THEE-315 3-611" ‘2' i “ '3 m m This is to certify that the thesis entitled ON THE SYMMETRIC DERIVATIVE presented by Lee Matthew Larson has been accepted towards fulfillment of the requirements for Ph. D. . Mathematics degree in CWQQM fly Major professor April 18, 1981 Date 0-7 639 i" ' l 'T 2w”! .' :3fffi \‘x’ y u\._ m. OVERDUE FINES: 25¢ per day per item RETURNING LIBRARY MATERIALS: Place in book return to remove charge from circulation records ON THE SYMMETRIC DERIVATIVE BY Lee Matthew Larson A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 1981 ABSTRACT ON THE SYMMETRIC DERIVATIVE BY Lee Matthew Larson A class of functions, 0*, is defined and is shown to contain all known symmetrically differentiable functions. It is proved that if f60*. then f is in the first Baire class. Using this result, it is shown that there is associated with each f€o* another function, g, which retains the symmetric differentiation properties of f while at the same time "maximizing" many of the more desirable properties of f such as differentiability, continuity and upper semi- continuity. Such a function. g, is in Baire class one and is uniquely determined up to its values on a set with countable closure. We call 9 the "nice copy" of {2 Using the properties of the nice copy, many of the standard theorems of ordinary differentiation can be refor- mulated in terms of the symmetric derivative. In particular, analogues of the mean value theorem and the Darboux property are presented. The methods also give simplified proofs of several well-known theorems. These results are then applied to develop an abstract Zahorski class structure for symmetric derivatives. In addition, several structure theorems for completely arbitrary symmetric derivatives are proved. ACKNOWLEDGEMENTS I would like to thank my parents, without whom I would not be, my wife, Ruth, who encouraged my dream-quest of unknown end and Professor C. E. Weil, whose patience was astounding and whose advice was invaluable. ii Introduction Chapter Section Section Section Chapter Section Section Chapter Section Section Section Chapter Section Section Section Section Bibliography I: 1.1: II: 2.1: 2.2: III: 3.1: 3.2: 3.3: IV: 4.1: 4.2: 4.3: 4.4: TABLE OF CONTENTS 0 O O O O O O O O O O O O O O O O O O O O 1 Notation. Definitions and Basic Theorems Notation O O O O O O O O O O I O O 0 O O O 5 Some Preliminary Function Theory . . . . . 12 A Covering Theorem . . . . . . . . . . . . 18 The Class of Arbitrary Symmetric Derivatives Comparison with Baire Class One . . . . . 21 Arbitrary Symmetric Derivatives . . . . 24 The Structure of Functions in 0* Nice Copies of Functions in 0* 28 Nice Copies of Functions in o . . . . . . 39 Monotonicity and Mean Value Theorems . . . 49 Symmetric Derivatives and the Zahorski Classes The Abstract Zahorski Classes . . . . . . 58 Symmetric Derivatives and the Class m2 . . 64 Symmetric Derivatives and the Class m3 . 67 Symmetric Derivatives and the Class M; . . 71 . . . . . . . . . . . . . . . . . . . . 79 iii INTRODUCTION If f is a real-valued function defined on IR” then the symmetric derivative of f at x (often called the first Schwarz derivative of f) is s _1im f baht-f Lx-h) f m" h-o 2h ' The symmetric derivative arises naturally in studies of the pointwise convergence of Fourier and Taylor series as well as other areas of harmonic analysis. In this work, however, we do not consider these applications of symmetric differ- entiation, but rather, we investigate the symmetric deriv- ative viewed as a generalization of the ordinary derivative. Specifically, our goal is to expose similarities between the well-known structure of ordinary derivatives and the struc- ture of symmetric derivatives. We begin in Chapter I by presenting much of the terminology used throughout this work and by stating the fundamental theorems needed in the succeeding chapters. In particular, we define a class of symmetrically differentiable functions, 0*, which is the "domain" for most of the later theorems. It is shown that 0* contains all measurable, symmetrically differentiable functions and therefore all known symmetrically differentiable functions, since the question of the measurability of such functions remains 1 unresolved. Chapter I is concluded with the proof of a 'partitioning theorem which was first stated in a slightly weaker form by B. S. Thomson [26]. One of the most useful theorems available for the study of ordinary derivatives, due to Zahorski [29]. is that any ordinary derivative belongs to the first class of Baire (81). It was proved by Filipczak [7] that the sym- metric derivative of an approximately continuous function is in 81. The main theorem of Chapter II is that this result can be extended to the more general case of 0*. In Chapter II, we also examine the question of whether there are any symmetrically differentiable functions which are not contained in 0*. While no answer to this question is reached, several results are obtained which strongly suggest that if any such function, f. exists, then fsis in $1. It is well-known that if f is a finite-valued ordinary derivative, then any primitive function for f is determined up to an additive constant. That this is not the case with a symmetric derivative can be seen by considering the following two functions. Let f(x)=o everywhere and let. x-2 for x=t1. il/Z. 9(X)= 0 otherwise Then it is easy to see that f8(x)=gs(x)=0 everywhere, but f(x)-g(x) is not constant. Because of this lack of a unique primitive. many of the standard theorems of ordinary differ- entiation are either false or much harder to prove with the symmetric derivative. A solution to this uniqueness problem is presented in Chapter III with the introduction of the "nice copy" of a function in 0*. The nice copy of f 60* is a.function, g. which in some sense "maximizes" several of the desirable properties of f such as differentiability and continuity, while at the same time retaining the symmetric differen— tiability properties of f. In particular, it is shown that there is a set. A, with countable closure, such that gs(x) agrees with f3(x) on A“: and further, that g is uniquely determined and upper semicontinuous on 54:. The existence of the nice copy for any f in 0* leads at once to the existence of a "nice primitive" which is uniquely determined up to an additive constant and its values on a set, A, with countable closure. This nice primitive solves the uniqueness problem presented above, and thus gives us a means of establishing many of the classical theorems of ordinary differentiation in terms of the symmet- ric derivative. For example, the quasi—mean value theorems of Aull [l] and Evans [5] and the monotonicity theorems of Weil [28] and Evans [5] can be generalized to 0*. Another consequence of the methods employed in Chapter III is a simplification of a proof due to Charzynski [4] showing that the set of discontinuities of any f ('0* such that f3 is finite-valued must be countable with no dense in itself subset. Finally, in Chapter IV, we extend the results of Zahorski [29] on the associated sets of derivatives to symmetric derivatives. In so doing, the results of Kundu [16] are considerably strengthened. In particular. it is shown that if f8 is the symmetric derivative of an f 6,0* such that f8 has the Darboux property. then £86 7/121. Kundu's theorems, in the cases of‘m3 and m4, are proved without his assumptions that f is continuous and f8 has the Darboux property. Examples are given to show that certain of the proved Zahorski class containments are proper with symmetric derivatives. CHAPTER I NOTATION. DEFINITIONS AND BASIC THEOREMS Section 1.1: Notation In this section we introduce most of the basic defin- itions and notation which will be used in later chapters. Several of the "classical" theorems concerning symmetric derivatives are also stated to motivate some of the new concepts. Throughout this work the real numbers will be denoted by IR and the extended reals, {-0, a], will be denoted by 111*. 2 will stand for the integers and 2+ will represent the positive integers. If ACJR is Lebesgue measurable, then the measure of A will be denoted by [A]. In fact, the only measure we shall have occasion to use is Lebesgue measure, so terms such as "measurable,“ "almost everywhere," etc., should be inter- preted accordingly. Ac will stand for the complement of A. Let f be a real-valued function defined on an open interval, I. If x 6 I, we define the upper (lower) symmet- ric derivative of f at x to he ?s (x)=1im suphdof (x+h)2-;lf jx-h) s =lim inf f( +h -f -h (_f (x) h~o x )2h (x ) ) ' 5 6 When fs(x)=§§(x). whether finite or infinite, we call their common value the symmetric derivative of f at x and denote it by fs(x). If fs(x) exists at every point of the domain of f. then f is said to be symmetrically differentiable. D+f(x°), D_f(x°). etc. stand for the Dini derivatives of f at x,: f+(x°). f'(x.) and f'(xO) denote the ordinary right, left and bilateral derivatives of f at x0, respec— tively. If both of the sums, D+f(x)+D-f(x) and D+f(x)+D_f(x), make sense, then it is easy to see that -:(D+f(x)+D-f(x)) s f‘(x) s fact) s %(D+f(x)+D_f(x)). Therefore, if both f+(x) and f‘(x) exist, then so does fs(x). and fs(x)=%(f+(x)+f-(x)) . Further, if f'(x) exists, finite or infinite, then fs(x)=f'(x). Thus, the symmetric deriv- ative is a generalization of the ordinary derivative. To see that it is a strict generalization, consider f(x)=[x] for which £3(0)=o, but f'(O) does not exist. f is said to be symmetrically continuous at xo‘if lim (f(x°+h)-f(x°4h))=0. h~0 As usual, a function which is symmetrically continuous at each point of its domain is called symmetrically continuous. It is clear that if fs(x.) exists and is finite, then f is symmetrically continuous at x0. It easily follows from the definitions that if f is continuous at x,. then f is also symmetrically continuous at x.. That the converse is not true can be seen from the function f(x)=cos% which is symmetrically continuous (even symmetrically differentiable) at x=O. but certainly not continuous there. Therefore, just as symmetric differen- tiability is an extension of ordinary differentiability, so is symmetric continuity an extension of ordinary continuity. We shall denote, for any function, f. D(f)=[x: f'(x) exists and is finite] and C(f)=[x: f is continuous at x ]. The following proposition will prove useful in later chapters and will be employed without constant reference to this section. A proof of it may be found in [12]. Proposition 1.1. Let I be an interval and f a function defined on I. Then C(f) is a G6 set. Let A CR and x be a limit point of A. If there exists at least one sequence from A increasing to x, we define A-lim sup f(t)=lim sup [f(t): t€(x-6. x)nA] trx- 6~0 and A-lim inf f(t)=lim inf [f(t): t€(x-6, x)flAQ,. tdx- 6-0 If both of the above limits agree. their common value will be denoted by A-lim f(t). The right-hand limits through tex- A are defined analogously. The meanings of A-lim sup f(t). t-x A-lim f(t). etc.. are now Obvious. If, in the above def- trx initions, A=nl, then it is omitted from the above expressions to conform to standard notation. Suppose A CR and x0619. . We denote the reflection of A through x° by Rx (A). For example, 31([O,4))=(-2,2]. 0 Further suppose that I is an open interval and f is a 8 function defined on I such that C(f) is dense on I. Let x061 and [biz i€Z+] be a sequence of positive numbers dec- reasing to 0 such that (xo-éi,x°+6i)CI for all iEZ+. Choose any kEZt Since any 65 set which is dense in an interval is residual in that interval, we see that C(f) is residual in both (xo-é . x0) and (x0, x°+6k). It is then k clear that R.X (C(f)n(x,-5 . x,)) is residual in (x,. xo+6kl O k and thus C(f)f]R. (C(f)f](x°-5 , x°))7=’¢. xo k Choose xk to be an element of the above intersection and let ykéflx°(xk). This procedure can be followed for each kEZ+ to generate two sequences, [xiz i€z+] and [y1 i€Z+], which satisfy (1) Rxo(xi)=yi for all i€Z+, (2) limn‘cxne-limn‘cyn=xo and (3) [xi: i€Z+]U[yi: iEZ+}CC(f). Given a function, f, we say two sequences, [xi: i€Z+] and [yi: i€Z+}, satisfying (1)-(3) converge C(f)-symmetric- ally to x,. From the above considerations, the following proposition is clear. Proposition 1.2. Let f be a function defined on an open interval, I, such that C(f) is dense on I. Then each xoél has a pair of C(f)-symmetric sequences converging to it. With f and I as in the proposition, we define fsc(x°) to be 9 f (yn) -f (xn) n-OO —x yn n lim if the limit exists and is the same for all C(f)-symmetric sequences, [xn: nEZ+} and [ynz n€Z+}, converging to x,. For example, if f3(x°) exists and C(f) is dense in a neighborhood of x0. then £s°(x°) exists and equals £3(x°). The following theorems, which motivated several of the definitions given above, are fundamental to the results in Chapters II and III. Theorem 1.3. (Fried [10]) Suppose the set of points at which f is symmetrically continuous is residual on an open interval, I. Then C(f) is also residual on I. Theorem 1.4. (Preiss [21]) If f is symmetrically contin- uous on an interval, I, then it is continuous a. e. on I. Theorem 1.5. (Khintchine [13]) Let f be a measurable function defined on an open interval, I. Then f has a finite ordinary derivative at almost all points for which _f_8 (X) >-°. Suppose f is a function defined on an open interval, I: SUCh that fs(x) is finite everywhere on I. Then f is symmetrically continuous on I and theorems 1.3 and 1.4 both imply that C(f) is dense. Further, if f is measurable on I and symmetrically differentiable (infinite values allowed), then by considering f and -£ we see from theorem 1.5 that f has a finite ordinary derivative a. e. and thus C(f) is again dense. The common thread which seems to bind 10 all three of these theorems is that if f is a reasonably behaved function which is symmetrically differentiable, then C(f) is dense. This observation motivates the following definition. Definition. Let I be an open interval. Define 0*(I) to be the class of all functions, f, such that C(f) is dense on I and fs(x) exists, finite or infinite, everywhere on I. Define 0(I) to be the class of all functions, f60*(I), such that fs(x) is finite at each x61. Analogously, we denote by A*(I) the class of all functions, f, such that f'(x) exists everywhere on I and by A(I) the class of all functions, f€A*(I), such that f'(x) is finite everywhere on I. Using the above notation, we write 0*s(I)=[fs: f€0*(I)} and A*'(I)=[f‘: f€A*(I)}. 05(1) and A'(I) are defined similarly. In order to make the notation slightly less cumber- some, we denote 0*(1R), A*(R), etc., as just 0*, A*, etc.. Most propositions will be stated with this simplification, it being clear in all such cases that the restriction of the statement to an arbitrary open interval is valid. As a further notational convenience, if I and u happen to be classes of functions, we denote Inn by In. The following corollaries are easy consequences of the definitions and the three theorems. ll Corollary 1.6. Let f be a measurable function such that fs(x), finite or infinite, exists everyWhere. Then (a) [[x: |£s(x)1=e}]=o (b) f'(x) exists and is finite a. e. (c) f€0* . Proof. By applying theorem 1.5 to f and -f, we see that th: £3(x)=e}]=o and [[x: fs(x)=—~}1=o and (a) follows. (a) and another application of theorem 1.5 yield (b). Since f'(x) exists and is finite a. e.. it follows that f is continuous a. e. so C(f) is dense and f60*. Corollary 1.7. £60 if and only if fs(x) exists and is finite everywhere. In this case, f is measurable. Proof. If £60, then fs(x) exists and is finite everywhere by the definition of 0. On the other hand, if fs(x) is finite everywhere, then f is symmetrically continuous and theorem 1.3 implies that f is continuous a. e.. Thus, C(f) is dense and £60. We note that any function which is con- tinuous a.e. is measurable. Putting these results together, we see that if i is the class of all measurable symmetrically differentiable functions, then ACOC£C0*. The question of whether there are any symmetrically.differentiable functions which are not measurable is unanswered and will be examined further in Chapter II. Using a theorem of Zahorski [29] that any f€A* is discontinuous on at most a countable set, we see that ACA“ C£C0*. 12 Section 1.2: Some Preliminary Function Theory Let f be a function defined in an open interval, ICIR, taking on values in IR*. f is said to be in Baire class one if there exists a sequence, [fn: n€Z+], of functions continuous on I such that f(x)=limnd.fn(x) for each x61. The class of Baire one functions defined on I is denoted by 251(1) . If I=1R , we write 81(I)=31. Following is the fundamental theorem characterizing 81. Proofs of it can be found in Goffman [12] or Natensen [20]. Theorem 1.8. The following statements are equivalent: (a) £681: (b) For all sent, the sets [x: f(x)$a] and [x: f(x)2a] are 35 sets: (c) For all aent, the sets [x: f(x)a} are FG sets: (d) If P is a perfect subset of hi, then the restriction of f to P has a point of continuity. The primary importance of the class of Baire one func- tions lies in the fact that it contains all ordinary deriv- atives. If f is a function such that f'(x) exists and is finite everywhere, then evidently f is continuous, and it is easy to see that f'éfil. If infinite derivatives are allowed, the situation is not as clear because f need not be continuous. For proofs that f'EEl, even in this case, see Zahorski [29] or corollary 2.6 of this work. In fact, 13 in Chapter II, we will prove that 0*SC31. A function, f, defined on an interval, I, is said to have the Darboux (intermediate value) property if whenever x and y are in I and o is any number between f(x) and f(y), then there exists a number, 2, between x and y such that f(z)=o. We shall denote the class of all functions defined on I which have the Darboux property by 3(1). As usual, .DUR) is just written as .0. We shall rarely have use of fi’by itself, but rather, will make much use of the properties of the class 381. There are more than a dozen known ways of characterizing £31. The following theorem contains the ones we will need. Theorem 1.9. Let féfil. The following are equivalent: (a) £6831: (b) For each xénl there is a sequence, [yn: n€Z+], increasing to x, and a sequence,[zn: n€Z+}, decreasing to x, such that llmnd°f(yn)=llmnflof(zn)=f(x): (c) For each x6111 , f(x)€[lim inft _f(t), lim supt_x_f(t)] n "X n [lim inf x+f(t) , lim suptfl+f(t) ] . t-o For a proof of theorem 1.9, as well as many other of the characterizations of 3%1, see Bruckner [3, p. 9]. It is well-known that the derivative of a continuous function has the Darboux property. That the same is not true of the symmetric derivative can be seen from the function f(x)=lxl, where fs(x)=1 for x>O, fS(O)=O and 14 fs(x)=-1 when x- everywhere and §§(x)20 a. e.. Then f is nondecreasing. Evans [5] extended this theorem from 381 to the larger class of all measurable functions, f, satisfying (1) lim inft‘xf(t)$f(x)slim supt‘xf(t) at each x. (Note the similarity to theorem 1.9(c).) In fact, he showed that this class is the largest class of measurable functions for which a statement like theorem 1.11 is true. For our purposes, we will need a slightly more general version of theorem 1.11. Theorem 1.12. Let f be a function satisfying (1) such that C(f) is dense, §§(x)>-an everywhere and §§(x)20 a. e.. Then f is nondecreasing. Proof. We pattern our proof after that of Weil [28]. First, let f§(x)>0 everywhere. Suppose f is not nondecreasing. Then there exist a0 and b0 with a°f(b°). Choose any o€(f(b°), f(a°)) and define Ea=[x€[a°, b0]: f(x)sa] and Ea=[x€[a°, b0]: f(x)2a]. Suppose that neither Ea nor Ea contains an interval. Then, if x€(a°, b°)nC(f), it is clear that f(x)=o. According to proposition 1.2, if x°€(a°, b0) we can choose C(f)-symmetric sequences, [xn: n21] and [yn: n21], converging to x0. Then _f_s(x°) slim f(xnl’fwn) =lim ‘3‘“ =0 n“ 11-90 x - - n yn xn yn which contradicts our assumption that §§(x°)>0. Thus, 16 either Ea or Ea contains an interval. Suppose, for example, that Ed contains an interval. (If E“ contains an interval, the argument is similar.) We can then choose an interval (c,d)CEa such that (2) c= inf [x: (x,d)C:Ea] . c>a°, for otherwise (2), (1) and a0, it follows that there is a 6 with O<6O and define f€(x)=f(x)+cx. Then £:(x)2e>0 everywhere and f6 satisfies (1), so according to the above argument, fe is nondecreasing. Since fe is nondecreasing for every c>O, we can take the limit as c~O to see that f is nondecreasing. Finally, let f be as in the statement of the theorem. In Zahorski [29], it is shown that for any set, A, such that [A]=O, there exists a continuous and nondecreasing function, g, which is differentiable everywhere and for which g'(x)=~ whenever x€A. Let e>O and A=[x: §§(x)<0]. Since ]A]=O, there is a function, g, as above. Define fe(x)=f(x)+cg(x). Then, clearly, §:(x)20 everywhere and fe(X) satisfies (1), so fe is nondecreasing. Again, letting e~O shows that f is nondecreasing, and the theorem follows. 17 Following Evans [5], we define the class m-l to consist of all functions, f, satisfying (1) such that C(f) is dense. Then from theorem 1.12, the following is clear. Corollary 1.13. Let f€M_l0* such that fs(x)>- everywhere and fs(x)20 a. e.. Then f is nondecreasing. A function, f, defined on an open interval, I, is upper semicontinuous at xEI if 1km suptdxf(t)sf(x). f is lower semicontinuous at x if -f is upper semicontinuous at x. If f is upper (lower) semicontinuous at each point of its domain, then it is said to be upper (lower) semicontin- uous. Note that this definition appears at first glance to be slightly different than that in some common books because of the way we defined the upper and lower limits of a function. The following theorem shows that our definition is the same. Theorem 1.14. Let f be a function defined on an open interval, I. The following statements are equivalent: (a) f is upper semicontinuous: (b) For each xEI, f(x)21im supt‘xf(t); (c) For each sent, [x: f(x)2a} is closed relative to I: (d) For each a632, [x: f(x)O associated with each xE(a, b). For each x, we define Sx=[[x-h, x+h] : O6(O). By the definition of a, for every c>O, Dh(a-c, a] is countable and 5h(c,a+e) is uncountable. But, if xEDcn(a-6(o), a), then C-contains a partition of [80(Rd(x)), Rd(x)], from which it follows that ENG, d+6(o))CRa(Dn(o-o(o) , 01)) , which is countable. This contradiction shows that o=b. Now, let x°€(0, b) and let -x°=ao-a and <1) (x: fs(x)=a]=(x: 93(x)=0}. Choose a 6>O. Using proposition 1.2, we see that for each xéni, there is an h€(O, 6) such that x+h and x-h are both elements of C(g). From this observation, it 21 22 makes sense to define A=[x: sup00, where x+hEC(g) and x-h€C(g)]. If x€A, then there are h, o and 6, each positive, satise fying the following inequalities: (2) x-h, x+h€C(g)n(x-o, x+5); (3) g(x+h)-9(x4h)>2a: '(4) lX+h-ylO with [x-x,]+BO such that x,-‘h'€C(g)flRl and x°+h'€C(g)nR2. From (7) and (8) it follows that h'>o so that xoeA. Since the only requirement on xo was that [x-x°]O such that (x—e, x+e)CA. Therefore, A is open. 23 Similarly, if we define, for each n€Z+, the set An to be [x: sup 19(x+h).g(x.h)>—%?-where x+h€C(g) and x-hEC(g)]. 00 for OO 1 . because yn xn0 for each n. This implies that fs(a)20, because f is symmetrically differentiable at a. But, fs(a)sa<0 because aEA. This contradiction shows the supposition to be false, so both A and B cannot be dense in I. Theorem 2.4 shows that the associated sets for an arbitrary symmetric derivative behave much like the assoc- iated sets for a function in Baire class one. In view of theorem 2.1, this is hardly surprizing. Theorem 2.5 rules out another form of pathological behavior for symmetric derivatives. Theorem 2.5. Let f be any function. Then [x: lfs(x)]=°] contains no interval. 26 Proof. Suppose, to the contrary, that there is an interval Ic[x: lf3(x)]=°]. According to theorem 2.4, both of the sets, A=[x: f8(x)=-¢] and B=[x: fs(x)=¢], cannot be dense in I. So assume there are a,B€I such that aO, there is a 6(x,p)>O such that if O2hp. For each n€Z+, define (10) Jh=[[x-h, x+h]: x€(a, B) and OZIE=1(Bi-oi)n 2xn. Since n is arbitrary and x>O, (14) clearly leads to a contradiction of (13). Thus, B can contain no interval. Similarly, A can contain no interval. 27 I * Corollary 2.6. A :81. Proof, According to theorem 2.5, the set on which f' is finite is dense. Whenever f'(x) is finite, f is continuous at x, so C(f) is dense. Now apply theorem 2.1. Corollary 2.6 was apparently first proved by Zahorski [29, p. 14]. CHAPTER III THE STRUCTURE OF FUNCTIONS IN 0* Section 3.1: Nice Copies of Functions in 0* The primary goal of this section is to prove the following theorem. Theorem 3.1. Let f60*. Then there are two sets, A1 and A2. each with countable closure, and two functions, 91 and g2, each in Baire class one satisfying: (a) g:c(x)=fs(x) everywhere, i=l,2; A“ (b) g:(x)=fs(x) everywhere on i' i=1,2; (c) 91 (92) is upper (lower) semicontinuous on A: (A3); (d) C(f)CC(gi) and f(x)=gi(x) for each xEC(f), i=1,2; (e) D(f)CD(gi) and f'(x)=gi(x) for each x€D(f), i=l,2: (f) If I is a component of Ag, then gi€m_1(I), i=l,2. The proof is accomplished with the aid of the following series of lemmas, some of which are interesting in their own right. Lemma 3.2. Let I be an open interval, C a dense subset of I and f any function defined on I. Define 28 29 u(x)=C-lim suptdxf(t) and.£(x)=C-lim.inft‘xf(t). Then u is upper semicontinuous and t is lower semicontinuous. Proof. Let den: and A=[x€I: u(x)2a]. If A=¢, then A is closed. Otherwise, we may choose a sequence, [xnz n€Z+]CA such that 1imnflaxnex. From the definition of u, for each + I C I 1 1 n62 , there is a tnGC satisfying [tn-xn]a. Then, clearly 1imhfiatn=x and . . . 1 _ C-lim.suptdxf(t)zlim.supn~°f(tn)zlim|supnqa(a—E)-a so u(x)2a and xeA. Thus A is closed and it follows that u is upper semicontinuous. The proof that z is lower semicontinuous follows by noting that -L is u for -f. Qemma 3.3. Let f, C, u and 1 be as in lemma 3.2. Then (a) C(f)CC(H) (C(f)CC(£)) and f(X)=H(x) (f(X)=¢(X)) for each x€C(f). (b) D(f)CD(H) (D(f)<'-'-'D(£)) and f'(x)=u'(X) (f'(X)=£'(X)) for each x€D(f). Proof. We may suppose without loss of generality that OEC(f) and f(O)=O. Then, given an e>O, there is a 5>0 satisfying [f(h)]O, there is a 6>O such that when ]h[<5, [f(h)]O such that (“(x)]<¢ whenever [x]O, there is a 66(O,n) such that when O-f(-t)l<2te. Fix a t6(0,6) and choose a sequence, (Sn: n€Z+]CC(f), such ‘0 that limn sn=t, O0 such 1 that when [sn-x]2tc. Fix a t€(O,5) and choose a sequence, (Sn: n€Z+]CC(f)n(-6,0) such that limnfiasn=-t and 1imnn°f(sn)=C(f)-lim supx‘tf(-x)=u(-t). In the same manner as above, we may choose a new sequence, (tn: n€Z+]CC(f)n(-6,0) such that 30([tn: n62+])CC(f), 32 1imnd°tn=-t and (8) 1imn”.f(tn)=u(-t). Then, using (8) and (7), u(t)-u(-t)=C(f)-lim.supxatf(x)-C(f)-lim.sudetf(-x) =C(f)-1im supx‘tf(x)-limnqaf(tn) (9) 21in supn‘;f(-tn)-limn‘°f(tn) 21in supnd¢(f(-tn)-f(tn)) >lim supnd¢2]tn]o=2to. By choosing a arbitrarily large in (9), we see that uS(O)30. The case when fs(O)=-° succumbs to a similar argument. Therefore, us(O)=fs(O). The assertion that 15(x°)=fs(x°) follows by noting that -L is u for -f. It should perhaps be noted that some condition such as requiring u and J to be finite in a neighborhood of x0 is necessary. To see this, for n€Z+, let . -n -n -1 El? 4 ”(€32 ) x6(2-n_4-n'2-n+4-n)_[z-n} 0 otherwise and fn(x)= Zn=lfn(x)+Z:=lfn(-x) . Then £60. but u(x)=°=-£(x) whenever x€[t2-n: nEZ+], so the difference quotients u(h)-u(-h) £(hl:£(-h) 2h and 2h are undefined whenever h=2-n for some n€Z+. Thus, us(o) and 13(0) are also undefined. 33 Lemma 3.5. Let f€0* with A=[x: lim.suptdxf(t)=¢] and B=[x: lim.inft‘xf(t)=-¢}. Then both A and B are countable closed sets. Proof. we prove the lemma in the case of A. The assertion for B then follows by considering -f. Using c=rz in lemma 3.2, we see that A is closed and because C(f) is dense, A.must be nowhere dense. Being closed, A may be written as A=PUN, where P is perfect and N is countable. Suppose P¥¢ and let (d,B) be a component of PC. (a or B could be infinite.) Since P#¢, a or B must be finite. Suppose B is finite. Then, since P is closed, BéP. P being perfect and (d,B)CPc, we see that for each O>Oo [5.3+6)OP is uncountable. Since An(o,B)CN is countable, we may choose a sequence, (Bu: n€Z+]CP, such that Bn decreases to B and RB({Bn: n€Z+])nA=¢. Using the facts that BnGA for each n and lim supt#RB(Bn)f(t)€[-°o“) we may choose a tn>B for each n62+ such that [tn-Bnl<% and f(tn)-f(Ra(tn))>n. Clearly, limnflatn=B and f(tn)mf(fic1(tn)) lim inf n 1im inf _____=. n”‘ 2(tn-B) 2 n‘°2(tn-B) so fs(B)=“. Similarly, if a>-, then fs(o)=-. Now, we note that c o P - Un=1(an.Bn) where (on,Bn) is a component of Pc for each n. Clearly, then since P is a nowhere dense set in hi, the sets 34 (on: n€Z+] and (Bu: n€Z+] are both dense in P. This implies, from the above, that the sets I+=[x: fs(x)=°] and I-={x: fs(x)=-°] are disjoint dense subsets of P. P, being closed, is a G set and according to theorem 6 2.1, I+and I- are 66 sets, so POI+ and POI- are dense Gd subsets of P. As such, both PnI+ and PnI' are residual in , + _ P. Since P is a Baire space, POI OI ¥¢, which contradicts the fact that I+nl'=¢. Therefore, we conclude that P=¢ and A=N, a countable set. Enough machinery has now been developéd‘to accomplish our primary goal. Proof. (Theorem 3.1) Let f€0*, u and 1 be as in lemma 3.4 and A and B be as in lemma 3.5. Define Al={x: lu(x)]=°l and A2=ix= l£(X>l=°} and let u(X) xEAi £(x) xeag 91(X)= and 92(x)= f (x) x€A1 f(x) xéA.2 Since A1UA2CAUB, the countable closure of both A1 and A2 follows from lemma 3.5. (b) follows from lemma 3.4. (c) follows from lemma 3.2. (d) and (e) follow from lemma 3.3. (a) follows from the definitions of u and L and (d). The rest of the theorem will be proved in the case i=1, the proof in the case i=2 being similar. Choose an aénl. The upper semicontinuity of 91 on 35 A: and theorem 1.14(c) imply that for each n€Z+, - f 0 l En-[xeAl . gl(x)za+I-1-] -c is closed relative to A1 and so is an F0 set relative to IR . Since —c - O F1=[x€A1 . g1 (x)>a]— Un=lEn' we see that F1 is also an F0 set relative to It. It is clear that (10) F2=[x€A1: gl(x)>a] is an F0 set, because A1 [x6112 : 91 (x) >a]=F1UF2 is countable. Thus, is also an F0 set. Similarly, the upper semicontinuity of g1 implies that - "C, F3-[x6A1. 91(X)lO. It suffices to show that the set Mn(-r,r) satisfies the lemma. To do this, we define Y(x)=f(x)-g(x). Since fs and g5 are finite everywhere, Ys(x)=0 everywhere, so by corollary 1.7, Y60 and C(Y) is dense. We may suppose without loss of generality that O€C(Y) and Y(O)=O. We will show that c=O satisfies the lemma. Let e>O. Because OEC(Y), there is a 61>O such that When O$]h]<61. [Y(h)]<§u Since Ys(x)=0 everywhere, for 40 each xEIl, there is a 6(x)6(0,61) such that when O llmn-«IDZer1 - 2 This contradiction shows that M has no nonempty dense in itself subset. A function, f, is said to be symmetric if for every x6}! there exists a 6(x)>O such that f(x+h)=f(x-h) whenever O0 such that f(x)=c when OO such that if O-f(yn>l<1. Using the fact that C(f) is residual, we may then choose, for each n, a new zn68x0((yn-6n,yn+6n)nC(f))flC(f). Then, using (27) and (28) C(f)-11m supt“R (y°)f(t)zlim supnfiaf(zn)2 x0 211m supnq¢(f(flxo(zn))-l)21im supnda(f(yn)-2)=o 44 so that Rx°(y°)€Al. A similar argument shows that if C(f)-1im supt‘y°f(t)=-". then Rx°(y°)EA1 once again. Therefore, axe(Aln(x°-6,xo+6))=Aln(x°-6,x°+6). Since x. was chosen arbitrarily, it follows that A1 is a symmetric set. Again, let xoenl. ‘We may, without losing generality, assume fs(x°)=0. Then, for each e>O there is a 6>0 such that whenever O-. Since x6A, either C(f)-1im suptdxg(t)=9, or C(f)-lim.suptdxg(t)=-. Assume the former. Using (49) and (50), we see (51) C(f) -1im suptdx+g(t)=a<° and (52) C(f)-1im suptdx_g(t)=°. According to (52), we may choose a sequence, [xn: n6z+]CC(f) such that xn increases to x and limnd¢g(xn)=°. Since C(f)CC(g) by theorem 3.1(c), for each n6Z+ there is a 1 . . 6n6(0,n) such that Iy-xn]<6n implies that Ig(y)-g(xn)]0 for each x6(a,B) and applying corollary 3.18, we see that F is strictly increasing on (0,B). Since 0,B6C(f), we see from theorem 3.1(c) that F is strictly increasing on [6,5]. This implies that F(G)- for all x€(a,B), then ]A(>o. Proof. If we proceed as in the proof of theorem 3.22 and assune that [A]=O, we arrive at a contradiction in the same way via theorem 3.18. Corollary 3.24. Let £60 with a,B, A and B as in theorem 3.22. Then both A and B have positive measure. Proof. Apply corollary 3.23 to f and -f. Propositions of the type of theorem 3.22, corollary 3.23 and corollary 3.24 are often called quasi-mean value theorems. Theorem 3.22 was apparently first proved by Aull [l] for continuous functions. It was later extended by Evans [5] and Kundu [14] to functions satisfying certain monotonicity conditions. As mentioned above, theorem 3.20 cannot take on the 54 form of the usual mean value theorem because fs may not satisfy the Darboux condition. For example, let f(x)=[x], =-l and B=2. Then f(B)-£(Q1_=l, B-a 3 and fSUR) =[-1,0,l] . However, if £863, then we do arrive at the usual statement of the mean value theorem. Corollary 3.25. Let f60* such that f3 has the Darboux property. Suppose o,B6C(f) such that ocx>0>lim.suph_.O 2 It is clear that f60S implies that f(--x)60s and g (x) =f (x)-I2-f ( -x) 608 . (57) may be rewritten as (58) 9(0)>a>0>lim sushd09(h). 55 (58) implies that there is a 6>0 such that g(h)<0 whenever O<[h]<6. By corollary 3.20, there is a primitive, G, for 9 such that G is continuous and decreasing on (-6,6). This implies that g(0)=GS(0)$O, which contradicts (58). Thus, we conclude that (57) is impossible and the rightehand inequality in (56) is true. The leftehand inequality is established similarly. Corollaryg3.27. Let f60*s and F be a nice copy of a primitive for f. Then C(f)CD(F). Proof. Let x6C(f) and e>O. Then there is a 6>0 such that [f(x)-f(y)]<§-whenever [y1s6. Let -60](](a,b)]. Proof. According to corollary 3.20, F is continuous. For each nez+ and each xerz, define l (59) Fn(x) n(F (xi-H) -F (x) ) . Fn is continuous for all n62+ and if x6D(f), then (60) limnflFn(x)=f(x) . According to corollary 3.28, (60) is true a. e.. Now 1im inf fbr (x)dx=lim inf nj‘b(r(x+l)-r(x))dx= 1'1"O a n 1 11"“. a n b+- =1im inf n(j' nF(x)dx-j“"l~(x)dx)= nee +l_ ‘ a _ n a+£ (61) =llm infn4.n(fb nF(x)dx-j‘a nF(x)dx)2 b+l- 1 al— 211m infnflcnj‘b nF(x)dx--lim supndcnja nF(x)dx= =F (b) -F (a) by the fundamental theorem of calculus. Using theorem 3.22 with (59) it follows that 57 Fn(x)$M for all n and x. Define, for each n6Z+, cn(x)=max(Fn(X) . 0) . Then, Gn is continuous for each n with O‘Gn(x)SM and from (60), limndaGn(x)=max(f(x), 0) a. e.. Applying the dominated convergence theorem, we see ° ° b x)dx 5 lim b lim inf ”[ F ( a G (x)dx: (62) nor“. -. " f ah,“ f(x,>o}f(x)dx ‘MI (x: f(x)>o)n(a.b)] A combination of (61) and (62) yields the lemma. CHAPTER IV SYMMETRIC DERIVATIVES AND THE ZAHORSKI CLASSES Section 4.1: The Abstract Zahorski Classes In 1950, Z. Zahorski [29] began a classification of derivatives based upon the structure of their associated sets. In the course of this work, he defined a descending sequence of subclasses of £51 which he called mi. i=0, ...,5. If we represent the classes of functions which are, res- pectively, approximately continuous, bounded in A' and both bounded and approximately continuous by an bA' and bdfi then Zahorski's conclusions can be represented schem- atically. ”(013ml 3 7”2 3’ 7”32’ 7724. D 77" ‘7 (l) U U U U U .681 23A“ :A' 3166' 31:4 5: Kundu [16], in 1976, defined abstract Zahorski classes and succeeded in demonstrating a similar structure for continuous functions, f, such that £863. In the following sections, we will extend Kundu's theorems to larger subclasses of 0*8. Definitions . Let ACIR . MO(A) is the collection of all F0 sets, F, such that 58 59 for all xeAnF, x is a bilateral limit point of F. M1(A) is the collection of all FO sets, F, such that for all x6AflF, x is a bilateral condensation point of F. M2(A) is the family of all F0 sets, F, such that for all x6AflF and all 6>0, [(x-6,x)nF[>O and l(x,x+6)]>0. M3(A) is the collection of all F0 sets, F, such that if xeAflF and (In: n6Z+] is any sequence of closed inter- vals converging to x (i. e., any neighborhood of x contains all but a finite number of the In) such that InnF=¢ for all n, then . I 1 lim 1 n =0. n“ d(x,In) M4(A) is the collection of all FC sets, F, such that there is a sequence of closed sets, [Fn: n6Z+], and a sequence of numbers, (nu: n6Z+]C(O,1), such that F= n=1Fn and for every c>0 and any x6AflFn there is an c(x,c)>0 such that for any two real numbers, h and hl, satisfying hh1>0, hnn is true, where J is the interval with endpoints x+h and x+h+h1. M5(A) is the collection of all F0 sets, F, such that for all x6AnF, x is a density point of F. For i=0,1, ...,5, define the abstract Zahorski class, mi(A), to be the collection of all functions, f, such that for any aEIZ, [x: f(x))a] and [x: f(x)O and each sequence, (In: n6z+], of closed intervals conver- ging to x such that for each n, f(y)2f(x) on In, or f(y)Sf(x) on In. liyEIn= if (y) -f (x) (26H n-m lIn]+d (onn) lim =0. If A=nl in any of the above definitions, we omit the reference to A: e. g., MiUR) =Mi' 7711f!!!) =7/(i and Z(JR)=Z. It follows easily from the definitions that for any ACJR , Mi+1 (A)C-'Mi(A), i=0, ,4. Therefore, ”(1+1 (A)(:77(i (A), i=0, ...,4. The following lemma is less obvious. Lemma 4.1. Let ACJR . Then Z(A)C?Iz3 (A) . Proof. Let f62(A) and a6nl. It must be shown that the sets [x: f(x))a] and [x: f(x)0 and a 6>0 such that fs(t)>fs(x)+6 when t6(x,x+€). Through the addition of an appropriate linear term to f, we may assume (6) fs(x)f(x) on (x,x+e), it is easy to see that there is an n>0 such that f(t)>f(x) on (x-n,x). Therefore, (10) D‘f(x)so. (3). (9) and (10) imply that D+f(x)=o‘f(x)=o. D'f(x)=0 implies that there is a sequence, [xn:n6z+}, increasing to x such that f (x) -f (xn) 1' =0. J’mn-oan x-xn D+f(x)=0 implies that . . f(ax(xn))-f(x) lim infn“ x-xn -20. Now, consider, f0 f(x)=' . 0 x=0 x . l 3 Sinx+x x<0. It is clear that f has a finite derivative on (-,O)U(O,°), so there is no problem with angularity or the Darboux prop- erty on either of these intervals. It is also easy to show that fs(0)=£u so f60. Since f is continuous and g; 8111;" - --]’—-2 cos}; x>0 3x f'(x)= %—sin!’--—1§cos}-+ l x<0, 3x x theorem 1.9(c) can be used to see that £6361. But, D+f(0)=%-and D_f(0)=%-shows that (2)is violated and f is angular at 0. Section 4.2: Symmetric Derivatives and the Class m2 The main purpose of this section is to prove the following theorem. Theorem 4.3. £0*sdm2. To prove this theorem, we use the following lemma, which should be viewed in light of theorem 3.18. Lemma 4.4. Let f6£0*s with F a primitive for f. If f(x)20 a. e., then any nice copy of F is nondecreasing. Proof. Without loss of generality, we may assume that F is the nice copy of itself. It then suffices to show that F is nondecreasing. To do this, according to theorem 3.18, 65 we must show that A=[x: f(x)=-°] is empty. Define B=[x: f(x)$-l] and C=[x: f(x)=-2]. Since f6£0*s, theorem 2.1 implies that £6361 from which it follows, using theorem 1.8(b), that A, B and C are G6 sets. We claim that A is relatively dense in B. To see this, suppose it is not. Then there is an open interval, I, such that IflA=¢ and IDB¥¢. An application of theorem 3.18 shows that F is nondecreasing on I, so IOB=¢, a contradic- tion. Therefore, A is relatively dense in B. We now claim that C is also relatively dense in B. To see this, let x63. Since A is relatively dense in B, we may choose a sequence, [xnz n6Z+]CA, such that lim x = ° . , 2 . nae n x. By assumption, the set (x. f(x) 0] is dense in It, so we may choose for each n6Z+, a yn such that [xn-ynl<%»and f(yn)20. Since £633 for each n62+, there is a 2 between x and y such that f(z )=-2. Because n n n n 1 . . _ [zn—xn]<]yn-xn]0]. 66 Suppose there exists an x6A and an c>0 such that (11) l(x,x+e)nA]=O. Then f(x)$O a. e. on (x,x+e). By the lemma, it follows that f(t)$0 everyWhere on (x,x+e). Now, theorem 1.9(b) implies that f(x)‘0 so that x6A. This contradiction shows that (11) is false for every x6A and every e>O. In a similar manner, it can be shown that [(x-e,x)fiA]>O for each x6A and each e>0. ' Therefore, A6M2. Through the addition of an appropriate linear term to a primitive of f, it may be shown that [x: f(x))a]6M2 for any a6[-¢,°). By considering -f, we see that [x: f(x)§ rn(-x)+x x6RO(In) It is an easy calculation to show that f is differentiable on (-°,0)U(0,°) with [f'(x)]SB whenever x#0. It is also 68 evident from the symmetry of its definition that fs(0)=%. Therefore, f60. For each nEZ, rn(x) attains its maximum value on In at 2 . From this, we see that 3n+1 + 2n+l 2 . 1 D “0) = llmn-oO—2_—r2n(—2h:—I)=lmn~¢(l "23) = 1 3 3 and similarly, D+f(0)=-1. From this and the definition of f, it follows that D-f(0)=0 and D_f(0)=-2. Thus, f satisfies (2) and (3) at x=0. f satisfies (2) and (3) everywhere else because f'(X) exists when x¥0. Therefore, f is nonangular. Since each Dini derivative of f is finite everywhere, f is continuous. Using the facts that f is continuous and nonangular, theorem 4.2 is applied to show that £86661. Let A=[x: fs(x)>0]. Since fs(0)=%u we see that 06A. From the definition of rn(x), it is clear that rfi(x)$0 _._jL_.Ja - whenever x6Jn—[3n+l,3n]. Observe that J2nnA-¢ for all n62 and [Jn: n6Z+] converges to O. In addition, .. .(lJin', . :21: n ' 2n n 2-3 Therefore, ffm3. Notice that this example also invalidates the next natural assumption from (1), that a bounded symmetric derivative with the Darboux property is in MA. The following theorem somewhat clarifies the situation. Theorem 4.7. Let f60*s with F any primitive for f. Then f6Z(D(F)). 69 Proof. According to theorems 3.1(d) and 3.8, there is no generality lost in assuming that F is the nice copy of a primitive of f, because D(F) is, at worst, made larger. Let x6D(F). Since f(x)6nl, through the addition of a linear function, it may be assumed that f(x)=0=F(x). Because x6D(F)CC(F), there is a 6>0 such that [F(t)]<1 whenever lt-x]<6. Let [a,b]C(x,x+6) be such that f(t)20 on [a,b] and let c>0. Define A=[t6[a,b]: f(t)zc]. We claim that (12) e\A|slimt_.b_F(t)-1im F(t) . t-Oa-I- To see this, first note that the limits in (12) make sense because theorem 3.18 guarantees that F is nondecreasing on (a,b). Define _ .1 _ Fn(t)-n(F(t+n) F(t)) . It is clear that if y6D(F), then limndaFn(y)=f(y), so that by corollary 3.28, Fn converges to f a. e.. Choose any [c,d]C(a,b) such that [c,d]CC(F). Then, since f is non- negative and measurable on [c.d], d . . c][x6[c,d] . f(x)=c] [if c lim infndan(t)dts d+l c+l . . - d . n . n slim Infnwchn(t)dt$11mn_.°fd nF(t)dt-limnwa nF(t)dt= =F (d) -F (c) because F is continuous at c and d. Now, choose two seq- uences, (ch: n6Z+] and [dn: n6Z+], contained in C(F),such that cn decreases to a and dn increases to b. Then c]A]=limndac][x6[cn.dn]= f(X)2€}l‘ 7O slim supn“ (F (dn) -r (cn) ) slimtdbj (t) -lim p (t) trad- which is (12) . Since x6D(F) and f(x)=0, we see that given an n>0 there is a C6(0,6) such that Oo]6M4(D(F)) . J Proof. We may suppose that F is the nice copy of a primitive for f because this at worst makes D(F) larger. Let o=0 and define B=[x: f(x))o]. According to theorems 2.1 and 1.8(c), E is an F0 set. Let x°6EnD(F) and f(x°)=a. Using the fact that x°6D(F), we may write, for h sufficiently close to 0 F(x°+h)=F(x°)+ah+hn(h) where limhyon(h)=o. If hh1>0 and Ih+h11 is sufficiently small, then F(x°+h+hl)-F(x°+h) h(n(h+h1)-n(h)) (l4) bl = a hl + n (h+h1) . Choose c>O with O<fill be an upper bound for f. Then lemma 3.28 implies (16) F(x°+h+h1)-F(x°+h)$Ml{x: f(x°)>0]fl(x°+h,x°+h+hl)l. It then follows from (15) and (16) that [[x: f(x))0]n(x,+h,x°+h+h1)] a >c-:>O. hl 2M Theorem 2.1 implies that [x: f(x)>%] is an F0 set for (17) + + each n62 , so we may choose, for each n62 , a sequence of : m6Z+], such that closed sets, [En m [X: f(X) >-} = UélEn m It is clear that E= -nU___1[X= f(x)>-}= Un___1Um== 1En,m and since x063, there are integers, n and m, such that x0613n m' Then a)% and from (17) we see [[x: f(x)>0]fl(x°+h,x+h+h )l l :> 1 O hl 2nM> ° Therefore, if we choose nn =2nM6(0,2), the definition of M4(D(F)) is satisfied. Corollary 4.12. bACm4. Theorem 4.10 improves on a result of Kundu [16] which was that if f60 is continuous such that £86.80s is bounded, then £86m2(D(f)). The corollary is due to Zahorski [29]. From (1), we see that baCbA'Cm4. It is easy to see that bAflcbfios and the example in section 4.3 shows that b.80s is not contained in m3. .Therefore bA' is properly contained in bfios. The next two examples show that even 73 the containments bA'C2b037/74 and A'dc0sd are proper. Example. There is a bounded symmetric derivative, f6m . which is not a derivative. Let In=[2-n,2_n+1] for n62+. Partition each In into 2n equal subintervals, 1:, k=l, ...,2n.. If we write [o.B]=I:, for some k and n, then we may define 4 B-0 0_ _B(x-o) x6[o,o+—Zf0 —4-(x- BEE) x6 [c+E-;—c,o+-u—fl4iz-] 9:(X)= B.“ a—‘f-Bm-s) xe (d+-3—$’i—'°—l.el o xetdplc Using these functions, we define a n Zn. 423:. (g: (no—gk n(-x))+x(_,,' O) (x) xao f1(x)= x=0 NIH We must show that flem4. Let 06H! and define F=[X61R : fl(x)>a] . If 02%» then it is clear from the continuity of each 9: that F is open and consequently FEM4. So, we suppose that G<%. Then we may write F= u" m=:1[x6]R f (mam-137?“) Using the definition of f1' we see that for each n6Z+, the set rn={xem: f1(x)2d+1‘:———9-} consists of the set [0] and a sequence of disjoint closed 74 intervals converging bilaterally to zero. Thus, Fn is closed for each n, which implies that F is an E0 set. We choose the sequence,[Fn], to be the sequence of sets in the definition of M4. Now, let A=[x>0: f1(x)z%]. In the same manner as above, we see that A is closed and also that ACFn for each nEZ+. n-3}. Pick Choose a c>O and let k°=min[n6z+: c<2 e(0,c)6(0,24k°) and let h and hl be positive numbers such -n°+1) for some that hk°. We claim that there exist integers, k and n, such that I:c(h,h+hl). To see this, suppose not. Then (h,h+hl) can intersect at most two if the intervals (1:: n6Z+,lSk$2n], because otherwise it must contain one of them. Using the -n°+l fact that h<2 , we see 2no 1 _ , —2n -2n +3 hlslrno 1+[Ino_1]—s 2 °<2 o , This implies that h 2-n, k -3 - >—-——__ )2 ° h1 2 n°+3 >c by the choice of k0. This violates the assumption that h-c(0,c), we note that RO(Afl(O,°))CA to establish (Fn(h+hl,h) l [An(h+hl,h) ] ___.— 2 2 "hi ‘hl (30(An(o,~) )r)(h+h1.h) I [An(-h,-h-h1] 2 = > T] . ‘hl ‘hl Therefore, in the definition of the class M , if we let nn=n for each n6Z+, then the definition is satisfied at O with the set F. If x6F such that x¥0, than there is a neighborhood (x-p,x+p)CF. From this it is evident that the criteria of the definition of M are satisfied at x. 4 Therefore, F6M4. 76 If B=[x: f1(x)0 fl(x)+f2(x)= 1 x=O 2 x<0 which violates the Darboux condition at x=O. Thus, either fl or f2 cannot be a derivative. Let f be either f1 or 152 such that f is not a derivative. ExamEle. There exists an féms which is a symmetric deriv- ative, but not a derivative. 77 For n62+ define In=[2-n, 2-n+2—2n] and let gn be a nonnegative continuous function supported in In such that -n-l g -2 IIn Then, let 9(X)=Z%:19n(x) and 9(x) X>O f (x)= O " x=0 -g(-x) x0, k=max{n62+: Z'nzc} and N={x€fii: f(x)=0}. Then _ a -i -2i_ ‘NM-c, c)122c-21U:=k1n‘-2c-221=k2 +2 - =2c-2(2 k+1+§2‘2k+2)22c-2‘k. From this it is clear that 1Nfl(-c, c)l___l limc-OO 2c Therefore, 0 is a density point of N. Since f(O)=O, we see that f is approximately continuous at 0. Let F(x)=f :f(t)dt. Then F' (x)=f(x) whenever xalo because of the continuity of f. Also, fgfItIdt-j’zgh f(t)dt =1 ~o 2h I29 (t) dt-J' ‘3, g (t) dt 11“h-oo 2h =0. Therefore, F60 and FS=onSd. 78 To see that fiA', we first note that + . n 2-n _ n 9 -m-1 D F(O)Zlimn_‘°2 f0 f(t)dt—2 23ng >l>f(0) so F is not differentiable at x=0 and therefore FKA. Since F is absolutely continuous, it must be the nice copy of itself. Suppose GEA is an ordinary primitive for f. Then Fs(x)-Gs(x)=0 everywhere and by corollary 3.19 there must be a cent such that F(x)=G(x)+c. But, this implies that FGA, which is a contradiction. Therefore f has no ordinary primitive and is therefore not an ordinary derivative. l. 10. 11. 12. 13. 14. BIBLIOGRAPHY C. E. Aull, The first symmetric derivative, Amer. Math. M., 74(1967). 708-711. C. L. Belna, M. J. Evans and P. D. Humke, Symmetric and Ordinary Differentiation, Proc. Amer. Math. Soc., 72(1978), 261-267. A. Bruckner, Differentiation of Real Functions, Lecture Notes in Mathematics #659, Springer-Verlag, Berlin, 1978. z. Charzynski, Sur les fonctions dont la dériveé symmetrique est partout finie, Fund. Math., 21(1931), 214-225. M. J. Evans, A symmetric condition for monotonicity, Bull. Math. Inst. Acad. Sinica, 6(1978), 85-91. M. J. Evans and P. D. Humke, Parametric Differentiation, Coll. Math., to appear. F. M. Filipczak, Sur les deriveés symmetriques des fonctions approximativement continues, Coll. Math. 34(1976), 249-256. J. Foran, The Symmetric and Ordinary Derivative, Real Anal. Exch., 2(1977), 105-108. M. Foran, Symmetric Functions, Real Anal. Exch., 1(1976) 38-40. H. Fried, fiber die symmetrische Stetigkeit von Funk- tionen, Fund. Math., 29(1937), 134-137. K. Garg, On a new definition of derivative, Bull. Amer. C. Goffman. Real Functions, Rinehart, New York, 1960. A. Khintchine, Recherches sur la structure des fonctions mesurables, Fund. Math., 9(1927), 212-279. N. K. Kundu, On some conditions of symmetric derivatives implying monotonicity of functions, Rev. Roum. Math. Pures. Appl., 15(1970). 561-568. 79 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 80 N. K. Kundu, On some properties of symmetric derivat- ives, Ann. Pol. Math., 30(1974), 9-18. , 0n symmetric derivatives and on properties of Zahorski, Czech. Math. J., 26(1976), 154-160. S. Mazurkiewicz, Sur la dériveé premiere generalisée, Fund. Math., 11(1966), 129-136. S. N. Muckhopadhyay, 0n Schwarz differentiability IV, Acta. Math., 17(1966), 129-136. I. P. Natanson, Theory of functions of a real variable, Vol. 1, Ungar, New York, undated. , Theory of functions of a real variable, Vol.2, Ungar, New York, undated. D. Preiss, A note on symmetrically continuous functions, Casopis pest. mat., 96(1971), 262-264. I. Ruzsa, Locally Symmetric Functions, Real. Anal. Exch., 4(1978), 84-86. S. Saks, Sur les némbres deriveés des fonctions, Fund. Math., 5(1924), 98-104. W. Sierpinski, Sur une hypothése de M. Mazurkiewicz, E. Szpilrajn, Remarque sur la derivee symetrique, Fund. Math. 21(1931), 226-228. B. S. Thomson, 0n full covering properties, Real Anal. Exch., to appear. C. E. Weil, A Property for Certain Derivatives, Ind. Math. J., 23(1973), 527-536. , Monotonicity, convexity and symmetric deriv- ates, Trans. Amer. Math. Soc., 222(1976), 225-237. Z. Zahorski, Sur la premiere dériveé, Trans. Amer. Math. Soc., 69(1950), 1-54.