx(b) (5)

Moreover, ωx is concave. Proof. Inequality 1 follows from Lemma 1. For the other inequalities the approach is more nuanced. Suppose that f is increasing. Then inequalities 2 – 5 and concavity follow from Lemma 2. Suppose that f is non-negative and non-decreasing. Consider the following table of values

a0 ≤ a ≤ b0 < b < c0 < a + b supA f = f(a0) ≤ f(a) ≤ supB f = f(b0) ≤ f(b) ≤ supI f = f(c0) ≤ f(a + b) Let M := ωx(a0) + ωx(b0) − ωx(c0) = f(a0) + f(b0) − f(c0) − f(x) Then there are two cases to consider: Let f(x) ≤ f(a0) ≤ f(b0) = f(c0). Then M = f(a0) − f(x) ≥ 0. Therefore, Lemma 2 can be applied as well. Let f(x) ≤ f(a0) = f(b0) ≤ f(c0). Then

M = 2f(a0) − f(c0) − f(x) ≤ f(a0) − f(c0) ≤ 0

However, in this case ωx(a) = 0 so it must be excluded for the assertion to hold. Inequality 5 follows from the concavity.  Conversely, for a super-additive bounded function f on I an analogous result can be stated.

Theorem 3 (Properties of ωx() for super-additive functions). Consider a super- additive bounded function f on I = [x, x + ]. Consider any two compact nested sub-intervals Ia = [x, x + a) ⊆ Ib = [x, x + b) ⊆ I, such that a + b ≤  and assume ωx(a) =6 0 and ωx(b) =6 0. Then

ωx(a + b) ≥ ωx(a) + ωx(b) (6) For a real λ ≤ 1 ωx(λa) ≥ λ ωx(a) (7) 10 DIMITER PRODANOV

ωx(a + λb) ≥ ωx(a) + λ ωx(b) (8) For a real λ, such that 0 ≤ λ ≤ 1,

ωx(λa) ≤ λ ωx(a) (9)

ωx(a + λb) ≤ ωx(a) + λ ωx(b) (10)

Moreover, ωx is convex. Proof. Inequality 6 follows from Lemma 1. Inequalities 7 – 10 and convexity follow from Lemma 2.  The next definition is given by Bartle [3][Part 1, Ch. 7, p. 103]. Definition 7 (Function of Bounded Variation ). Consider the interval I = [a, b]. A partition of I is a set of n + 1 numbers P[I] := (a < x1 . . . xn 1 < b). The − function f : R 7→ R is said to be of bounded variation on I if and only if there is a constant M > 0 such that n+1 X V [I] := (P) |f(xi − xi 1)| ≤ M P − i=1 The total variation of the function is defined as V ar(f, I) := sup V [I] P P where the supremum is taken in all partitions P. The class of function of bonded variation in a compact interval I will be denoted as BV [I]. Under this definition the following proposition can be stated: Proposition 2. If f is either monotonously increasing or decreasing on I = [x, x+ + ] then ωx() = |f(x+)−f(x)| = |∆ f(x)|. If f is either monotonously increasing or decreasing on I = [x − , x] then ωx() = |f(x − ) − f(x)| = |∆−f(x)|. Under ∼ the same hypotheses ωx = BV [I]. n Proof. Consider an increasing collection of intervals U(n) = {[x, x + ak]}k=1 such Sn that a1 < ··· < an. Then these form a partition P[x, x + an] over I = k=1[x, x + ak]. Then V ar[ωx,I] = ωx(an) − ωx(a1), which is bounded.  The next theorem is a consequence of the Darboux–Froda theorem. The Darboux– Froda theorem states that the set of discontinuities of a monotone function is at most countable. Hence, by Th. 1 it is also totally disconnected. In fact, the latter inference can be strengthened to arbitrary functions as Th. 5. Theorem 4 (Continuity set of oscillation). Consider a bounded function f defined on a compact interval I = [x, x + h] and let it be given. The discontinuity set ∆ω[I] of the oscillation function ωx is a null set. If f is strictly increasing (respectively decreasing) on I then the continuity set of the oscillation ωx can be written as [∞ Cω[I] = (ak, bk), bk ≤ ak+1 k=1 Proof. Consider the interval I = [x, x + h] with length h = |I| and denote the left- open interval Jh = (0, h] then ∃q ∈ Q ∩Jh. Therefore, there is a map Jh 7→ q. Since ωx(h) is non-decreasing, it has only jump discontinuities (since bounded, increasing LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 11 of numbers have limits). Indeed, by the LUB and GLB properties, if ωx is (right-)continuous about h

sup ωx(h + ) − L1 ≤ µ/2,  :: µ 

inf ωx(h + ) − L2 ≤ µ/2 −→ 

sup ωx(h + ) − inf ωx(h + ) − L1 + L2 = |ωx(h + ) − ωx(h) − L1 + L2| ≤ µ   by the non-decreasing property (Prop. 1). On the other hand,

|ωx(h + ) − ωx(h)| = sup f − sup f ≤ sup f − f(x) = ωx(h) h+ h h

Therefore, if h∈ / Cω, that is, if L1 =6 L2 by the Second Oscillation Lemma there is some number x0, such that

ωx0 (h) ≥ L1 − L2 > 0 and there is a jump discontinuity at h. In such a case, ∃p ∈ Q ∩ [L1,L2] so that [L1,L2] = Lp can be labelled for a uniquely chosen rational p. Let dp = |L2 − L2|. If the number dp is the same for all x ∈ I then we are done. On the other hand, suppose that dp varies in function of x. Since ωx is non-decreasing then for another p0 =6 p −→ Lp ∩ Lp0 = ∅. Therefore, there is an isomorphism p ←→ h. Therefore, the set of continuity of ωx is Jh \{h} which is an open countable interval. Hence, the second claim follows.  Having established these properties, we will characterize the set of discontinuities of a function using the following definition: Definition 8. Define the set of discontinuity for the function F in the compact interval I as  ∆[F,I] := x : osc±[F ](x) > 0, x ∈ I or if the context is known ∆[F,I] ≡ ∆[I]. In particular, under this definition osc[F ](x) = ∞ is admissible. Theorem 5 (Disconnected discontinuity set). Consider a bounded function F de- fined on the compact interval I. Then its set of discontinuity ∆[F,I] is totally disconnected in I. Proof. Consider a decreasing collection of closed nested intervals from a partition of n the interval I1 = [x, x+h](h > 0, not necessarily small); that is {Ik = [x, x + ak]}k=1, Ik+1 ⊂ Ik ⊂ ... ⊂ I1. Since the case when the function is locally constant is trivial we consider only two cases: increasing and decreasing. Let En = In ∩ In 1 and define − ∆n := sup F − inf F In In ∆n 1 := sup F − inf F − In 1 In 1 − − ∆E := sup F − inf F En En 12 DIMITER PRODANOV as indicated in Fig. 1 : F increasing F decreasing y y

∆n 1 ∆n 1 − −

∆n ∆n

x x In In 1 In In 1 − − sup F = sup F sup F = inf F En In 1 En In − sup F = inf F inf F = inf F En In 1 En In − inf F = inf F sup F = sup F In In 1 In In 1 − −

Figure 1. Schematic of the interval construction

Increasing case: Suppose that F is increasing in In 1 then −

∆n 1 − ∆n = sup F − inf F − sup F + inf F = − In 1 In 1 In In − − sup F − inf F − inf F + inf F = ∆E = osc f In 1 En In 1 En En − − Decreasing case: Suppose that F is decreasing in In 1 then − ∆n 1 − ∆n = sup F − inf F − sup F + sup F = ∆E = osc f − In En In En En

On the other hand, osc f = ∆n and osc f = ∆n 1. If F is continuous on the In In 1 − − opening of In 1, in limit − lim osc f − osc f = 0 ⇒ lim osc f = 0 n In In 1 En 0 En →∞ − | |→ and ∆[F,I] \{x} = ∅. On the other hand, F can be discontinuous on the boundary points {x, x + an 1}, which are disconnected. Suppose then− that both endpoints are points of discontinuity. Hence, {x, x + an 1} ⊆ ∆[F,I]. If F is discontinuous on the opening of In 1, we take In and proceed− in the same way. Therefore, ∆[F,I] =6 ∅ is a union of totally− disconnected sets and hence it is totally disconnected.  4. Moduli of continuity The moduli of continuity will be discussed as second-order properties of the preimage functions and will be studied in a point-wise manner. A modulus will be indexed by a point x in the domain of the preimage function but no other restrictions will be placed there. LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 13

Definition 9 (Modulus of continuity). A point-wise modulus of continuity gx : R 7→ R of a function f : R 7→ J ⊆ R is a (1) non-decreasing , such that (2) gx(0) = 0 and (3) |∆± [f](x) | ≤ K gx() holds in the interval I = [x, x ± ] ⊂ J for some constant K.

A regular modulus is such that gx(1) = 1. Unsurprisingly, under this definition every continuous function admits a modulus of continuity: Theorem 6 (Modulus characterization theorem). Every continuous function ad- mits a modulus of continuity on an interval, which is a subset of its domain. Any modulus of continuity is BVC[I] for such interval I.

Proof. Consider the point oscillation function ωx(). Then ωx is non-decreasing. Trivially, |∆± [f](x) | ≤ ωx() holds. Finally, ωx is continuous and ωx(0) = + osc [f](x) = 0 by Prop. 5. Then gx() = ωx()/ωx(1).  The point oscillation function used is the proof of Th. 6 will be called a canonical modulus of continuity of a continuous function.

4.1. Classification of the moduli of continuity.

Proposition 3. Suppose that ωx is strictly sub-additive in I = [0, h]. Then ωx0 (0) = ∞. Suppose that ωx is additive. Then ωx is linear and homogeneous in h and ωx0 (0) exists. Suppose that ωx is super-additive. Then ωx0 (0) exists. Proof. The proof proceeds in three cases. Strictly sub-additive case: Suppose that the derivative exists finitely and let M > ωx0 (0) ≥ m > 0. By sub-additivity there is h, such that 2 h 1 2 ω (h/2) > ω (h) ⇒ M > ω > ω (h) ≥ m x x h x 2 h x Then by induction: 2n  h  M 1  h  m M > ω ≥ m ⇒ > ω ≥ h x 2n 2n h x 2n 2n Taking the limit in n → ∞ leads to

0 > ωx0 (0) ≥ 0 which is a contradiction. Therefore, the limit does not exist finitely and ωx0 (0) = ∞. Additive case: By additivity, for all integer k : ωx(kh) = k ωx(h). Then by change of variables z = kh. ωx(z) = k ωx(z/h). Therefore, ωx(qh) = qωx(h) for all rational q. Then by continuity, ωx(h) = Kh for some K > 0. Super-additive case: 2 h 1 2 ω (h/2) ≤ ω (h) ⇒ ω ≤ ω (h) x x h x 2 h x 14 DIMITER PRODANOV

Then by induction: 2n  h  2n 1  h  1 ≤ ≤ − ≤ ≤ 0 ωx n ωx n 1 ... ωx (h) h 2 h 2 − h Since the is bounded from below and decreasing it has a limit. Taking the limit in n → ∞ leads to 1 ω0 (0) ≤ ω (h) x h x therefore, the derivative exists finitely at h = 0.  Example 4. An illustrative example of the super- additive case is the function 1 f(x) := e− x2 . For every n > 0

1 e− x2 lim = 0 x 0 n → x   ∈ − √3 √3 The function is super- additive in the interval x 2√log 2 , 2√log 2 since there

3 2e− 4x2 < 1 Based on this classification result it is useful to apply the following definition. Definition 10 (g-continuous class). Define the growth class C g[I] induced by the modulus of continuity g(|I|) by the conditions: If f =∼ C g[I] on the compact interval I then ωx() (1) Cx = lim exists finitely,  0 g() → (2) Cx is non-zero and (3) |∆± [f](x) | ≤ Cx g(). g for  = |I|. To emphasize the dependence on x we may write Cx and skip I if it is known. This definition encompasses the definitions of the H¨older and Lipschitz functions. So that L ≡ H 1 ≡ C g for g(x) = x or H α ≡ C g, for g(x) = xα, 0 < α < 1. By Prop. 3 the modular functions can be classified into three distinct types Lipshitz: for which lim ωx() < L for some L. These are either linear or  0  → otherwise ωx0 (0) exits finitely and hence they are Lipschitz. Singular: (or strongly non-linear) for which the ratio ωx()/ diverges and ωx0 (0) = ∞ by the L’Hˆopital’srule.

5. Generalized maximal ω derivatives Definition 11. For a function f define superior and inferior, and respectively forward and backward, maximal ω modular derivatives, as the limit numbers L

∆± [f](x) ¯ sup − L < µ =⇒ Dω±f(x) = L  ωx ()

∆± [f](x) inf − L < µ =⇒ Dω±f(x) = L  ωx () ¯ for all  :: µ,  > 0. LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 15

Remark 2. These derivative functions obviously generalize the concept of Dini derivatives (Def. 19) Equipped with the above definition we can state the first existence result: Theorem 7 (Bounded ω-derivatives). For a continuous function the four derivative functions exist as real numbers. Moreover, if f is non-decreasing about x+ ¯ Dω±f(x) = 1

0 ≤ D±f(x) ≤ 1 ¯ ω while if f is non-increasing about x−

D±f(x) = −1 ¯ ω ¯ 0 ≥ Dω±f(x) ≥ −1 Proof. Let I = [x, x ± ] be given and x is fixed but we can vary . Consider the auxiliary function  ∆± [f](x) υω± [f](x) := (11) ωx ()  The supremum definitions are restatements of the LUB property for υω± [f](x) in terms of the variable , while the infimum derivatives are restatements with the GLB property of the reals again for the same variable. Therefore, all four numbers exist for a given argument x and, therefore, under the above hypothesis. Moreover, since ¯ |∆± [f](x) | ≤ ωx () then for an non-decreasing function |Dω±f(x)| ≤ 1. Therefore, by the supremum property sup ∆±[f](x) = 1. For a decreasing function f it is  ωx() sufficient to consider −f and apply the same arguments.  Corollary 2. Suppose that f is monotone and continuous function on a compact interval. If f is increasing in [x, x + ] = I then D¯+f(x) = D+f(x) = 1. If f is ω ¯ ω decreasing in [x, x + ] = I then D¯+f(x) = D+f(x) = −1. If f is increasing in ω ¯ ω [x − , x] = I then D¯ f(x) = D f(x) = 1. If f is decreasing in [x − , x] = I then ω− ¯ ω− D¯ f(x) = D f(x) = −1. ω− ¯ ω− Proof. Fix x and consider I = [x, ±]. The proof follows from the fact that in both cases |∆± [f](x) | = ωx ().  We can give generalized definition of local differentiability (called ω-differentiability) as follows Definition 12. A function f is ω-differentiable at x if at least one of the two limits exist

∆± [f](x) − L < µ =⇒ Dω±f(x) = L ωx () where the conventions for L, µ and  are as above. Note that the definition only supposes that the one-sided limits of the increments + – Dω f(x) (respectively Dω−f(x)) exist as real numbers. That is, + Dω f(x) =6 Dω−f(x) is admissible. This is the minimal statement that can be given for the limit of an in- crement of a function. Nevertheless, based on two strong properties – monotonicity and continuity – it can be claimed that 16 DIMITER PRODANOV

Proposition 4 (Monotone ω-differentiation). If a function f is monotone and continuous in a closed interval I then it is continuously ω-differentiable everywhere in the opening I◦.

Proof. The continuity follows directly from Corr. 2, while the restriction comes from the fact that at the boundary only one of the increments can be defined without further hypothesis for the values of f outside of I. 

Proposition 5 (BVC ω-differentiation). If the function f is BVC[I] in a closed interval I then it is ω-differentiable everywhere in the opening I◦.

Proof. The proof follows from the Jordan theorem, since a BV[I] function can be decomposed into a difference of two non-decreasing functions. On the other hand, without further hypotheses we can not claim anything about the eventual equality + of Dω f(x) and Dω−f(x) since Jx = [x − , x] ∩ [x, x + ] = {x} so that we can form only the trivial map x 7→ {x}, which without further restrictions of the domain of x (i.e. by means of some topological obstructions) is uncountable. 

We can further utilize the concept of oscillation to give a concise general differ- entiability condition as

∆± [f](x) lim osc = 0 (12)  0 → ωx () Theorem 8 (Characterization of ω-derivative). The following implications hold

D¯±f(x) = D±f(x) = D±f(x) =⇒ f =∼ C[x±] ω ¯ ω ω

∆± [f](x) ¯ lim osc = 0 ⇐⇒ Dω±f(x) = Dω±f(x) = Dω±f(x)  0 ¯ → ωx () so that if Eq. 12 holds at x then f is ω-differentiable (and hence continuous) at x.

Proof. Continuity implication: Consider the inequality

¯ ∆± [f](x) Dω±f(x) = Dω±f(x) =⇒ − L ≤ µ/2,  :: µ ¯ ωx () so that ∆ [f](x) L − µ/2 ≤ ± ≤ L + µ/2 =⇒ ωx ()

sup ∆± [f](x) ≤ (L + µ/2) ωx () 

(L − µ/2) ωx () ≤ inf ∆± [f](x) =⇒ 

sup ∆± [f](x) − L ≤ µ/2 ωx ()

inf ∆± [f](x) − L ≤ µ/2 ωx () LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 17

Let sup ∆± [f](x) = M and inf ∆± [f](x) = m. Then by triangle in- equality sup ∆ [f](x) inf ∆ [f](x) M − m  ± −  ± = ≤ µ ωx () ωx () ωx () M − m ≤ µ ωx ()

Therefore, in limit M − m ≤ 0, hence M = m and f is continuous. This sequence of operations reminds the fact that real numbers are constructed by a limiting process. Forward statement: Suppose that D¯ f(x) = L and D f(x) = L Then ω± 1 ¯ ω± 2 by LUB

∆± [f](x) sup − L1 ≤ µ/2  ωx ()

∆± [f](x) inf − L2 ≤ µ/2  ωx () so that

∆± [f](x) ∆± [f](x) sup − L1 + inf − L2 ≤ µ  ωx ()  ωx () Then by the triangle inequality

∆± [f](x) ∆± [f](x) |L1 − L2| ≤ sup − inf +L1 − L2 ≤ µ  ωx ()  ωx () | {z } ∆±[f](x) osc ωx() Then in limit by lemma. 6

|L1 − L2| ≤ 0 =⇒ L1 = L2 Further, starting from ∆ [f](x) ∆ [f](x) ∆ [f](x) inf ± ≤ ± − ≤ sup ± =⇒  ωx () ωx ()  ωx ()

∆± [f](x) ∆± [f](x) ∆± [f](x) ∆± [f](x) ∆± [f](x) 0 ≤ − inf ≤ sup − inf = osc ωx ()  ωx ()  ωx ()  ωx () ωx () Therefore,

∆± [f](x) ∆± [f](x) ∆± [f](x) ∆± [f](x) − inf ≤ µ =⇒ sup − ≤ µ ωx ()  ωx ()  ωx () ωx () Therefore, all three limits coincide. Converse statement: Suppose that

D¯±f(x) = D±f(x) = L > 0 ω ¯ ω 18 DIMITER PRODANOV

By hypothesis

∆± [f](x) L − inf ≤µ/2  ωx () | {z } A

∆± [f](x) sup − L ≤µ/2  ωx () | {z } B

∆± [f](x) ∆± [f](x) sup − inf ≤ |A| + |B| ≤ µ  ωx ()  ωx () | {z } ∆±[f](x) osc ωx() Therefore, in limit

∆± [f](x) 0 ≤ lim osc ≤ 0  0 → ωx ()

∆±[f](x) so that lim osc = 0.  0 ωx() → 

Corollary 3 (Range of Dω± ). The range of Dω± is given by the discrete set

Dω±f(x) ⊆ {−1, 0, +1}

Proof. Let I = [x, x + ] be given. If f is constant in I trivially Dω±f(x) = 0. If f is increasing in I then Dω±f(x) = 1 and by duality if f is decreasing in I then Dω±f(x) = −1.  The ω non-differentiability set of a continuous function can be characterized by the following theorem. Theorem 9 (ω non-differentiability set). Consider the function f =∼ C[I] on the compact interval I. Then the sets +  + +  ∆ [I] := x : D¯ f(x) > D f(x) ∩ I, ∆−[I] := x : D¯−f(x) > D−f(x) ∩ I ω ω ¯ ω ω ω ¯ ω are null sets. That is for a continous function the ω non-differentiability set is null. Proof. Consider the case wherever the right ω-derivative does not exist. That is, the defining quotient oscillates without a limit. Then for 0 < u, v ≤ δ + + ∆u [f](x) ∆v [f](x) − > µ (D1) ωx (u) ωx (v) for some µ > 0. We can consider a variable ξ ∈ [x, x + u] ∩ [x, x + v] = [x, x + min(u, v)] = J. There is a rational r ∈ Q ∩ J. Associate (r, J) ≡ Jr so that Jr can be counted by an enumeration of the rationals and index δ :: r. Therefore, the set

[∞ ∆ω := {z : D1 true, z ∈ Jk} k=1 LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 19 is countable ∀δ > 0. Since ∆ω is totally disconnected by Th. 5 we can select k δk = δ/2 and Jk ⊂ Jr. Therefore, X∞ X∞ δ |∆ | = |J | ≤ = δ ω k 2k k=1 k=1 and ∆ω is a null set. The left derivative case holds by duality.  Note that this is the best possible result for the local-type of derivatives and partially corresponds to the expectation of Ampere.

6. Continuity sets of derivatives In the following we re-state the classical result of the Lebesgue differentiation theorem. The poof is given using the machinery of ω-differentiation. In the follow- ing argument I reserve the term ”strictly monotone function” to mean only a strictly increasing or strictly decreasing function in an interval. Theorem 10 (Lebesgue monotone differentiation theorem). Suppose that f is strictly monotone and continuous in the compact interval I. Then f is continu- ously differentiable almost everywhere. The set  ∆f [I] := x : f+0 (x) =6 f 0 (x) ∩ I − is a null set.

Proof. Let Dωf(x) = L > 0. By Corr. 2 for  :: µ

∆+ [f](x) ω () + −  x ≤ L + µ/2 ωx ()  | {z } A

∆− [f](x) ωx ()− − L ≤µ/2 ωx ()−  | {z } B

+ + ∆ [f](x) ωx () ∆− [f](x) ωx ()− − ≤ |A| + |B| ≤ µ ωx ()  ωx ()− 

+ ωx () ωx ()− − ≤ µ   Therefore, by monotonicity using the original notation + 2 ∆ [f](x) ∆− [f](x) ∆ [f](x) − = ≤ µ    hence f+0 (x) = f 0 (x) and ∆f [I] = ∅.  − Recall the definitions of nowhere monotone functions: Definition 13. A function f is non-decreasing on I = [a, b] if given any a < x < y < b f(y) − f(x) ≥ 0 20 DIMITER PRODANOV and non-increasing on I if f(y) − f(x) ≤ 0. A function, which is neither non-decreasing nor non-increasing changes direction of growth in I. A function is nowhere monotone (NM[I]) if given any a < x < y < z < b (f(y) − f(x)) (f(z) − f(y)) ≤ 0 so that NM[I] function is neither non-decreasing nor non-increasing on any sub- interval of I. A function, which is nowhere monotone at a point (NM[y]), is treated as above while y is fixed. From the Lebesgue monotone differentiation theorem it follows that a nowhere differentiable function on an open interval I is simultaneously nowhere monotone on I. Brown et al. establish that no continuous function of bounded variation BVC is MN[y]. [5][Th. 12. Corr. 3]. That is to say NM[x] for x ∈ I as above. Therefore, it is of interest to establish the following result. Theorem 11 (NM continuous ω-differentiability). Suppose that f =∼ C[I] and f =∼ NM[I]. Then ∼ Dω±f(x) = C[I] =⇒ Dω±f(x) = 0

Proof. The set {x : Dω±f(x) = 1} is totally disconnected. By duality, the set {x : Dω±f(x) = −1} is also totally disconnected. Hence, only the set {x : Dω±f(x) = 0} has connected components.  Theorem 12 (Continuity of derivatives). Consider a bounded and continuous func- tion f on a compact interval I. Suppose that f+0 (x) and f 0 (x) are separately con- tinuous then the following holds: −

(1) f+0 (x) = f 0 (x) = f 0(x) − (2)∆ f,I := {x : f 0 ∈/ C, x ∈ I} is totally disconnected with empty interior. (3) The total discontinuity set can be written as ∆f,I = ∆1,f ∪ ∆2,f , where ∆1,f is Fσ and ∆2,f is a null set. (4) The continuity set Cf is Gδ. Proof. Consider the interval I = [u, v]. Then there is rational r ∈ Q ∩ I. Associate (r, I) ≡ Ir so that Ir can be counted by an enumeration of the rationals. Assume that f+0 (x) and f 0 (x) are separately continuous on the opening of Ir◦ = − Ir − {u} − {v}. Fix x, such that u ≥ x > v. u > v u ≥ x > v f(u) − f(v) f(u) − f(x) + f(x) − f(v) = = u − v u − v f(u) − f(x) u − x f(x) − f(v) x − v + = u − x u − v x − v u − v | {z } | {z } 1 λ λ − f(u) − f(x) f(x) − f(v) (1 − λ) + λ u − x x − v ↓ lim ↓ lim u x v x → →  (1 − λ)f+0 (x) + λf 0 (x) = f+0 (x) − λ f+0 (x) − f 0 (x) − − LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 21

By continuity  lim f+0 (v) = f+0 (x) = f+0 (x) − λ f+0 (x) − f 0 (x) v x − → However, since x and hence λ =6 0 is arbitrary f+0 (x) = f 0 (x) must hold ∀x ∈ Ir◦. − Hence, f 0 is continuous on Ir◦. ∼ By this argument we establish that the set ∆1,f := {x : f 0 =6 C} ∩ I is Fσ, where we also assume that whenever f 0(x) does not exist it is replaced by a value that makes f 0 discontinuous. By Th. 5 the discontinuity set is totally disconnected and with empty interior. Let us further consider the case wherever left and right derivatives do not exist (either diverge or oscillate without a limit). It is enough to consider the right derivative. Then we have that for 0 < u, v ≤ δ + + ∆u [f](x) ∆v [f](x) − >  > 0 (D2) u v for some . We can consider a variable ξ ∈ [x, x+u]∩[x, x+v] = [x, x+min(u, v)] = J. There is a rational r ∈ Q ∩ J. Associate (r, J) ≡ Jr so that Jr can be counted by an enumeration of the rationals and index δ :: r. Therefore, the set [∞ ∆2,f := {z : D2 true, z ∈ Jk} k=1 is countable ∀δ > 0. Since it is totally disconnected by Th. 5 we can select k δk = δ/2 . Therefore, X∞ X∞ δ |∆ | = |J | ≤ = δ 2,f k 2k k=1 k=1 and ∆2,f is a null set. The same argument can be applied to the left derivative considering f(−x). The total discontinuity set can be written as

∆f,I = ∆1,f ∪ ∆2,f c Therefore, the continuity set can be written as Cf = (∆1,f ∪ ∆2,f ) hence it is Gδ.  7. Modular derivatives As indicated in Sec 1, the derivatives can be generalized in different directions. If locality is the leading requirement, then the most natural way for such general- ization is to replace the assumption of local Lipschitz growth with the more general modular-bound growth. In such way one can generalize, previously introduced fractional velocity of Cherbit [6]. Definition 14. Define g-variation operators as

 ∆± [f](x) υ ± [f](x) := (13) g g() for a positive  and a modular function g. Condition 1 (Modulus-bound growth condition). For given x and a modular func- tion g. osc±f(x) ≤ Cg () (C1) for some C ≥ 0 and  > 0. 22 DIMITER PRODANOV

Condition 2 (Vanishing oscillation condition). For given x and  > 0  osc±υg± [f](x) = 0 (C2) where the limit is taken in . Define the modular derivative as: Definition 15 (Modular derivative, g-derivative). Consider an interval [x, x ± ] and define ∆± [f](x) Dg±f (x) := lim (14)  0 → g() for a modulus of continuity g(). The last limit will be called modular derivative or a g-derivative. + NB! We do not demand equality of Dg f(x) and Dg−f(x). We are ready to establish the existence conditions of the g-derivative. + Theorem 13 (Conditions for existence of g-derivative). If Dg f (x) exists (finitely), then f is right-continuous at x and C1 holds, and the analogous result holds for Dg−f (x) and left-continuity. Conversely, if C2 holds then Dg±f (x) exists finitely. Moreover, C2 implies C1. Proof. We will first prove the case for right continuity. Condition C1 trivially  implies the g- continuity, which according to our notation is given as υg± [f](x) ≤ Cg(). Forward statement: Without loss of generality suppose that L > 0 is the value of the limit. Then by hypothesis + ∆ [f](x) − L < µ g() holds for every µ :: δ,  < δ . Straightforward rearrangement gives |f(x + ) − f(x) − Lg()| < µ g() Then by the reverse triangle inequality |f(x + ) − f(x)| − Lg() ≤ |f(x + ) − f(x) − Lg()| < µ g() so that |f(x + ) − f(x)| < (µ + L) g(). Further, by the least-upper-bound property there exists a number C ≤ µ + L, such that |f(x + ) − f(x)| ≤ Cg(), which is precisely the Modulus bound growth condition. The left continuity can be proven in the same way. Converse statement: In order to prove the converse statement we can observe that condition C2 + + implies that osc υg [f](x) = 0 so that ∆+ [f](x) osc+  ≤ µ  g() for µ ::  (and in particular for a Cauchy null-sequence µ) so that + + ∆ [f](x) ∆ [f](x) sup − inf ≤ µ  g()  g() LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 23

by lemma 6 and ∆+ [f](x) ∆+ [f](x) sup  ≤ µ + inf  ,  g()  g() so that taking the limits in µ (and hence ) implies

∆+ [f](x) ∆+ [f](x) lim sup  = lim inf   0 g()  0 g()a → → + + Hence lim υg [f](x) = L = Dg f (x) for some L.  0 However,→ the latter limit can be rewritten from its definition as

+ ∆ f(x) − Lg() < µ g() for an arbitrary µ :: . Then since µ is arbitrary by the least upper bound property there is 0, such that

+ + ∆ f(x) = osc [f](x) ≤ (µ + L)g(0) 0 0

for µ :: 0 and we identify condition C1. The left case follows by applying the right case, just proved, to the reflected function f(−x). 

7.1. Generalized Taylor-Lagrange property.

Proposition 6 (Generalized Taylor-Lagrange property). The existence of Dg±f (x) =6 0 implies that

f(x ± ) = f(x) ± D±f (x) g() + (g()) (15) g O for the modular function g. While if

f(x ± ) = f(x) ± Kg() + γ g() uniformly in the interval x ∈ [x, x + ] for some Cauchy sequence γ = and  Ox K =6 0 is constant in  then Dg±f (x) = K. Proof. We prove only for the forward modular derivative. The case for the backward modular derivative is proven in the same way following a reflection of the function argument x. Forward statement: By the definition of the modular derivative ∃γ, such that f(x + ) = f(x) + D+f (x) g() + γ. Moreover, γ = (g()). g O Converse statement: Suppose that

f(x + ) = f(x) + Kg() + γ g()

uniformly in the interval x ∈ [x, x + ] for some number K and γ = .  Ox Then this fulfils both Modulus bound growth and Vanishing oscillation + conditions. Therefore, K = Dg f (x) observing that lim γ=0.  0 →  24 DIMITER PRODANOV

7.2. Characterization by ω-derivatives. Proposition 7. Consider the modular function g. Then

Dω±f (x) = KDg±f (x) for some constant K wherever all limits exist. Proof. Let ω () K = lim x  0 → g() and suppose that the limit exists as a finite number. Then ∆+ [f](x) ∆+ [f](x) ω ()  =  x g() ωx () g() Therefore, the statement of the result follows.  In view of Prop. 4 this means that a function can change its modulus of continu- ity point-wise. Since the cases of H¨olderand Lipschitz functions have been treated extensively in literature we will consider only the general case. 7.3. Continuity of g-derivatives. Gleyzal [13] established that a function is Baire class I if and only if it is the limit of an interval function. Therefore, Dg±f (x) are Baire class I from which it follows that Dg±f (x) must be continuous on a dense set. Moreover, since the continuity set of a function is a Gδ set, (i.e. an intersection of at most countably many open sets), from the Osgood-Baire Category theorem it follows that the set of points of discontinuity of Dg±f (x) is Fσ meagre (i.e. a union of at most countably many nowhere dense sets or else it has empty interior). Since in the previous sections it was established that the modulus of continuity can be conveniently classified as used conventionally in applied literature we are ready to state an important result concerning the continuity of g-derivatives. First, we have the following theorem: Theorem 14 (Continuity of g-derivatives). Suppose that g is a strictly sub- additive modular function on the compact interval I. Then wherever Dg±f (x) is continuous it is zero. That is ∼ Dg±f (x) = C[I] ⇒ Dg±f (x) = 0 + Proof. Let Dg f (x) = K > 0. ∆+ [f](x) f(x + ) − f(x + /2) f(x) − f(x − /2)  = + = g() g() g() f(x + ) − f(x + /2) g(/2) f(x) − f(x − /2) g(/2) = + g(/2) g() g(/2) g() Therefore, in limit supremum and by hypothesis of continuity 2 g(/2) K = K lim sup  0 g() | → {z } G By strict sub-additivity 2g(/2)/g() < 1, therefore, the limit G exists. So it is established that K = GK < K, which is a contradiction. Therefore, K = 0 on the first place. The case for the left derivative follows by duality.  LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 25

Corollary 4. The continuity requirement is equivalent to requiring that 2 g (/2) lim 0 = 1  0 → g0() Theorem 15. Consider a function f having a strictly sub-additive modulus func- tion g on the compact interval I. Then the set

χg±(f) := {x : Dg±f (x) =6 0} ∩ I is totally disconnected and of measure zero, that is |χg±(f)| = 0. The set χg± will be called the set of change of f. Proof. Using the same argument as in the proof of Th. 14 we establish that either K = 0 allowing for continuity of Dg±f (x) or K =6 0 but then Dg±f (x) can not be continuous. Furthermore, by Th. 12 it follows that |χg(f)| = 0.  β Corollary 5. Under the same notation, let g() =  , for β ∈ (0, 1]. If |χg(f)| > 0 then β = 1 and f is Lipschitz.

Corollary 6. Under the same hypotheses the image set Dg±f is totally disconnected. 8. Discussion The relaxation of the differentiability assumption opens new avenues in describ- ing non-linear physical phenomena, for example, using stochastic calculus or the scale relativity theory developed by Nottale [20], which assume fractal character of the space-time geodesics and hence of quantum-mechanical paths. In contrast to the Riemann-Liouville or Caputo fractional derivatives, the geo- metrical, and hence physical, interpretation of a modular derivative is easier to es- tablish due to its local character and the demonstrated generalized Taylor-Lagrange property. That is, presented results demonstrate that the modular derivative pro- vides the best possible local non-linear approximation for its natural modulus of continuity function at the point of interest. The desirable properties of the derivatives, such as their continuity, are estab- lished from the more general perspective of the moduli of continuity. From the perspective of approximation, derivatives can be viewed as mathematical idealiza- tions of the linear growth. The linear growth, i.e. the Lipschitz condition, has special properties, which make it preferred. Importantly, the statements of the Th. 14 and 15 give further insight on why the ordinary derivatives are so useful for describing physical phenomena in terms of differential equations.

Appendix A. General definitions and conventions The term variable denotes an indefinite number taken from the real numbers. Sets are denoted by capital letters, while variables taking values in sets are denoted by lowercase. The action of the function is denoted as f(x) = y. Implicitly the mapping acts on the real numbers: f : R 7→ R. If a statement of a function f fulfils a certain predicate with argument A (i.e. P red[A]) the following short-hand notation will be used f =∼ P red[A]. Square brackets are used for the arguments of operators, while round brackets are used for the arguments of functions. The term Cauchy sequence will always be interpreted as a null sequence. 26 DIMITER PRODANOV

Everywhere,  will be considered as a small positive variable. Definition 16 (Asymptotic small notation). The notation (xα) is interpreted O O as the convention (xα) lim O = 0 x 0 α → x for α > 0. Or in general terms (g(x)) (g(x)) ⇒ lim O = 0 O x 0 → g(x) for a decreasing function g on a right-open interval containing 0. The notation will be interpreted to indicate a Cauchy-null sequence possibly indexed by the Ox variable x. Definition 17. Define the parametrized difference operators acting on the function f(x) as + ∆ [f](x) := f(x + ) − f(x),

∆− [f](x) := f(x) − f(x − ) for the variable  > 0. The two operators are referred to as forward difference and backward difference operators, respectively. Definition 18 (Anonymous function notation). The notation for the pair µ ::  will be interpreted as the implication that if Left-Hand Side (LHS) is fixed then the Right-Hand Side (RHS) is fixed by the value chosen on the left, i.e. as an anonymous functional dependency  = (µ). Definition 19 (Dini derivatives). Define the Dini derivatives as the functions

∆± [f](x) D¯±f(x) = lim sup  0  → ∆± [f](x) D±f(x) = lim inf ¯  0 →  For the function f. Definition 20 (Baire categories). Let X be a metric space. A set E ⊆ X is of first category if it can be written as a countable union of nowhere dense sets, and is of second category if E is not of first category.

For example Q and ∅ are I category, while the class of continuous functions is of category 0.

Definition 21 (Baire function classes). The function f : R 7→ R is called Baire- class I if there is a sequence of continuous functions converging to f point-wise.

Definition 22 (Gδ and Fσ sets). Let X be a metric space.

• The set E ⊆ X is Gδ if it is countable intersection of open sets, and it is Fσ if it is countable union of closed sets. • The set E ⊆ X is meagre if it can be expressed as the union of countably many nowhere dense subsets of X. • Dually, a co- is one whose complement is meagre, or equivalently, the intersection of countably many sets with dense interiors. LOCAL GENERALIZATIONS OF THE DERIVATIVES ON THE REAL LINE 27

Appendix B. The First Oscillation Lemma The lemma was stated in [22]:

Lemma 6 (Oscillation lemma). Consider the function f : X 7→ Y ⊆ R . Suppose that I+ = [x, x + ] ⊆ X, I = [x − , x] ⊆ X, respectively. If osc+[f](x) = 0 then −f is right-continuous at x. Conversely, if f is right- + continuous at x then osc [f](x) = 0. If osc−[f](x) = 0 then f is left-continuous at x. Conversely, if f is left-continuous at x then osc−[f](x) = 0. That is,

lim osc±[f](x) = 0 ⇐⇒ lim f(x ± ) = f(x)  0  0 → → Then the negation of the statement is also true. Corollary 7. The following two statements are equivalent

lim osc±[f](x) > 0 ⇐⇒ lim f(x ± ) =6 f(x)  0  0 → → Proof. Forward case: Suppose that osc+[f](x) = 0. Then there exists a + pair µ :: δ, δ ≤ , such that oscδ [f](x) ≤ µ. Therefore, f is bounded in I+. Since µ is arbitrary we select x0, such that

|f(x0) − f(x)| = µ0 ≤ µ

and set |x − x0| = δ0. Since µ can be made arbitrary small so does µ0. Therefore, f is (right)-continuous at x. Reverse case: If f is (right-) continuous on x then there exist a pair µ :: δ such that

|f(x0) − f(x)| < µ/2, |x0 − x| < δ/2

|f(x) − f(x00)| < µ/2, |x − x00| < δ/2 Then we add the inequalities and by the triangle inequality we have

|f(x0) − f(x00)| ≤ |f(x0) − f(x)| + |f(x) − f(x00)| < µ

|x0 − x00| ≤ |x0 − x| + |x − x00| < δ

However, since x0 and x00 are arbitrary we can set the former to correspond to the minimum and the latter to the maximum of f in the interval. There- fore, by the least-upper-bond property we can identify f(x0) 7→ inf f(x), + f(x00) 7→ sup f(x). Therefore, oscδ [f](x) < µ for |x0 − x00| < δ (for the pair µ :: δ ). Therefore, the limit is osc+[f](x) = 0. The left case follows by applying the right case, just proved, to the mirrored image of the function: f(−x).  References [1] F. Ben Adda and J. Cresson. About non-differentiable functions. J. Math. Anal. Appl., 263:721 – 737, 2001. [2] F. Ben Adda and J. Cresson. Corrigendum to ”About non-differentiable functions” [J. Math. Anal. Appl. 263 (2001) 721 – 737]. J. Math. Anal. Appl., 408(1):409 – 413, 2013. [3] R. Bartle. Modern Theory of Integration, volume 32 of Graduate Studies in . American Mathematical Society, Providence, R.I., 2001. [4] E. I. Berezhnoi. The subspace of C[0,1] consisting of functions having finite one-sided deriva- tives nowhere. Mathematical Notes, 73(3–4):321–327, 2003. [5] J. Brown, U. Darji, and E. Larsen. Nowhere monotone functions and functions of non- monotonic type. Proceedings of the American Mathematical Society, 127(1):173–182, 1999. 28 DIMITER PRODANOV

[6] G. Cherbit. Fractals, Non-integral dimensions and applications, chapter Local dimension, momentum and trajectories, pages 231– 238. John Wiley & Sons, Paris, 1991. [7] P. du Bois-Reymond. Versuch einer Classification der willk¨urlichen Functionen reeller Argu- mente nach ihren Aenderungen in den kleinsten Intervallen. J Reine Ang Math, 79:21–37, 1875. [8] G. Faber. Uber¨ stetige Funktionen. Math. Ann., 66:81 – 94, 1909. [9] R. P. Feynman. Space-time approach to non-relativistic quantum mechanics. Reviews of Mod- ern Physics, 20(2):367–387, apr 1948. [10] R. Flood. Mathematics in Victorian Britain. OUP Oxford, 2011. [11] V. Fonf, V.I. Gurariy, and M.I. Kadets. An infinite dimensional subspace of C[0,1] consisting of nowhere differentiable functions. C. R. Acad. Bulg. Sci, 52(11–12):13 – 16, January 1999. [12] R. Girgensohn. An infinite-dimensional subspace of C[0,1] consisting of functions with no finite one-sided derivatives. Mathematica Pannonica, 12:129–132, 2001. [13] A. Gleyzal. Interval functions. Duke Math. J., 8:223 – 230, 1941. [14] Ida Kantor. Mathematics. American Mathematical Society, 2015. [15] K. Kolwankar and A.D. Gangal. Fractional differentiability of nowhere differentiable functions and dimensions. Chaos, 6(4):505 – 513, 1996. [16] K. Kolwankar and J. Vehel. Measuring functions smoothness with local fractional derivatives. Frac. Calc. Appl. Anal., 4(3):285 – 301, 2001. [17] B. Mandelbrot. Fractal Geometry of Nature. Henry Holt & Co, 1982. [18] S. Milanov, A. Petrova-Deneva, A. Angelov, and N. Shopolov. Higher mathematics , part II. Technika, 1977. [19] E. Nelson. Derivation of the Schr¨odingerequation from Newtonian mechanics. Phys. Rev., 150:1079–1085, Oct 1966. [20] L. Nottale. Fractals in the Quantum Theory of Spacetime. Int J Modern Physics A, 4:5047– 5117, 1989. [21] L. Nottale. Scale Relativity and Fractal Space-Time: Theory and Applications. Foundations of Science, 15:101–152, 2010. [22] D. Prodanov. Conditions for continuity of fractional velocity and existence of fractional taylor expansions. Chaos, Solitons & Fractals, 102:236–244, sep 2017. [23] D. Prodanov. Fractional velocity as a tool for the study of non-linear problems. Fractal and Fractional, 2(1):4, jan 2018. [24] D. Prodanov. Characterization of the local growth of two cantor-type functions. Fractal and Fractional, 3(3):45, aug 2019. [25] C E Silva. Invitation to ergodic theory. Student mathematical library. American Mathematical Society, Providence, RI, 2007. [26] W.F. Trench. Integral Calculus of Functions of One Variable, chapter 3, pages 171 – 177. Trinity University, 2013. [27] G. E. Uhlenbeck and L. S. Ornstein. On the theory of the Brownian motion. Physical Review, 36(5):823–841, sep 1930. [28] S Willard. General Topology. Dover Publications Inc., 2004. [29] X.-J. Yang, D. Baleanu, and H. M. Srivastava. Local Fractional Integral Transforms and Their Applications. Academic Press, 2015.

EHS, Kapeldreef 75, IMEC, 3001, Leuven, Belgium