a: nut] ‘.. . hfiui- Lufiwuwnmmm..m - H I O .- 155. v o ‘Qflp t. ft 3.1.. .f ..A. 1|.vov1 J .1. 1.0.3 I’obfi. ul‘IIJ.‘ i'vf. FOIV’?‘ .I .c» 1.1.5.7.. Y; 7 . .5:- va'AaIIIco 1"}, ' 31.9..l. ex . t h L Brit's I: 9.. . . .. :c} .5. ‘V 17‘3.’: I... I V 0' '1’. Ir... agatzhi... ; u .r. .L. 7125:: 2’ .i . at y. .39.: 2,1311!!! . 1 XI: :13: I III“! .i. .51.... s If ‘11 .ll Ii..é;c.l .rla nun-“h..4t oiled-IO: I... I. v I! ’9!!!” i.‘1t.§ar‘l.£ .n.£....:.ftlra§.-.! LIL .v«..ll.... . I. x . . 1 . v-slvh‘ll", 71’ .J In!" 'HFSlfl lllllllll\llllllllllllHlllllllll llllllljlll 31293 0079 91 LIBRARY ‘ Mich'Qan State nlverslty This is to certify that the dissertation entitled THE t’EPxNO DEK‘WA'UVES presented by HMfwm N F E32; \ C has been accepted towards fulfillment of the requirements for Pill. D degreein ma‘f’tlepMCti‘l‘Cs owed a my ”1/ Major professor DateOdsywa THE PEANO DERIVATIVES by Hajrudin Fejzié A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 1992 Let } has a it! that fur The (on l‘th 0rd but the Lot by .X' a have sh ditteron deriva: sets .4" they p the firs that {( shown Peano A. ll\'eg_ Ofdln.‘ tires Wit ABSTRACT THE PEANO DERIVATIVES by Hajrudin Fejzié Let f be a function defined on an interval [a,b] and that k 6 N. We say that f has a Ic-th Peano derivative at :1: E [a, b] if there exist coefficents f1(x), . . . , fk(:c) such that f(a:r+ h) = f(x) + hf1(:r) + . . . + %fk(a:) + hkek($, h) where limh_.o ek(:c, h) = O. The coeflicent fk(:r) is called the k-th Peano derivative of f at x. The existence of a k-th ordinary derivative, f(")(:r), implies the existence of fk(:c) and fk(:c) = fl")(:z:), but the converse is not true for k 2 2. Let A’ be the class of all derivatives, and let [A’] be the vector space generated by A’ and O’Malley’s class Bf. S. Agronsky, R. Biskner, A. Bruckner and J. Marik have showed that every function [A’] has the form g’ + hk’, where g,h and k are differentiable. They also proved that f 6 [A’] if and only if there is a sequence of derivatives {22“} and closed sets {An} such that U§°=1An = R and f = vn on A... The sets An and corresponding functions 1),. are called a decompositon of f. The question they posed is whether every Peano derivative belongs to this class of functions. In the first part of this thesis a. positive answer to this question is given. Also it is shown that for Peano derivatives the sets An can be chosen to be perfect. Moreover it is shown that every k—th Peano derivative is the composite derivative of the (k — 1)-th Peano derivative relative to the sequence {An}. A. Bruckner, R. O’Malley and B. Thomson introduced the notation of path deriva- tives. They showed that path derivatives have many of the properties possessed by ordinary derivatives. In the second part of this thesis it is shown that Peano deriva- tives are also path derivatives and hence they have all the properties possessed by path derivatives. This gives another proof of the many properties possessed by Peano ' .n "‘ 'l .‘l’lfl’fi derive Tl derivz quest Fl Peanc and s derivatives and also answers the question posed by the above authors. The third part of this thesis shows that a k-th Peano derivative is a selective derivative of the (k - 1)-th Peano derivative, and hence gives a positive answer to the question posed by C. Wei] regarding Peano and selective derivatives. Finally the last part of this thesis shows that these results are still true if we replace Peano derivatives with generalized Peano derivatives, introduced by M. Laczkovich, and studied by C. Lee. To my brother-in-law MIDHAT DRINA, who was killed by Serbian irregulars on June 22, 1992. iv ACKNOWLEDGMENTS I would like to thank my advisor, Professor Clifford Weil, for all his help, encouragement and advice. His knowledge and enthusiasm were invaluable. TABLE OF CONTENTS INTRODUCTION .......................................................... 1 CHAPTER I .............................................................. 10 1.1 Decomposition of Peano derivatives ................................. 10 CHAPTER II ............................................................. 30 2.1 Peano and path derivatives ......................................... 30 CHAPTER III ............................................................ 41 3.1 Relationship between fk and f)._1 ................................... 41 3.2 Peano derivatives and Property Z ................................... 48 3.3 Peano and selective derivatives ...................................... 51 CHAPTER IV ............................................................ 57 4.1 Decomposition of Generalized Peano derivatives ..................... 57 4.2 Generalized Peano, path and selective derivatives .................... 68 BIBLIOGRAPHY ......................................................... 76 vi The in he defi\' to fir prevh One Inofix Pean< Defir crfit uhrrg f 011 17 introc Poirm Innst; fin k u ¢ It iSe defiVaI thf‘re i INTRODUCTION The definition of the k-th ordinary derivative of a real valued function is iterative in nature and thus easily comprehended if one initially understands what a first derivative is. This nice feature can present a problem, however, because in order to find the k-th derivative of a function f at a point 1:, one must know all the previous derivatives, not only at x, but at every point in some neighborhood of 3:. One type of generalized k-th order differentiation, having Taylor’s theorem as its motivation, attempts to skirt this drawback. This kind of differentiation is called Peano differentiation. Definition 0.0.1 A function f is said to have a lc-th Peano derivative at x if there exist numbers f1(a:), f2(a:), . . . ,fk(:c) such that k f(1 + h) = ftx) + W) + - - - + $4142) + ektx. h» (1) where 6;.(3, h) —+ 0 as h —) 0. The number fk(:c) is called the k-th Peano derivative of f at at. It will be convenient to denote f (2:) by fo(x). With this notation (1) becomes k hi h" f(1' + h) = Z fifflx) + Felix, h)- j=o ° ' This concept was presented in 1891 by the italian mathematician G. Peano. Peano introduced this type of derivative, obtained a product rule, a quotient rule, and pointed out that if a function f has an ordinary k-th derivative at :c, f(*)(z), then it must have a k-th Peano derivative at a: and fk(:r) = f‘k)(z). The converse is not true for k 2 2 as can be seen from the following example. Let f(x) = 16"“H sin-i: for a: 76 0 and f(0) = 0. It is easy to see that fk(0) = 0 but f(") at 0 doesn’t exist. Thus the k-th Peano derivative is a true generalization of the ordinary k-th derivative although obviously there is no difference for k = 1. ~.:'- In exhib derive tive. Barre funrti where the D. and f, and e; Cliftcr flinch} fl: p0,: dem'at Defini intf'f‘t‘q FE Wing In 1954 H. W. Oliver published the first extensive work devoted exclusively to exhibiting properties of k-th Peano derivatives. (See [10].) He showed that such a derivative has several of the pr0perties known to be possessed by an ordinary deriva- tive. Oliver established that if f], exists for all :1: in some interval I, then fk is of Baire class one; i.e., f). can be written as a pointwise limit of a sequence of continuous functions (A). (Denjoy had obtained this result earlier in the more general setting where fl. is defined relative to a perfect set H.) Oliver also showed that f]. must have the Darbouz property; i.e., that for any interval [a, b] C I if y is a point between fk(a) and fk(b), then there is c 6 (a, b) so that fk(c) = y (B), another property well known and easily verified for ordinary derivatives. Moreover, he showed that if f]. is bounded above or below on some [a, b] C I, then f]. is the ordinary lc-th derivative off on [a, b] (C). In particular, this yields the monotonicity theorem which states that if fk Z 0 on [a,b], then fk_1 is nondecreasing and continuous on [a, b] (D). Combining this with the fact that fk is of Baire class one, it follows that fk is an ordinary k-th derivative on an open, dense subset of I (E). R. J. O’Malley and C. E. Weil showed that if fk attains both values —M and M on some interval [a, b] C I, then there is an open interval J C [a,b], on which f1, = f“) and fl") attains both values —M and M on J (F). (See [12].) Ifg is an ordinary derivative on I, then for any open interval, (a,b), g‘1(a,b) either is empty or has positive Lebesgue measure, a result first proved by Denjoy. A function having this property is said to have the Denjoy property. Oliver showed that fk possesses the Denjoy property on I (G). (See [10].) Z. Zahorski proved that the following property is possessed by every ordinary derivative. Definition 0.0.2 A function g is said to have the Zahorski property if for each open interval (a,b), for each a: E g‘1(a,b), and for each sequence of intervals {In} con- verging to 2:, (The end points of the In converge to a: but 2: belongs to no In.) with Defin for er: that g' “I still is In Proacl genera merit i< look-in Th alapt for ap: anolho dm..., {Unclio fOraD m(g"1(a, b) n I") = 0 for every n, implies limn_.o‘> 3% = 0, where m(I,,) denotes Lebesgue measure of In and dist(:r, 1,.) denotes the distance between a: and In (H). C. E. Weil showed that a k-th Peano derivative also has the Zahorski property, and he introduced a property somewhat stronger that the Zahorski property, which he called property Z. (See [18].) Definition 0.0.3 A function g defined on an interval I is said to have property Z if for every 6 > 0, each :2: E I, and each sequence of intervals {In} converging to a: such that g(y) 2 g(m) on In or g(y) S g(x) on In for each n, we have 1. m({y e I. = lg(y) -9($)| 2 €}) = 0 n...» m(1n) + dist(a:,1,.) ' Weil showed that this property is strictly stronger than the Zahorski property, yet still is possessed by every k-th Peano derivative. In [3] the authors introduced the concept of a path derivative as a unifying ap- proach to the study of a number of generalized derivatives. Namely since many other generalized derivatives like approximate derivative, possess most of the properties mentioned above that are possessed by Peano derivatives, the authors in [3] where looking for a framework within which all of these derivatives could be presented. The perspective they chose was to consider just those derivatives of a function F at a point x which can be obtained as Hm F (v) - F (x) 146E... 14-”: y — a: for appropriate choices of sets Ex. One generalized derivative, then, differs from another only by the choice of the family of sets {Ex : a: E R} through which the difference quotient passes to its limit. For example, an approximately differentiable function F permits a choice of sets {E3 : a: E R} so that each E: has density 1 at 2:; for a Dini derivative each set may consist only of a sequence converging to m. This framework includes any generalized derivative for which the derivative at a point is a derived number of the function at that point. Since Weil has proved that fk(:v) is a derived number of fk_1 at a point x, we see that this concept of path derivatives also includes k-th Peano derivatives. But in order to get some properties for path derivatives, like those possessed by Peano or approximate derivatives, we require that the family of sets {E3 : :r E R} satisfy various “thickness” conditions. These conditions relate to the “thickness” of each of the sets E, and the way in which two of the sets intersect. The authors proved that path derivatives with certain type of conditions imposed on the family of sets {Ex : a: 6 R}, have many of the properties possessed by approximate and Peano derivatives. We will show that Peano derivatives are path derivatives with {E3 : a: E R} satis- fying some of the intersection conditions introduced by the authors mentioned above. This will give a positive answer to the question posed in [3]. In proving this assertion, we won’t use any known results for Peano derivatives. So this can be regarded as a new approach to studying Peano derivatives. Namely all of the properties (B), (C), (D), (E), (F), (G) and (H), that we mentioned before, we will get for Peano derivatives directly from the corresponding properties of path derivatives. The main tool will be a decomposition of Peano derivatives which we will discuss next. Let C be the family of all continuous functions on R, A the family of all differ- entiable functions on R and A’ the family of all derivatives on R. If I‘ is a family of functions defined on R, then by [I‘] we denote the family of all functions f on R with the following property: for each n E N there exist vn E I‘ and a closed set A, such that f = v" on An and ugg,A,. = R. In [1] (Theorem 2) it is shown that the following four conditions are equivalent: (i) There are g, h and Is: in A such that h’, k’ E [C] and f = g' + hk’. (ii) There is a (p E A’ and d) E [C] such that f = (p + 1!). (iii) The function f E [A’]. (iv) There is a dense open set T such that f is a derivative on T and f is a derivative on R \ T with respect to R \ T. The statement (ii) implies that [A’] is the vector space generated by A’ and [C]. In [1] (Theorem 3) it is shown that each approximate derivative, each approximately continuous function and each function in Bf = [C] belongs to the class [A’]. In [10] O’Malley showed that for approximate derivatives, the sets An from the definition of [A’] can be chosen to be perfect. The following question is raised in [1]. “Does every Peano derivative belong to [A’]?”. We will give a positive answer to this question, plus we will prove even more. Definition 0.0.4 Let f be a function defined on R. If there exist a function g, and closed sets A", n = 1,2,. .. such that Ufi°=1An = R and gl’ "(3) = f(a:) for :c E An, then we say that f is a composite derivative of g. We will prove that f;‘ is a composite derivative of fk_1 with respect to the sets P1,“, where for e > 0, 6 > 0 we define P5 = P(f,e,6) = {2: : Iek(:c,h)| < e for |h| < 6}. These sets were first introduced by A. Denjoy. He showed that with respect to these sets for 0 S l < i, i = 1,...,k — 1, f.- is an l-th Peano derivative of f,-_1, with ( f1|p6)(,-_,)(2:) = f.-(.'c) for at 6 P5, where the (l — i)-th Peano derivative is computed relative to P5. Using different techniques, we are able to improve his result. Namely, we show that the result also holds for the case i = 1:. Since Ufi°=1 P” = R, we have that fl, is a composite derivative of fk_1. This gives a positive answer to a question raised by C. Weil. (See [19].) From this result it is easy to conclude that f]. E [A’]. We just need to recall the fact that for any function g defined on a closed set A, such that at every point a: E A, a derivative g'(:r) computed relative to A exists, there exists a function G differentiable at every point a: E R so that 0],, = g and G’ I A = g’. We can enlarge the sets PM so that they are perfect and that f], is still a composite derivative of fk_1 with respect to these perfect sets. Therefore one more property possessed by approximate derivatives is also possessed by Peano derivatives. Because every composite derivative is a Baire 1 function, we see that f], is a Baire 1 function. Although this property is very easy to establish for Peano derivatives, for generalized Peano derivatives it is not so easy. We will discuss these derivatives in Chapter IV, but using some techniques similar to those that we use for Peano deriva- tives we will prove that generalized Peano derivatives are also composite derivatives and therefore, they are Baire 1 functions. Another immediate corollary is that f], is the approximate derivative of fk_1 almost everywhere. This result was first proved by Zygmund and Marcinkiewicz. (See [20] page 75.) We will generalize their result showing that fk is the l-th approximate Peano derivative of fk_1 with (fk-;)(a,,_;)(:r) = fk(:c) almost everywhere, where (ap—l) denotes l-th approximate Peano derivative. The sequence of sets {Pu} satisfies the condition Pn C PM]. This fact together with results already established will enable us to construct a system of paths {Ex : a: E R} satisfying the LC. property (as it was defined in [3]), so that f], is a path derivative of fic— forpa prope (E, : ta far and ( Fi condi‘ that s Defin called wists, sclcrti of f1,_1 with respect to this system. Using an induction argument and known results for path derivatives with such a system of paths, we get that Peano derivatives possess properties (B), (C), (D), (E), and (G). Since there is a nonporous system of paths {E3 : x E R}, such that f), is the path derivative of fk_1 with respect to that system, (a fact established in [3]), we get that Peano derivatives possess also properties (F) and (H). Finally we show that there is a system of paths {Ex : x E R} satisfying the I.I.C. condition (as it is defined in [3]), with fk the path derivative of fk-1 with respect to that system. This implies that f;‘ is the selective derivative of fk_1. Definition 0.0.5 If for a given function F there is a function p of two variables called a selection, satisfying p(x,y) = p(y,x) and p(x,y) E (x,y), so that . F(p(x,y))-F(Iv) Iii-'12 p(x.y)-$ (2) exists, we say that F is selectively differentiable at x, and the limit in (2) we call the selective derivative ofF' at the point x and denote it by F;(x). Selective differentiation was introduced by R. O’Malley. Motivation for introduc- ing selective differentiation was the fact that approximate derivatives are selective derivatives, which was proved by O’Malley. Showing that fk is a selective derivative of fk_1 we give a positive answer to a question raised by C. Weil. (See [19].) So Peano derivatives possess one more property possessed by approximate derivatives. Generalized Peano derivatives were introduced by C. Lee. (See [9].) He showed that every absolute Peano derivative on a compact interval is a generalized Peano derivative. Absolute Peano derivatives were introduced by M. Laczkovich. (See [7].) Definition 0.0.6 Let f be defined in a neighborhood of x. We say that the absolute Peano derivative off at x exists and is A (in symbols f‘(x) = A) if there is a function g, a nonnegative integer k, and a 6 > 0 such that (i) g, = f on (x — 6,2: + 6) and (ii) gk+1(x) = A. ,. bk.) ' .f' "will wher II indefi that 1 qtold any t' same. Note t the C0 d9fl\'at C. . Wells they art We . that (a; p(98110 r} al50 for PPanO (A Laczkovich showed that this concept is unambiguously defined, that if f “ exists on an interval, it is a function of Baire class one, it has the Darboux property, and if f‘“ is bounded above or below on an interval, then f“ = f’ on that interval. Definition 0.0.7 Let F be a continuous function defined on R, and let it E N. We say that F is n-th generalized Peano difierentiable at x E R, if there is a positive integer q, and coefficients F[,](x), i = 1, . . . ,n such that 0'1 (-a +1) F F(-q)(x + h) =2: h1____ F 11(3) +2 hq+j(q 814:5! + h0+n¢£fln(x, h) (3) 1:0 j=0 where limhno eqq+n(x, h): 0. Here F[o](x) = F(x) = F(°)(x) and F(‘j)(x) = fI F('j+1)(t) dt; i.e. F(‘jl is an indefinite Riemann integral of the continuous function F (‘j“) for j = 1,. . . ,q. Note that the definitions of F[,-](x), i = 0,1, . . . , n and of 65141:, h) don’t depend on which q-fold indefinite Riemann integral F ('9) of the continuous function F, is taken because any two differ by a polynomial of a degree less than q. The above definition is the same as the definition of (q+n)-th Peano derivative of a function F (‘9) at the point x. Note that every n-th Peano derivative is also a n-th generalized Peano derivative, but the converse is not true. Namely M. Laczkovich has constructed an absolute Peano derivative on an interval which is not an ordinary Peano derivative of any order. C. Lee showed that all properties (A), (B), (C), (D), (E), (F), (G), (H) and Weil’s Z property are possessed by generalized Peano derivatives and in particular they are possessed by absolute Peano derivatives. (See [8] and [9].) We will take a different approach to studying generalized Peano derivatives than that taken by C. Lee. Our approach will be similar to the one we used in studying Peano derivatives, so many results that we established for Peano derivatives will hold also for generalized Peano derivatives. In particular we will obtain that generalized Peano derivatives are composite derivatives and hence belong to [A’]. Also we will show that generalized Peano derivatives are path derivatives with respect to a bi- lateral, nonporous system of paths satisfying I.I.C. condition. Therefore generalized Peano derivatives are also selective derivatives. show that generalized Peano derivatives are path derivatives with respect to a bi- lateral, nonporous system of paths satisfying I.I.C. condition. Therefore generalized Peano derivatives are also selective derivatives. Thrc on R. B? 1.1 I Let C l) functior defined prOpertl f = I». four (01 Th. Corlilm prOVe , deriya, CHAPTER I Throughout this theses all the properties will be established for functions defined on R. But it can be easily seen that R can be replaced by any connected subset of R. 1.1 Decomposition of Peano derivatives Let C be the family of all continuous functions on R, A the family of all differentiable functions on R and A’ the family of all derivatives on R. If I‘ is a family of functions defined on R, then by [I‘] we denote the family of all functions f on R with the property that for every n E N there exist v7, 6 I‘ and a closed set A, such that f = vn on An and U;,'°=1An = R. In [1] (Theorem 2) it is shown that the following four conditions are equivalent: (i) There are g, h and k in A such that h’, k’ E [C] and f = g’ + hlc’. (ii) There is a cp E A’ and 2/) E [C] such that f = cp + 11). (iii) The function f E [A’]. (iv) There is a dense open set T such that f is a derivative on T and f is a derivative on R \ T with respect to R \ T. The statement (ii) implies that [A’] is the vector space generated by A’ and [C]. In [1] (Theorem 3) it is shown that each approximate derivative, each approximately continuous function and each function in Bf = [C] belongs to the class [A’]. The main goal of this chapter is to show that every Peano derivative is in [A’]. We will prove even more. Namely we will prove that every Peano derivative is a composite derivative. 10 11 Definition 1.1.1 Let f be a function defined on R. If there exist a function g, and for n 6 N there is a closed set An with gI’An(x) = f(x) V2: 6 An and Ug‘fflAn = R, then we say that f is the composite derivative of g. The following result due to Maiik says that every composite derivative is in [A’]. (See [14].) Theorem 1.1.2 Let a function g be defined on a closed set H. If 9’ exists on H, where g’ is computed relative to H, then there is a function G differentiable on R so that CI" = g and C’IH = g’. O’Malley proved that every approximate derivative is the composite derivative of its primitive. (See [11].) In this chapter we will prove that a k-th Peano derivative is the composite derivative of the (k — 1)-th Peano derivative. Thus we will get that every Peano derivative is in [A’], and hence possesses all the properties possessed by functions in [A’]. We will start with an elementary lemma. Lemma 1.1.3 For m E N we have 0 ifi=0,...,m—l m! ifi=m %(m+l)! ifi=m+1. f:(-1)"‘"'(';‘)j‘ = i=0 Proof: Let Bf" = m (—1)"“‘j(’;.‘)j’. Then B? = 0, B11 = 1, and B? = 1. i=0 Now we will proceed by induction on m. Suppose { 0 ift=0,...,m—2 Bid: m (m-l)! ift=m—l m2:_1mg ift=m. Note that B3, = (1 — 1)"‘ =0. Let 1 gig m+1. SinceiZ 1, 0’ =0. Thus Bf. = mi(—1)’”‘j(';‘:,‘)j"‘ .=1 12 111-1 = m §(-1)"‘“" (m;‘)(j + 1)"-1 m-l i—l = m 2(—1)m-1—1(mj—1)): (i:1)J-i-1-r j=0 r=0 i-l = m 2 (':1)B;,,'lf' , and by the induction hypothesis r=0 0 if i — 1 < m — 1 - mB,',",:] ifi—1=m-l m((T)B,',’,‘:11+('g)B,T,_1) ifi _ 1 = m 0 ifi < m = m(m — 1)! if i = m m(m(m — l)! + 951ml) ifi = m +1 0 ifi=0,...,m—1 ml ifi = m Lg-(m+1)! ifi=m+1. Cl Definition 1.1.4 The Riemann diflerence A?f(x) of order m ofa real valued func- tion f at a point x is defined by Arflz) = 2;":0(—1)’"‘j (T)f(x +jt). If f is continuous on R, then Atf(y) is continuous on R. This is the case if the k-th Peano derivative f1, exists on R. The relationship between A? and A?“ is given by the following simple lemma. Lemma 1.1.5 AI""‘f(fl=) = Al‘fla? + t) - ANGIE)- Proof: Arftz + t) - AI"f(-’v) = Z(_l,m-.(?),(, +(j+1)t)— Z(-—1)"‘"'(';-‘)f(x + it) = i=0 i=0 f(x + (m +1)t)+ i(—1)m+1-i( :31) + (3‘)) f(x + jt) + i=1 for ti 1. Ila in Leninia 1 Emma Then for Hmf A?“ TM TilEOr Then .. m. with n l3 (—1)"‘+1f(x) (since (if!) +(';.‘)= ("‘1“) it is equal to) m+l . . = §(-1)m+“’(mf‘)f(x +Jt) = Arenas)- :1 J: For the remainder of this chapter I: will be a fixed positive integer greater than 1. If a function f at some point x E R has a k-th Peano derivative fk(x), then using Lemma 1.1.3, we have the following formula for Riemman differences. Lemma 1.1.6 Let f be a function defined on R. Let x E R be such that fk(x) exists. ThenforOS m S k t"zjzot—1)m-j(';)j‘e.(z,jt) ifm>i- Proof: Arftx) = :(—1)m-J'(’;‘)f(x+jt) = fan) 1) MC; )(§;:O(j) t)"——’(;”+ I) +(jt)e.(x,jt)) = :t'flif) §(-1)m”(,)j’ +t‘Z(-1)"‘ 1(7')j 6,‘($, jt) and by Lemma 1.1.3 we have tmfmtx) + ,. 2,-.(-1)m-J( ),... emtz jt) if m— _.- t'E,"f_.0(—l)m’1(';.‘)je,-(x, jt) ifm>i. C1 The next theorem is an easy consequence of Lemma 1.1.6. Theorem 1.1.7 Let f be a function defined on R such that fk exists for each x E R. Then f), is a Baire 1 function. Proof: For each n 6 N let g..(x) = n"A" %.f(x) Then by Lemma 1.1.6 (applied with m = i = k) we have lim.._.oo g..(x) = fk(x). It remains only to notice that each 14 gn is a continuous function. Therefore fl; is a pointwise limit of continuous functions; i.e., ft is a Baire 1 function. E] In order to prove that every Peano derivative is a composite derivative, we need to construct a sequence of closed sets {An} whose union is R, and with respect to which f], is a composite derivative. The obvious candidate for a primitive is fk_1. We will investigate a relationship among the Peano derivatives f,- for i = 1, . . . , k on certain sets Fk( f, e, 6) which are defined next. Definition 1.1.8 Let c > 0 and 6 > 0 be given. Let P = Pk(f,e,6) = {x : [ek(x,h)| < 6 whenever |h| < 6}. These sets were first introduced by A. Denjoy. (See [5].) He proved that for i = 1,. . . , k—l and for 0 S I g i, f,- is the (i—l)-th Peano derivative of f; with respect to the closure of these sets with the expected values; i.e., with (fIIF)(,-_1)(x) = f,-(x) for every x E F. (He proved that the same conclusion holds if f is defined on some perfect set H having finite index provided fk, computed relative to this set, exists on H.) (See [5].) When f is defined on R so that fk exists on R (or f), exists on a perfect set H), we will show that the result also holds for i = Is. First, using different techniques than Denjoy used, we will prove his result. We begin with some elementary formulas. Formula 1.1.9 Let f be a function on R and let x E R. Forl E N, suppose that fz(x) exists. Then for each t E R £1-1(x, t) 2 Ag?)- + te;(x, t). 15 Proof: The assertion follows directly from the definition of Peano derivatives. II] Formula 1.1.10 Let f be a function on R and let x,t E R. Forl E N, suppose that fz(x) and f;(x +t) exist. Then f:_1(:c+ )— fz1(z)—tfz(x)=tZ(—1)’j(,’)j’q(:v,jt)+ i=1 l-l ;(_1)l—l-j (l;l)jl—l€l_l(x,jt)_ 1 (—1)"1—’(’j1)j"1q-1(x+t,jt). 1 ~ I J Proof: By Lemma 1.1.5 A[f(x) = Ai’1f(x + t) — A[‘1f(x). Applying Lemma 1.1.6 to both sides of above equality we get 1 t’ f;(x) + t’ 2(—1)'-J‘ (9.4.1., jt) = j=1 tl— lfl— 1($+t) )-—+tl 11-21(_ 1)l—1—j(l— 1)J’l-l Cl— 1(1: +t,jt)— :0 Wm. wa— 1)-"-J( )-'-I..-.. Dividing both sides of the above equality by t"1 gives the desired formula. [:1 Formula 1.1.11 Let f be a function on R and let x,t E R. Forl E N, suppose that fl($) and f1(:c + t) exist. Then 1+1 fz(:c+)- f:(=x) Z(— 1'+‘-J('+‘)J 61(rv 20+ j=0 1 1 Z(-1)l-j(,i)jléz($ajt) — EI—lll-’ (Djlfle‘ + t1jt)‘ i=0 i=0 16 Proof: By Lemma 1.1.5 A[+If(x) = A[f(x + t) — A[f(x). Applying Lemma 1.1.6 to both sides of the above equality we get 1+1 _ t' Zt—IWH(‘:‘)j'e.(x,jt) = l was +t) +t'zt—1)’-j(;)j'e.(z + wt) — I tlMil?) —t12(—1)"’C)1161($,fl)o j=0 Dividing both sides of the above equality by t’ gives the desired formula. C] Formula 1.1.12 Let f be a function on R and let x,t E R. For I E N, suppose that f1(x) and f1(x + t) exist. Then fz-1($ + t) - ftlxl t l 41x)=¥(—1)’*’G)i’a+ l-l ’—;l(f.(x) - fz(ar + t» + gnarl-1' (fireman) — .,(. + tut». Proof: From Formulas 1.1.10 and 1.1.9 we get fz-1(w+t)—fz(z) Hm )=tz(- 1)::C),z.,(.,,-.)+ l-l g<—1>'-“"(‘;‘)j'-‘(jt, tff" —)+ +1:th jt))- E(-1)"""(’§‘)i"1(’zf"‘(—."‘$ 1+ ) +Jt61(1‘ +t jt)) Dividing both sides of the above equality by t, and applying Lemma 1.1.3 we get the desired formula. [3 17 Theorem 1.1.13 Let f be a function defined on R such that f), exists for each x E R. There is a positive constant M such that V e > 0 and 6 > 0 if x and y are in P with Ix — y] < 1%, then lfk(y) - M3” 5 M‘ (1) fk-lIyz : ilk-1”) — fk($) 5 Me. (2) Moreover for [h] < 6 l: 1,. . . ,k f; and e;( ~ ,h) are bounded on P 0 [a,b] inde- pendent of h, for any interval [a, b]. Proof: Let x and y be in P such that |y — x] < ti], and let t = y — x. Set B = 2:3 ( k j 1 ) j". Then the left hand sides of the equalities in Formulas 1.1.11 and 1.1.12 are bounded by 3B6 and 3B6 + 3Beig—1 respectively. Hence (1) and (2) follows for M = 3B 5-125- Let [a, b] be any interval. From (1) we see that ft is bounded on P n [a, b]. From Formula 1.1.9 (applied with l = 11:) it follows that for |h| < 6 Iek_1( - , h)] is bounded on P n [a, b] independent of h. Now from Formula 1.1.10 (applied with l = k) we see that fk_1 is bounded on P (1 [a,b], and again going back to Formula 1.1.9 (applied with l = k— 1) we see that for Ihl < 6 ]6k_2( - , h)| is bounded on Pfl[a, b] independent of h. Continuing we can deduce that there is a constant C so that I f;(y)] S C and |ez(y,h)| S C whenever y E P 0 [a,b], |h| < 6 and 1 _<_ I S k. C] The next theorem says that if we replace P by P, then the conclusion of Theo- rem 1.1.13 still holds. Theorem 1.1.14 Let f be a function defined on R such that f]. exists for each x E R. Then P Q P(f,3e,6). 18 Proof: Let x E P, and let {xn} E P be a sequence such that limn...» xn = x. Let [a, b] be such that {xn} C Pfl[a, b]. From Theorem 1.1.2 we see that f; for 1 S l S k is bounded on P n [a, b]. Therefore we can choose a subsequence {xnj} converging to x such that for i = 1- - - k the sequence {f,-(x,,j)} is convergent. Let these sequences converge to F,-(x), i = 1,... , 1: respectively. Let h with [h] < 6 be given. Suppose that [h + x — 3.1,] < 6 for everyj E N. Therefore Iek(x,,j , h+x-xn,)| < 6 so we can also suppose that this sequence converges. (If not, then extract a convergent subsequence.) Denote its limit by E(h). Since fk(x,,j) exists, (h + x — x,,1.)"‘1 f(x + h) : f(xn,) + (h +37 "' $n,)fl(xn,) + + fit—10311,) (k — l)! + (h + x — xnj)"_l(h + x — xn1)(fk(:!n’) + ek(xn’., h + x — 33%)) . (3) Lettingj —) oo in (3) we get k—l x f(x + h) = f(x) + hF1(x) + - -- + MIT’s-1(3) + h""1 h (5% + E(h)). Since limh_.o Milk?) + E (h)) = 0, by the uniqueness of Peano derivatives we have E-(x) = f,-(x) for 1 S i S 19—1 and fig) + an, h) = Eff) + E(h). (4) Since |E(h)| S e, from (4) we have that fk(z) - F.(::) k! = |E(h) - 61(x,h)| S 6 + l€k(-’v,h)|- (5) The left hand side of (5) doesn’t depend on h so letting h —+ 0 in the right hand side of (5) we get MW Finally this estimate and the formula ek(x, h) = E(h)+m9fi’—*Jfl gives lek(x, h)| _<_ _<_e. 26 < 36 for [h] < 6. Hence x E P(f, 3e, 6) and the theorem is proved. Cl 19 In order to prove that Theorem 1.1.7 holds for i = k, we need a formula that involves more than two variables. We will derive a formula, (Theorem 1.1.17 below) involving three independent variables. The proof of the formula is elementary, but the formula itself is the crux in what follows. Lemma 1.1.15 Let 0 S s S k — 1, x1 E R and let a function f be defined on R, having a k-th Peano derivative at x1. Then I: s Aif($l) = t'fa($1) + 2 Ex- 1)‘-J(;j) lt___ flu-"(31) l=s+1j=0 tk Z(-1)"’(,-)J €k($1,Jt) Proof: .1 I: Am.) = gem-1L) (it ‘(It)‘f—'——‘)+(Jt) 6.6. m) a k = gt—w- (,);:‘t"’(x‘)+t"2( wet-k 6:.(x1 m k 3 s- a .llltf—J—(fl) I: .— = ;;(-1) 101 +t ;(— 1) J(;)j 6.61.11) by Lemma 1.1. 3 = zit- w-Jc)'1‘—,——f' 0. Then f, is (k — 3) times Peano diflerentiable with respect to PU, 1,6) with the expected value; i.e., (f,|-,-,-)k_,(:c) = fk(:c). Proof: Let a: 6 F(f, 1,6) and let 1 > e > 0 be given. Then there is a0 < 17 < 6such that |ek(a:, h)| < 6 whenever |h| < 17. Let M be the constant from Theorem 1.1.19. Let 2:1 6 P(f,1,6) so that Ian —:r| < 7:313. Let t- — (ml—flei. Then forj = 0,1,...,k—1 we have 25 1:1 — a: +jt|= (1+je%)|:r1 — :cl < k|xl — xl < 17. Hence (7) lek(:r,a:1 — a: +jt)| _<_ c and (8) |€k(rt1,jt)| S 3- (Since-17f. 1.5) C P(f.3.5) and kltl _<_ klxl — 36| < n < 61) (9) By Theorem 1.1.17 I: ($1 _ 3))l!-s f,($1)— ”—2 (l _ 3)f1($) = l: s a k '3' $1 _ it l-i E(—1)“"(§) Z 27‘” (2 fifth) —fi($1)) + j=0 i=s+1 2! l=i ' . , — + 't " . §(_1)8—J (j) ((131 :9 J ) €k($a$1 _ $+]t) _ tk” :(— 1)‘ ‘1 (Djk £k(x1,jt) and by Theorem 1.1.19 j=0 3 (ah-.11-.) 1? f.(1)2—1———(1_f:()) s k 2 C): —|:rl — :1:|' ’eTMle — xlk‘ "‘ j=0 i=s+lz s a -8 1+ ‘él/lc k . 2(j)|x1_$lk ( :13; )|€k($,$1—$:Jt)|+ j=0 Ix. — sunk-2‘? 2 (;)j"lc.(z1,jt)| and by (8) and (9) j=0 8 k 'i . a - k S E (3) 2 lel -— :rlk”c"E£M + 2 (film — :cl:”-(ii.ll—e + j=0 i=s-lv-1z j=0 (I |x1- w- 2‘14 2 (;) *3- — le new): (9f; 2%? + 1:0 j=0 i=s+1z lml — :rIk-’ 1‘: ()(1 +j)ch + |a:1 — xl" ’3: C)j" 5T 1:0 j=0 Since 6 was arbitrary, we have that hm f.$121_.L17—)yI-n:mfz($)_o. 1'1 €P.$1-'Z ($1 " 37),“. 26 Therefore the assertion of the theorem is proved. C] Theorem 1.1.20, together with the simple observation that U°° Pk(f,1,1/n) = R n=1 has many applications. Corollary 1.1.21 Let f be a function defined on R such that fk exists for each x E R. Then f), is the composite derivative of fk_1. Definition 1.1.22 A function f is said to have a k-th approximate Peano derivative at x if there exist numbers fap(1)(x), fap(2)(x), . . . ,fap(k)(x) and a set V with density 1 at 0, such that hi: f(x + h) = f(x) + hfauo(1)(17) + ' ° ° + k—!fap(k)($) + “(13: h) (10) where ek(x, h) —+ 0 as h E V, h —» 0. For k = 1 we have the definition of the approximate derivative. Corollary 1.1.23 Let f be a function defined on R such that fk exists for each x E R. Then f, is almost everywhere (I: — 3) times approximately Peano differentiable with the expected values; i.e., (f,)a,,(k_,)(x) = fk(x) for s = 1, . . . , k - 1. Proof: Let x be a point of density of P = Pk(f,l,% . By Theorem 1.1.20, (f,|p)k_.($) = fk(x). Since x is a point of density of P, we see that (f,)a,,(k_,) exists at x and equals fk(x). Finally the Lebesgue Density Theorem and the fact that U$,°=1Pk( f, 1, i) = R proves the corollary. D The case .9 = k — l was proved by Zygmund and Marcinkiewicz. (See [20], page 77.) Corollary 1.1.23 can be regarded as a generalization of their result. 27 Corollary 1.1.24 Let f be a function defined on R such that fk exists for each x E R. Then f1, 6 [A’], and hence (i) there are g, h and q in A such that h’,q’ E [C] and fk = g' + hq’, (ii) there is a go 6 A’ and 11) E [C] such that fk = so + w, (iii) there is a dense open set T such that fk is a derivative on T and fl. is a derivative on R \ T with respect to R \ T. We will end this chapter with a different decomposition of R into closed sets so that fl. is the composite derivative of fk_1. Definition 1.1.25 Let f be a function defined on R such that f], exists for each x E R and let H(f.M.6)={x= “1 k-l no) + 2(-1)"“"’(*E‘)j*e.(a=.jt) s M for M < 6} where M and 6 are positive constants. Lemma 1.1.26 Let f be a function defined on R such that fk exists for each x E R. Then for any 6 > 0 we have U°fi=,H(f, M, 6) = R. Proof: The assertion follows from the fact that 6;.(x, jt) is a continuous function oftforj=0,1,...,lc-1.D Theorem 1.1.27 Let f be a function defined on R such that f}, exists for each x E R. Then H = H ( f, M, 6) is closed and fk-1 is differentiable on H relative to H with fk_1|’,,(x) = fk(x), also Ifk(x)| 5 2M for every x E H. 28 Proof: Let x E H. Let 1 > e > 0 be given. There is 0 < 17 < 6 such that l€k($,h)l < 6 whenever |h| < 17. Let xn E H so that Ixfl — xI < 53-. Then for t=(xn—x)e% wehaveltl <6and |x,,—x+jt| x and let 0 # |t| < 6. Then (11) yields _ k—l _ x _ z ' k t (k 2 1f.(.)+ 2(—1)"-“’(";‘)( " .5") mm — 2: +jt)) s t (k _ 1fk(xn) + E(“Uk-l-j (k71)jk€k(xn .7”) + 2 i=0 1 ’ |fk-1($n) - fk_1($) — (17a - a¢)fk(-”B)| S lth + |fk-1($n) - fk_1(:v) - (3n - 1‘)fk($)|- (13) Letting n -—+ co the left hand side of (13) becomes t k‘lf k-l lk-l—j k—l -k 't 2 .(x)+2(— ) (, )2 “(an while the right hand side of (13) is Ith. Hence a: E H. That I fk(x)| 3 2M on H follows from Definition 1.1.25 taking t = 0. E] Theorem 1.1.27 and Lemma 1.1.26 combine to say that fk is the composite deriva- tive of fk..1, a result that we already established. But this can be regarded as a simpler proof of that result, because the only tool we used was a special case of The- orem 1.1.17, whose proof is even more elementary than the proof of Theorem 1.1.17. On the other hand the sets H ( f, M, 6) are already closed. CHAPTER II 2.1 Peano and path derivatives We will start this chapter with the notion of a path derivative that was introduced in [3]. Definition 2.1.1 Let x 6 R. A path leading to x is a set E, C R such that x 6 E3 and x is a point of accumulation of Ex. A system of paths is a collection E = {Ex : x E R} such that each Ex is a path leading to x. Definition 2.1.2 Let F : R —+ R and let E = {Ex : x E R} be a system ofpaths. If Hm F(y) - F(x) yeExw-w y — x = f(x) is finite, then we say that F is E-difierentiable at x and write Fg(x) = f(x). IfF is E differentiable at every point x, then we say simply that F is E differentiable; we call F an E-primitive and f an E-derivative. Definition 2.1.3 Let E = {E3 : x E R} be a system of paths.(IfE has any of these properties at each point, then we say that E has that property.) E is said to be bilateral at x if x is a bilateral point of accumulation of Ex. E is said to be nonporous at x if E: has left and right porosity 0 at x. The basic definition of porosity of a set E at x from the right (left) is the value lim sup,_,0+ l(x,r,E)/r , where l(x,r, E) denotes the length of the largest interval contained in the set (x, x + r) H (R \ E) ((x -— r, x) n (R \ E)). Porosity 0 at x means both right and left porosity 0. Note that a nonporous system is necessarily bilateral. 30 31 Definition 2.1.4 Let E = {Ex : x E R} be a system of paths. E will be said to satisfy the condition listed below if there is associated with E a positive function 6 on R so that whenever O < y — x < min{6(x),6(y)}, the sets E, and Ey intersect in the stated fashion: i) intersection condition 1.0.: Ex 0 E3, 0 [x, y] 9i 0; ii) internal intersection condition I.I.C.: E, D E” 0 (x,y) 95 0; iii) external intersection condition E.I.C'.: Eanyfl(y,2y-x)5£0 andEanyfl(2x—y,x)7é0 We will prove that for every k-th Peano derivative fk there is a nonporous bilateral system of paths E satisfying LC. and I.I.C. conditions, for which f), is the E-derivative of fk_1. In this chapter we will prove that E satisfies only the LC. condition. To show this first we will prove the following theorem due to Mafik. (See [13].) Theorem 2.1.5 Let k E N, x E R. Suppose that a function f has a k-th Peano derivative at x. Define P(y) = 215:0 y — x)‘—‘},9 (y E R). Let c > 0, n > 0. Then there is a 6 > 0 such that if I is a subinterval of (x — 6,x + 6), j an integer with 0 < j S k and if either f(j) S P”) on I or f(j) Z P”) on I, then m({y E I : |f(5)(y) — P(j)(y)| Z er — xlk’j}) _<_ 1] ~ (m(I) + d(x, 1)). (Here m denotes Lebesgue measure and d(x, I) denotes the distance from x to I.) In order to prove Theorem 2.1.5 we need two lemmas. Lemma 2.1.6 Let f be a monotone differentiable function on a bounded interval I . Let c > 0, [3 > 0 and let m{x 6 I: |f’(x)| Z e} 2 fl. Then there is an interval J C I such that m(J) = ,6/4 and that |f| 2 63/4 on J. 32 Proof: We may suppose that f’ 2 0 on I. Let (a, b) be the interior of I. There is 3. c6 [a,b] such that f S 0 on (a,c) and f 2 0 on (c,b). Set B = {x E I: f’(x) 2 e}. If m(Bn (c,b)) 2 fl/2 and if x E (b— fl/4,b), then f(x) 2 f: f’ 2 e-m(Bfl (c,x)) 2 6 ° (m(3 n (c,b)) - (b - 93)) Z 6 - (fl/2 - W4) = 6fl/4- If m(3 n (0,6)) 2 3/2, then, analogously, f S —cfl/4 on (a,a + 3/4). C] Lemma 2.1.7 Let I be a bounded interval and let j be a natural number. Let g be a function such that either gm 2 0 on I or gm 5 O on I. Let c > 0, 5 > 0 and let m{x E I: IgU)(:c)l Z 6} _>_ 6. Then there is an interval J C I such that m(J) = ,6/4j and that Igl Z cflj/4‘+"'+j on J. Proof: The assertion follows by induction from Lemma 2.1.6. C] Proof of Theorem 2.1.5: Let g = f — P, and let a = 41+"'+". There is a 6 > 0 such that for each y E (x—6,x+6) we have 3"alg(y)| S f’lkly- xlk- (1) Now let I be a subinterval of (x — 6,x + 6) and let j be an integer, 0 < j S k. Let B = {y E I : |g(j)(y)| Z ely — xlk‘j}, B = %m(B). Suppose that 3 > 0. Let C = B\ (x — ,B,x + fl). Now Igml Z 6,3(k-j) on C and m(C) 2 B. If either g(j) _>_ 0 on I or g”) S 0 on I, then by Lemma 2.1.7, there is an interval J C I such that m(J) = ,8/4j and that '9' Z léflk_j ' flj = 16.3,‘ on J. (2) a a Together (1) and (2) yield (3fl)" _<_ nkIy-xl" for every y 6 J. Hence m(B) _<_ 17d(x, I). C] 33 Definition 2.1.8 A real valued function f defined on an interval I is said to have the intermediate value property if whenever x1 and x2 are in I, and y is any number between f(xl) and f((Eg) , there is a number x3 between x1 and x; such that f(x3) = y. A function having the intermediate value property is called a Darboux function. It is known that a k-th Peano derivative, fk, is a Darboux function. Also it is known that if f], is bounded either from above or below, then the k-th ordinary derivative, f(k), exists with the obvious equality, f“) = fk. In the next theorem we will only assume that these two properties hold for any l-th Peano derivative where O S l S k — 1. We know that any continuous function is Darboux, so for k = 1 the above assumptions trivially hold. Theorem 2.1.9 Let lc, l E N, with l S k — 1. Assume for each function g defined on an interval I having an l-th Peano derivative, g,, on I, g; is Darboux and if g; 2 0 on I then 9; = g“) on I . Suppose f is a function defined on R so that fk exists for each x E R. Then there is a bilateral nonporous system of paths E = {E, : x 6 R} satisfying the I. C. condition such that f], is the E-derivative of fk-1. We will need some lemmas before we prove this theorem. Lemma 2.1.10 Suppose that the assumptions of Theorem 2.1.9 hold. Then for every 6 > 0 and r] > 0 there is a 6 > 0 such that if] is a closed subinterval of (x — 6,x + 6) with x not in I such that Ifk-1(y) — fie-105) y — x for all y E I, then m(I) S 17d(x, I). - fk($)| Z 6 (3) Proof: Let 6 be chosen according to Theorem 2.1.5 applied with :7 replaced by m = 17/(1 + 7]) and with j = k — 1. Let I be as above, and let g(y) = f(y) — 34 yk‘lé'fff—ffiz — (y - x)"£‘,§f-l. Then g has a (k — 1)—th Peano derivative and gk_1 (y) = fk_1 (y) — fk-1(x) — (y — x) fk(x). So by assumptions gk_1 is a Darboux function. By (3) ng-1(y)| Z cly — x| on I. Since x is not in I, ng_1(y)| > 0 on I and since gk_1 is a Darboux function, we have either gk—l > 0 on I or -gk-1 > 0 on I. Hence by the assumptions, g]._1 is the (k — 1)-th ordinary derivative of g on I. Therefore f is (k — 1) times ordinarily differentiable on I and by the uniqueness of Peano derivatives, f(k'll = fk-1 on I. Now we can apply Theorem 2.1.5 with j = k — 1, which gives that m(I) _<_ 171 - (m(I) + d(x, 1)). Hence m(I) S nd(x, I). C] Next we will prove a lemma using ideas from the proof of 3.6.1 in [3]. Lemma 2.1.11 Under the assumptions of Theorem 2.1.9, for each point x E I there is a path E3 leading to x and nonporous at x so that lim fir-1U)" fk-IW) = 116131.31“: y — 13 fk($)- Proof: For each 6 > 0 let 6(6) be as in Lemma 2.1.10 applied with n = 6/2 and let {6;} be a sequence so that 0 < 6, _<_ 6(1/l) and 61+; < 61/2. Set E:={x}u0{y>6z.2= f""(”)‘f""(””) -fk(x) <1/I}. l=1 y_$ It is certainly true that lim fk—1(3/)- fin-10") véng-w y -— x = fk($)- (Although this assertion is true if x is not a point of accumulation of EL, in fact we will prove that E; is nonporous from the right at x.) Suppose E; is porous from the right at x. Then there must exist a number 1/2 < 0 < 1 and a sequence of numbers h; l 0 with [x + 0h;,x + h] n E; = 0 for every index I. Choose an integer lo larger that (1 — 0)"1 (i.e. so that if I _>_ lo, 35 then 1 — 1/1 > 0) and let jo be the first index for which hjo < 610. Fix I so that 61+, S hjo < 61, and note that I 2 I0. Since hjo < 61 _<_ 6(1/1), by Lemma 2.1.10, there must be a point z with x + hj0(1— ’*-'<’2::*-1‘:1— fk(x)l < 1/1. 1/1) S 2 S x + hjo such that We then have the inequalities, x+61+2_ 6(y), then P(f, l,6(x)) C P(f,1,6(y)) and hence x E Ey- Therefore Ex 0 E" f] [x,y] yé 0. Hence E satisfies the LC. condition. This completes the proof of the theorem . E] 36 Now we will prove that we can drop the assumptions concerning the arbitrary function g from Theorem 2.1.9. In order to do that we list some theorems from [3] about path derivatives. Theorem 2.1.12 Let E = (E, : x E R} be a system of paths that is bilateral and satisfies the LC. condition. If f is an exact E-derivative and is Baire 1, then f has the Darboux property. Proof: This is Theorem 6.4 in [3]. D Theorem 2.1.13 Let E = {E,, : x 6 R} be a system of paths that is bilateral and satisfies the LC. condition. If _F_’E Z 0 on [a,b], then F is nondecreasing on the interval [a, b]. Proof: See 4.7.1 in [3]. El Theorem 2.1.14 Let E 2 {Ex : x 6 R} be a system of paths and suppose F is monotonic. If E is nonporous at a point x, then Ego) = E(x) and T’Ee) = 7’s). Proof: See Theorem 4.4.3 in [3]. D Theorem 2.1.15 Let f be a function defined on R and let lc E N. Suppose fk(x) exists for each x E R. Then there is a bilateral nonporous system of paths E = {Ex : x E R} satisfying the I. C. condition such that f], is the E-derivative of fk_1. 37 Proof: The proof is by induction on k. For k = 1 there is nothing to prove. Let 1 S l S lc — 1, and let a function g defined on some closed interval I have a l-th Peano derivative on I. Suppose the assertion of the theorem is true for every 1 S j S lc — 1, and every function h defined on some closed interval J, so that h,- exists on J. (Note that we can restrict ourselves only to closed subintervals because we can always extend h to R so that h j exists on R. For example if J = [a, b], then, we can define h(y) = {=o(y — x)‘1‘§,fl for y E (—oo,a) and h(y) = {=o(y — x)‘£§,fl for y E (b, 00).) By Theorem 1.1.7 g; is a Baire 1 function. By the induction hypothesis and Theorem 2.1.12, g; is a Darboux function. Suppose that g; 2 O on I. Again by the induction hypothesis but now using Theorem 2.1.13, g1_1 is nondecreasing on I. By Theorem 2.1.14 g[_1 = g, on I. Also there is an a such that g1_1 — a Z 0 on I. Let h(x) = g(x) — 0%. Then hz_1 = gz_1 — a and hence h1_1 Z 0 on I. Proceeding as before h[_2 = h1-1 on I. This implies g{_2 = g;_1 on I. Continuing in this fashion one can deduce that g“) exists on I. Now we can apply Theorem 2.1.9. Cl Corollary 2.1.16 Let f be afunction defined on R such that fk exists for each x E R. Then f]. is a Darboux function. Proof: The assertion follows directly from Theorems 2.1.15, 2.1.12 and Theo- rem 1.1.7. Cl Definition 2.1.17 A perfect road of a function f at a point x is a perfect set P such that (I) x is a bilateral point of accumulation ofP (2) f|P is continuous at x. The assertion of the next corollary follows directly from the properties of Baire 1, Darboux functions. (See [2]) 38 Corollary 2.1.18 Let f be afunction defined on R such that fk exists for each x E R. Then (1) For each x, there exist sequences xn /' x and yn \, x such that fk(x) = limit—too fl:($n) = limit-"00 fk(yn)- (2) For each x h(x) 6 [lizrgggf fk(z), limsup fk(2)] n [lizrggyf fk(2),1im 83p fk(2)]- 2-02- 2—03 (3) For each real number a, the sets {fk _<_ a} and {f1c Z a} have compact compo- nents. (4) The graph off). is connected. (5) The function f], has a perfect road at each point. (6) Each of sets {f1c < a} and {f}, > a} is bilaterally c-dense in itself. ( See [2].) (7) Each of sets (f;‘ < a} and (f;‘ > a} is bilaterally dense in itself. Definition 2.1.19 Let E = (E; : x E R} be a system of paths and F a function on R. We say that F has the monotonicity property relative to E if for any interval [a,b] the conditions FI’;(x) exists a.e. in [a,b] and Fl’;(x) Z a a.e. in [a,b] (resp. Fax) 5 0) imply that the function F(x) — ax (resp. ax — F(x) ) is nondecreasing on [a, b]. Theorem 2.1.20 Let E = {Ex : x E R} be a system ofpaths and let F be a function. If E is bilateral and satisfies the intersection condition, and F is E -difl'erentiable, then F has the monotonicity property relative to E. 39 Proof: See Theorem 6.6.1 in [3]. [3 Corollary 2.1.21 Let f be a function defined on R such that fk exists for each x E R. Let [a,b] be an interval, and a be any constant. If f1, 2 a (or fk S a) on [a,b], then a) fk_1(x) — ax (ax — fk_1(x) ) is nondecreasing and continuous on [a, b] b) f“) exists and fl” = f,c on [a, b]. Proof: The assertion follows directly from Theorems 2.1.15, 2.1.20 and 2.1.14. El Corollary 2.1.21 was first proved by Oliver in [10] and Corominas in [4]. See also Verblunsky [15]. Definition 2.1.22 Let f be a function defined on R. If for any interval (a,b), f‘1(a,b) # 0 implies m({x : f(x) 6 (a,b)}) > 0, then we say that f has the Denjoy property. Theorem 2.1.23 Let E 2 {Ex : x E R} be a system of paths and let F be an E-difierentiable function that has the monotonicity property relative to E. If F g; is Darboux Baire 1, then F}; has the Denjoy property. Proof: This is Theorem 6.7 in [31°C] Corollary 2.1.24 Let f be a function defined on R such that fk exists for each x E R. Then f), has the Denjoy property. Proof: The assertion follows directly from Theorems 2.1.15, 2.1.20, 2.1.23 and Corollary 2.1.16. CI Corollary 2.1.21 first was proved by Weil in [17]. 40 Theorem 2.1.25 Let E = {E_., : x 6 R} be a nonporous system of paths satisfying the intersection condition. Suppose that F is an E-difierentiable function with F 1’; Baire I. If F}; attains the values M and -M on an interval Io, then there is a subinterval I of Io on which F is differentiable and F’ attains both values M and —M. Proof: This is Theorem 8.1 in [3]. An immediate consequence of Theorems 2.1.25 and 2.1.15 is the following corollary first proved by O’Malley and Weil in [12]. Corollary 2.1.26 Suppose fk(x) exists for all x in Io and let M 2 0. If fk attains both M and —M on Io, then there is a subinterval I of Io on which fl. = f(k) and f(k) attains both M and —M on I. This corollary has some nice and immediate applications. The reader is referred to [12] for the details and proofs. CHAPTER III In Chapter I we have shown that for any k-th Peano derivative, fk, defined on R, there is a countable decomposition {Ha} of R into closed sets and a sequence of differentiable functions {vn} so that for each n 6 N, v;ly,, = fk. In [10] O’Malley showed that the same holds for approximate derivatives. Moreover he proved that for any approximate derivative there is a decomposition of R into perfect sets with the above property. In this chapter we will show that the same holds for Peano derivatives. Also we will show that any k-th Peano derivative, fk, is a path derivative of fk-1 with respect to a system of paths that is bilateral, nonporous and that satisfies the internal intersection condition I.I.C.. This will enable us to give a positive answer to the question posed by C. Weil, regarding the relationship between Peano and selective derivatives, (See [19].) namely, the last result of this chapter is that f], is a selective derivative of fk_1. 3.1 Relationship between f1. and fk_1 We will begin this section with a very well known lemma. Lemma 3.1.1 Let n E N and let f and g be functions on R having n-th Peano derivatives fn(x) and gn(x) at some point x. Then the function fg has n-th Peano derivative at x and fl (fg)n(x) = 2 (:)f.(x)g.-.-(a=). i=0 Proof: Let x 6 [a,b] be such that f,,(x) and gn(x) exist. Thus the following formulas hold 41 42 f(x + h) = f(x) + hf1(-’€)+° +-,,-, n($) + h"€n(5r h) 9(32 + h) = 9(3) + h91($)+ + h—ignm + h" €n(x h) where en(x, h) and En(x, h) tend toward 0 as h —9 0. Then f(x+h)g(z+h)= 2 3,2; 'f‘ff)f}‘i(f))g+ h"e..(:c, h)g(x + h) + h"En(x, h)(f(x + h) — 5,,(x, h)) = 50% g ( )L-(z) )9..le +h"e:.(z,h) where en(x, h) = 6,,(x, h)g(x + h) + €n(x, h)(f(x + h) — 6,,(x, h)) Since obviously linu._.o 6;,(x, h) = 0 we have (mm) exists and (19m) = 2"; (;)f.(x)g.-.(x).g i=0 Lemma 3.1.2 Let f and g be functions on R such that the n-th Peano derivative, fn( x), and the n-th ordinary derivative, g(")(x), exist at some point x. Then :(-1)’(2)(fg“’)n—j(x) = mum. Proof: By Lemma 3.1.1 i<—1)j(:)(fgm).-.(x) = :(-1)j(?) E: ("f-j)fi($)(9(j))(n_j-e)($) = genie) Z (":’)f.-(x)g<"~"(x) = n n—i Z Z<—1)J‘ (3) ("?’)fa(w)g‘”“’(x) ——- i=0j=0 43 n n—i gig—1)"??? (i)fl(x)g‘""’(rv) = n n-i E (2‘) EH)" ("rhesus-”(s = 5:0 j=0 mean) + 33 (2)0 - 1)"“f.-($)g‘"“’(z) = semen i=0 Lemma 3.1.3 Let H be a continuous function in an interval [a,b] containing 3;. Suppose that H is n times Peano diflerentiable at each x 6 [a,b] and that H" is m times Peano difl'erentiable at y. Then H is (n + m) times Peano differentiable at y, and H(n+m)(y) = (Hn)m(y)- This lemma was first proved by Corominas in [4]. We will use ideas of his proof to prove this lemma, but before we give the proof we need some other properties of Peano derivatives that are known. The following definitions and Lemma 3.1.5 are due to Oliver. (See [10]). We will give a simpler proof of Lemma 3.1.5, than is given in [10]. Definition 3.1.4 If f has an n-th Peano derivative at each point of an interval [a, b], we say that f satisfies the mean value theorems M1,“, 1: = 0,1, . . . , n —l (or that f 6 M5), iffor each x and x + h 6 [a,b], there is an x’ between x and x + h such that: ms + h) — fk(x) — h fig) _ ... _ z—fiili-ll)!.f.-1(x) = My). (1) (n-k)! When n = 1 and Ir 2 0, we have the ordinary mean value theorem for first derivatives. The mean value theorem (Lemma 3.1.5 below) is that if fn exists on some interval [a,b], then f 6 M3,”, 1: = 0,1, . . . ,n — 1. 44 The special case of M: when the left hand side of (1) equals 0, we refer to as Rolle’s Theorem, Hi. In the usual manner, M: follows from Rf, by adding a suitable polynomial to f. If f has an n-th Peano derivative on [a,b] and if y and y + h are given in [a, b], for each x E [a, b] we set my + h) — 2:: (gratis) . (x _ y)" g(x) = f(x) _ hn'k/(n — k)! n! It follows immediately that hn—k-l 9:.(11 + h) - 9:.(31) - h9k+1(3/)-"' - (n _ k_1),gn_1(y)=0 i.e., that g satisfies the hypothesis of Hi; and that fk(!/ + h) - 23:1: fifflyl 9’10”): fn(3’) " hn-k/(n __ k)! ’ Applying the conclusion of Rf, to g, i.e., replacing x by x’ and gn(x’) by O, the conclusion of M]: follows for f. It is also possible to deduce Rfi, k = 0,1,...,n — 2, from R”‘1 and M,’f_1, as n follows. We may write fk($ + h) —' fk(x) - ' ‘ ° -' (—"_—::;k 22)Ifn-2(x)— (—_—::;:.11)gfn—l($) hn-k/(n _ k)! =0 in the form I (um-1 (r)----—(—;—W’:."."‘_".J.- (z) k h:—h-I](n_k-1)§ 2 " fit-1(3) _ 0 h/(n — k) ‘ and replace the first ratio, using M: _1, to obtain fn-1(x ”)— fn_1(x)= O for some x” between x and x + h. We use R3" to deduce from this last equation the existence of 15’ between x and x” for which fn(x’) = 0, the conclusion required by Rfi. Lemma 3.1.5 Letn E N and let f be a function defined on some interval [a, b] so that fn exists on [a,b]. Then fn satisfies the mean value theorems M,’f, k = 0,1,. . . ,n — 1, on [a,b]. 45 Proof: The proof is by induction on n. Since the assertion holds for n = 1, the remarks before the lemma show that the induction will be completed by proving f 6 R2“. So it is enough to prove that if fn-1(x + h) = f,,_1(x), then there is an 93’ between x and x + h for which fn(x’) = 0. We may assume h > 0, because a proof for the case h < 0 is similar. If fn is positive on [x,x + h], then, by Corollary 2.1.21, f("l exists and hence by Rolle’s Theorem for derivatives there is an x’ between x and x + h so that f(“)(x’) = 0. Similarly if f” is negative. If fn takes on both positive and negative values, then since f" is Darboux, fn attains the value zero. The induction is complete. C] Now we are ready to prove Lemma 3.1.3. Proof: For each x E [a, b] let T(1‘) ___ H(JS) _ $05—31 jHJ(y__)_ wi(_ y)n+i (Hn)i(y) (2) i=0 i=1 (T! + 2)! Then To) = My) = = any) = 0. Tum = H.(z)— 2.-..(x- vii—W"; " and Tn(y) = (Tn)1(y) = - - ~ = (Tn)m(y) = 0. Since T 6 M2, for each x 6 [a,b] there is a c, E [a, b] between x and y such that Tm = (x ;!y)"r,.(c,). (3) Since Tn(y) = (Tn)1(y) = - - . = (Tn)m(y) :2 0, we have Tn(x) = (x — y)mem(y,x — y) where em(y,x — y) —> 0 as x -—) y. (4) Combining (3) and (4) we get T0”) = (x _ y)n+m%€m(yacx — 31): Now ( 2) becomes LI(—$) Z($_ y)j__ HJ(y) + Z($_ y)n+i (Hn)i(.y_____)+ +1:( __ y)n+mC ;+m(y,x _ y) j=0 i=1 (n + 3)! 46 where (“mg/,3; __y) = fifigemwm, -— y) —§ 0 as x —+ y. This proves that Hn+m(y) exists and equals (Hn)m(y) . EJ Lemma 3.1.6 Let f be defined in an interval [a,b] containing 0. Suppose that the lc-th Peano derivative of f at 0 exists, and that the l-th Peano derivative of f exists on [a,b], where lc and l are positive integers with l 3 Ir — 1. Also suppose that f(0) = f1(0) = = fk(0) = 0. Let g(y) = y‘(""). Then the function h defined by h(y)=(:,)f(y —(;)/ mg was. —1)’(i)/.y/.n"'2:f(t)g<'>(t)dt-~dx. fan/790, and h(O) = 0 has an l-th Peano derivative on [a, b]. Moreover w:{ i? ;;;:3 Proof: By assumption f (y) = ykek(0, y). Consequently all of the above integrals are integrals of continuous functions. Hence h is well defined. Moreover for y 96 0, y 6 [a,b] H(y) = f: (f2 3”" f(t)g(‘)(t) dt~ udxg i = 1,. . . ,l is i times ordinarily differentiable and H(‘)(y) = f(y)g(‘)(y) for i = 1,...,l. By Lemma 3.1.1, fg“) is l — i times Peano differentiable at y. Therefore by Lemma 3.1.3, H is l times Peano differentiable at y and H¢(y) = (H(‘))1_;(y) = (f(y)g(’)(y))(;_,). Hence h is 1 times Peano differentiable at y and y)=§(-1)J (3190)) is) and by Lemma 3.1.2, h1(y) = fz(y)g(y)- 47 It remains to prove that h)(O) exists and that h)(O) = 0. For y 96 0 hill = % {(é)y’ek(0.y)+(k-I)( (’) )/,,y""€k(0,t)dt+~-+ (k—I)(k—l+1)-(k—1)(,)/y/n-/%e,.(0t)dt .132}. Hence lim——— h(y) =0. Therefore h(O) = h)(O) = = h)(O) = 0. E] v-*0 y Now suppose that f has an l-th Peano derivative in an interval [a, b] containing x, and that fk(x) exists. Consider a function T(y) = f(y) — f(x) - (y - x)f1(9=) - - (y - x) T and its translate C(t) = T (x + t). Then G satisfies the hypothesis of Lemma 3.1.6 and by that lemma the function H defined by H(y) = (3)G(y)g(y) - (1) joy G(i)g’(t) dt + . . . + _1)l [) [01' /0x2 . - ,v/OW G(t)g(l)(t) dt . . . ([32 for y 7g 0 and H (0) = 0 has an l-th Peano derivative on x — [a,b]. Moreover by the same 9&9? if y if 0 = y - lemma, H1(y) { 0 ify = 0 . But Gz(t) = T)(t + x) = f:(t + x) — f)(x) — tf1+1(x)— —t"" ”£17,. Therefore we have proved the following theorem. Theorem 3.1.7 Suppose that a function f defined on an interval [a,b] containing a point x has an l-th Peano derivative on [a, b] and a k-th Peano derivative at x, where 0 S l S h. Then the function F defined on [a,b] by 1()-";‘1'——31 (z) . 0 ify = x is an l-th Peano derivative. 48 Corollary 3.1.8 Suppose that a function f defined on an interval [a,b] containing a point x has a (k — 1)-th Peano derivative on [a, b] and k-th Peano derivative at x. Then there exists a perfect set P C [a, b] of positive measure such that x is a bilateral point of accumulation of P and lim fk—1(y) ‘ flu—1(3) 146 P. y-w y — = fk($)- Proof: The function F from Theorem 3.1.7, applied with l = k — 1 is a (lc — 1)-th Peano derivative and hence Baire 1, Darboux and has the Denjoy property. Therefore, by Corollary 2.1.18 there is a perfect set H such that F is continuous at x with respect to H. Since F has the Denjoy property there is a perfect set P of positive measure, containing H, so that F is still continuous at x with respect to P. The set P satisfies the assertion of the corollary. C] 3.2 Peano derivatives and Property Z Property Z was introduced in [18] by Weil. He proved, that f), has the property Z at every point of R. In [13] Marik gives a different proof of this fact. Moreover he proved there that a k-th approximate Peano derivative has property Z. Also he generalized this result to an assertion which when specialized to lc—th Peano derivatives is the following theorem. Theorem 3.2.1 Let j and k be integers, 1 S j S 1:. Let x E R and let f be a function such that fk(x) exists. Define P(y) = f=o(y —x)’—‘-§,—‘52 for y E R. Let c > 0, 17 > 0. Then there is a 6 > 0 with the following property: If I is a subinterval of (x — 6,x + 6) such that fj exists on I and that IfJ-(y) — P(j)(y)| _>_ ely — xlk‘j for all y E I, then m(I) S nd(x,I). 49 For the special case of lc-th Peano derivatives, the proof of Mafik’s result is simpler than for approximate lc-th Peano derivatives. Moreover the proof given here concludes the case 1 S j S k — l as a consequence of the casej = lc. The casej = k, using Theorem 3.1.7, is an immediate consequence of property Z for Peano derivatives. Proof of Theorem 3.2.1: Casej = k. Let g(y) = f(y) — P(y), and let 6 > 0 be such that _ ml]: (5) my» < «liqr'y a Let I be a subinterval of (x — 6, x + 6) such that lgk(y)| Z 6 for y E 1- (6) By the Darboux property, either gk(y) 2 e on I, or g), S e on I. By Corollary 2.1.21, g“) exists on I, and hence by Lemma 2.1.7, there is a subinterval J of I such that Igl 2 . - @915 (7) Combining (5) and (7) give m(I) < fi—nly — x] for every y E J. Therefore m(I) S T1? - (m(I) + d(x, 1)), hence m(I) S n - d(x, I). Casej < k. By Theorem 3.1.7, the function hJ-(y) = W”- for y 9i x and h,(x) = 0, is a j-th Peano derivative, and by what was just proved, for any 6 > 0 and n > 0 there is a 6 such that whenever I is a subinterval of (x — 6, x + 6) and such that |hj(y) — h,(x)| 2 e for y E I, then m(I) S 17 - d(x, I). This is exactly the claim of the theorem for j < lc. E! This theorem enables us to prove the following analogous of Theorem 2.1.15. Theorem 3.2.2 Let k E N and let f be a function defined on R with f),(x) existing for all x E R. Then for each integer 1 < r S k — 1 there is a bilateral nonporous system of paths E = {E,, : x E R} satisfying the I. C. condition such that — ’5‘"1.(!-_3li .3 h(x): lim fr”) 1:0 j! fr+J() veExw-n' (y — x)"" 50 The proof of this theorem is similar to the proof of Lemma 2.1.11 and Theo- rem 2.1.9. Proof: Let c > 0 and let 6(6) be the 6 from Theorem 3.2.1 applied with r] = 6/2. Let {6)} be a sequence so that 0 < 6) S 6(1/l) and 6H1 < 61/2, and define the set E; by 00 Z _ 6:6'11flli r J ={x}UIL_J{z>x+61+2: fr( ) zf:-x)""' f+(x x)—fk(.’t) < l/l}. It is certainly true that k—r— 1 y-xJ It) fk($) : “m fr(y) J: LTLfr+J() yEE;.y—~z _(y — 2:)" " Now we will prove that E; is nonporous on the right at x. Suppose not. Then there must exist a number 1/2 < 0 < 1 and a sequence of numbers h; l 0 with (x + 9h), x + h)) n E; = 0 for every index I. Choose an integer lo larger that (1 — 0)"l (i.e. so that if I 2 lo, then 1— l/l > 0) and let jg be the first index for which h,0 < 610. Fix l so that 614.1 S h,0 < 61, and note that I 2 lo. Since h,o < 61 S 6(1/l), by Corollary 3.1.8 there must be a point z with x + (1 — 1/l)h,-0 S z S x + h,o such that rz- k—r—1(Z_-j;_r_£rj$ m (2%)“ m )—f;.(x) <1/I. We then have the inequalities, x+61+2 6(y) > 0 so that (y — 26(y),y + 26(y)) fl H(f, M, 1) = {y}. Let P, be a perfect set containing y so that y is a bilateral point of accumulation of P, satisfying lim fk-1(Z) '- fk-1(y) ZEPyszfly Z - y = fk(y) 52 and fk-1(Z::;"-l(y) _ h(y) S 1 for every z E Py- Corollary 3.1.8 assures the existence of P,. If P, H (y + "—11-,y + %) 9i 6 , for n 6 Z \ {—1,0}, then by the Baire category theorem there is a perfect set Qn(y) C P, 0 (y + that] + fi), there is Mn(y) E N with Qn(y) C H(f,M,,(y), 1). Let Qy = U Qn(y) n (y — 62(.11).:1 + 62(31)) U {y}. and let n€l\{—l,0} HM = H(f,M,1)U {Q,: y E H(f,M,1), y is isolated in H(f,M,1)} Theorem 3.3.1 H M is a perfect set, and f)._1 is differentiable on H M relative to H M with (fk_1|HM)’(x) = fk(x), for each x 6 HM. Proof: By the construction of H M we see that no point is an isolated point. Note that each Q, is perfect and that Q, 0 Q, = 0 if y,z E H ( f, M, 1) are two different isolated points of H ( f, M, 1). Let {2,} be a sequence in H M such that limn_.°o 2,, = 2. If 2,, E H(f,M,1) for infinitely many n, then 2 E H(f,M,1) since H(f,M,1) is closed. Assume 2,, not in H ( f, M, 1) for each n E N. Then for each n 6 N there is an isolated point yn 6 H ( f, M, 1) such that 2,, E Q,,,. If there are only finitely many different y,,, then 2,, E Q, for infinitely many n. Since Q, is closed, 2 E Q, C H M. Assume there are infinitely many different yn. Since Izn — ynl < 6(yn) < 1, and since “an“, 2,, = 2, there is a subsequence {2",} of {2,} such that {ynj} converges. Let y = limj.”o ynj. Then y E H(f,M,1) and it follows that z = lim,-_.co 2n]. = y. So 2 E H M» Therefore H M is closed. Now if x E H M is an isolated point of H ( f, M, 1), then clearly f,’,_l at x relative to HM, exists and is equal to fk(x). If x E Q, for some y E H(f, M, l) where y is an isolated point of H(f, M, 1), then there is n E Z so that x 6 Qn(y) C H(f,M,,(y), 1) 53 and by the fact that there are two numbers a < b so that (a, b) 0 HM = Q,,(y), we see that f,’,_1 at x relative to HM exists and is equal to fk(x). Finally let x E H(f, M, 1), and x not an isolated point of H(f, M, 1). Let e > 0 be given. Then there is e > r] > 0 so that fk-Ify) — fie-1(3) y — x — fk(x) < 6 whenever y E H(f, M, l) and Iy — x] < 7}. Let y be an isolated point of H(f, M, l) and let 2 E Q, with |z -—- x| < 17/2. Since ly - 2| < 62(y) < 5(31) and Lu —e| > 25(31), we have 71/2 > le - 2| 2 le — yl - ly— 2| > 25(31) - 6(y) = 5(31). Hence ly — el S ly - 2| + lz - e| < 5(11) + n/2 < 77. Thus "‘4”: : ;"“(“’) — fk(x)l = (”“02 1"”) — h(x)) 3’ j “’+ (fern-1e) — h(y)) :1 + :: gm) — h(x)) s fk-l(y3::k-l(x) — fk($)[l1— :1: + "4‘": : f4“) - My) 41+ :1] use» + mm» s my) .93) 62m ‘(H 6(y)) +1 6(y) + mums 26 + 6(y)(1+ 4M) 3 2.: +%(1+ 4M) and since 6 was arbitrary we have that f,;_l at x relative to HM exists and equals fk($). D We end this chapter showing that a k-th Peano derivative is a path derivative of the (k — 1)-th Peano derivative with a system of paths satisfying the I.I.C. condition. As a corollary to this result we will obtain that a. k-th Peano derivative is a selective derivative of the (k -— 1)-th Peano derivative. 54 To define the system (E, : x E R} of paths with respect to which a given k-th Peano derivative, fk, is the path derivative of fk_1, we begin with some notation. Notation For x,y 6 R let 6(x,y) = min{1, [lg—’11}. For x E R and M E N let R, = U{P, n [y,y + 62(x,y)) : y E H(f, M, 1) and y is right isolated from H(f,N,1) for N E N} and let L, = U{P, n (y — 62(x,y),y] : y E H(f, M, l) and y is left isolated from H(f, N, 1) for N e N}. Definition 3.3.2 Let x E R. If there is an M, E N such that x is a bilateral point of accumulation of H(f, M,,1), then let E, = HMxUR,UL,. If x is a right isolated point of H ( f, M, l) for every positive constant M but there is an M, so that x is a left point of accumulation of H(f,M,, 1), or ifx is a left isolated point of H ( f, M, 1) for every positive constant M but there is an M, so that x is a right point of accumulation of H(f, M,, 1), let E,=HM,UP,UR,UL,. Finally if x is an isolated point of H (f, M, 1) for every positive constant M then let M, = 1 and let E,=HM,UP,UR,UL,. Definition 3.3.3 Let E be the system of paths {E, : x E R}. Lemma 3.3.4 Let k E N and let f be a function defined on R such that fk(x) exists VII 6 R. Then E is bilateral and satisfies I.I. C. condition. 55 Proof: Clearly E is bilateral. We will prove a stronger condition than I.I.C. . Namely we will prove that for any two points x and y E, n E, F] (x,y) .-,£ 0. Let x < y be any two points. Suppose M, S M,,. If x is a right point of accumulation of H(f,M,,1)C H(f,M,,l), then E,fl E,fl (x,y) 75 0. If x is a right isolated point of H(f,M,, 1), then by choice of M,, x is a right isolated point of H(f, M, l) for every M E N and x E H(f,M,, 1). Thus 0 e P..- n [m + 62m» n (x,y) c e. n E. n (x,y). If M, > M, and if y is a left point of accumulation of H(f,M,, 1) C H(f,M,, 1) then E, F) E, 0 (x, y) 75 0. If y is a left isolated point of H ( f, M,, 1), then by an argument similar to the above E, n E, n (x, y) 96 0. Therefore E satisfies the I.I.C. condition. Cl Theorem 3.3.5 Let k and f be as in Lemma 3.3.4. Then fk_1 is E differentiable With f(k—l)E(-’”) = f),(x). Proof: Let x E R, and e > 0 be given. Then there is an e > 17 > 0 such that fk_1(y) - fie—10'?) y — T - fk(x) < 6 whenever |y — x] < r) and y E H(f,M,,l) or y E P,. Let 2 E E, be such that [z - x| < 321. If 2 E P, for some y E H(f,M,,l) such that y is an isolated point of H ( f, M, 1) from either left or right, and for every positive constant M, then 52‘ > [z - x| _>_ Ix — y| — |y - 2| 2 26(x,y) — 6(x,y) = 6(x,y). Therefore Iy - x] S Iy — z] + Ix — z] < 6(x,y) + 17/2 < 17. Hence fk-1(y) - fie-1W) y - x - h(x) < e (8) 56 Thus fk-1(Z) - fk-1($) _ Z—$ h(x) —f. 0, 6 > 0 let Pa = P110515) = {e E A, |€q+n(e' h)I < 6. for I’ll < 5} (2) 57 58 Note that if x E A,, then x E A, for every p Z q. Also x E A, iff F(-q) has an (q + n)-th Peano derivative at x with (F(‘9)),+,,(x) = F[,](x). Lemma 4.1.3 For q S p, P,(e,6) C P,(e :1: !,6). Proof: Let x E P,(e,6). Then x 6 A, and for t E R (- Q+Jl F(- (”(3 + t).— _ (’2th + Z tq+j F_lJ__l($)' + tq+n6 [,q.],n(x,t). (3) j=0 1' j=0 (q + 1') Integrating both sides of (3) from 0 to h we get F(‘q’1)(x + h) — F(""1)(x) = q-1 :H)——(—$) n 2),: th+1 +2}, hq+1+j___ +/h tq+n Min )dt. Thus x 6 A,+1. By(] the remark after Definition 4.1.1, we have hq+1+n 65:31,“ (x h) =/: tq+ne 5"],(x, t)d (4) and since x E P,(e, 6) for 0 # |h| < 6 from (4) we have |h|q+l+n lhlq+l+nlclqfilil($a M] < jlh'tq+"e dt = e——— . q n o q +1 + n Hence |c[9331,(x, h)| < e/(q + n + 1) whenever Ihl < 6. Therefore P,(e,6) c P,“ (adj: :)l)!’6) . The general result follows by induction. [:1 Definition 4.1.4 For x E A, and fori = 1, . . . ,n, define egqlxx, h) by F(-q 0+J') F("’)(x+h)- 21.1—— flm +21: 11"” (Ffl—J—j) ), +h°+‘ elflxx, h). (5) j=0 j=0 59 Note that eflfix, h) doesn’t depend on which q-fold indefinite Riemann integral, F (”1), of F is taken. The following formula follows directly from Definition 4.1.4. Formula 4.1.5 Let x E A,. Then fori E N with 2 S i S n we have ngl,_l($,t)= t(q quj——::}I+ “iii-T t)’ Recall Lemma 1.1.3, Definition of Riemann difference A?f(x) and Lemma 1.1.4 from Chapter 1. Lemma 4.1.6 For m E N the following holds: m . . 0 ifi=0,...,m—1 Z(—1)m-J(';:)j: = [ m! ifi = m i=0 Ig-(m+1)! ifi=m+1 Definition 4.1.7 For any function f defined on R the Riemann difference Ag"f(x) at a point x, of order m is defined by AM =Z(-1)"’(’?fl)(e+fl> The relationship between A? and A?“ is given by the following assertion. Lemma 4.1.8 For any function f defined on R, for any m E N andt E R we have A2"”‘f(e) = AI"f(=v + t) - A7705)- 60 Lemma 4.1.9 Let x E A,. Then for each i = 1, . . . ,n A3+mF(—9)($) = t°+mF1m1(x)+tq+m :3:3‘<—1)°+m-J' (IEM)j°+me£°lm(x.jt) ifm =z‘ ,.+.~ ::r(-1)q+m-i(”jm)jv+‘e.~(x.jt) ifm > 2'. Proof: 9+m . A3+mF("')(x) = ;(_1)Q+m-J (qtm) F("’)(x + jt) J: = if (_1)q+m-I (07") (:0 0151:3121. j=0 l=0 :0 W (F "1, ,——(%+ +(jt)q+‘e£°l.(x.jt)) 1:0 115%- q+I)__(_,) 9+": =1)”: ——,.—— zu- 1...-.(..-),.+ l=0 J=0 F (33)"+ q+l____ l’l 1)q+m—j q+m q+l gt q+I)!,-§,((_1) (, )’ + .Q+m . tH-a Z(_1)q+m-J (9'37”)jq+i 41914:; jt) i=0 which by Lemma 4. 1. 6, is equal to tq+mF[m 1(2)) + tq+m zq+m(_1)q+m—J (9+m)jq+m€q+€l¢1lm($,jt) if m _____. Z tq+i 29+m(_1)q+m-j (9+m)jq+i6q+l01i($, jt) if m > i. [1 Formula 4.1.10 Let x,x +t E A, and let i E N with 1 S i S n. Then 9+i+1 F[:1(9=+t)-Fm 2:) Z(- ,,.+.-+1_.(.+;;+1),.+. 511:: 36+ J=0 q-l-i q+i P(_ 1) q+i—J (q+i)] q+i 651],” jt) _£(_ _1) q+i-J' (q-H), «1+5 file] +_(,, +t jt). J=0 61 Proof: By Lemma 4.1.8, A3+’+1F(‘°)(x) = A?+’F("’)(x + t) ._ A‘,’+’F(’q)(x). (6) Applying Lemma 4.1.9, with m = i + 1 to the left hand side and with m = i to the right hand side of (6) we get q+i+1 tq+i z: (_1)q+i+1-.i (q+;+1)Jq+i [alga jt)- =0 . n q+£ o . - twig—1,1” + t) + W“ Z(_1)Q+t-J (9:1)jq+i£[Q]i(x + t,jt) _ i=0 s . q+£ . . . [ ] tq+'F[,](x) _ tq+t Z(_1)q+t—J (Q:I)Jq+i€ 6:“(3 jt) i=0 Dividing both sides by t?“ gives the above formula. C] Theorem 4.1.11 For any interval [a, b], F[n] is bounded on P,(e, 6) f) [a, b]. Proof: Let [a, b] be an interval. Let x, y E P,(e,6)fi [a, b], so that for t = y —x we have It] < 6 / (q+n+ 1), and let B: 2"?“ (q+g+1)jq+“. Then the right hand side of Formula 4.1.10 applied with z = n, is bounded by 3315. It follows that F[,,] is bounded on P ,(e,6)fl [a, b]. From Formula 4.1.5 (applied with i = n) it follows that for |h| < 6, |eL°1,_,( . , h)| is bounded on P,(e,6) 0 [a,b]. Now from Formula 4.1.10 (applied with i = n — 1) we see that F[,,_1] is bounded on P,(e,6) n [a,b], and again going back to Formula 4.1.5 (applied with i = n — 1) we see that for |h| < 6, I491“, ( ,h)] is bounded on P,(e, 6) 0 [a, b]. Continuing we can deduce that there is a constant C so that |F[q(z)l S C for 1 S i S n, for x E P,(e,6) 0 [a,b]. Let x E P,, and let {xm} be a sequence in P,(e, 6) such that lim,,,_.,° xm = x. Choose p 2 q such that x e A,. Let [a, b] be such that {xm} C P,(e, 6) n [a, b]. From the first part of the proof we see that for l S i S 11, FM is bounded on P,(e, 6) n [a, b]. 62 Therefore we can choose a subsequence {xmj} converging to x such that {F[,-](xm1)} converges for each 1 S i S 11. Let these sequences converge to G;(x), i = 1,... ,n respectively. Let [h] < 6, and, as we may, suppose that |h+x —xm,I < 6 for everyj 6 N. Since 6 P,(e, 6), by Lemma 4. 1. 3 we have |ep+,,(xmj,h + x - xmj) )I < 612—3131. Thus we may also suppose that the sequence ep+,,(xm1, h + x— me) converges. Denote its limit by E(h). Now letting j -—-1 oo in the formula F(-P)(m + h) : F(-P)(xm1) + (h + z _ 3m, )Ff-P+1)(xm1) + . . . + (h+x—xm,)” (h+x—x,,,1)'°+""l pl (p+n— l)! +(h + 1' - $m1)p+n ( F[0](xm,) 'i" + Fl"1($m1) (p + n)' F'[n—l]($m, )‘i' + e,+,,(xm1, h + x — me)) we get 1"“ h: p+1 _ hp F( p’($+h 12? F’ ”m(x+';!$F[o]()+mG1($)+m+ (1):”: 1) G.-.( )+hp+" ((1)—70:“), +1300) (7) + E(h) 1s bounded, by the uniqueness of Peano derivatives from (7) Gn x p+n ! we have G,-(x) = Fm(x) for 1 S i S n — 1 and Since . + E(h). (8) Since |E(h)|_ —..+m) 9‘1 I —s QFK—q+0(x)q+n-1 n— -' n- 's =Zz(1_$)($l )—l )tT Z (_1)q+ 1 1(q+j 1)] + l=08=0 ‘ 1:0 71 0+1 F( ) q+n— —1 9+1 2: _ III 0+l-8t8___ [ll( 1? 1) q+n—l -J (Hm—1 , g§(s)(l ) (q+l)l12(- (1)]. by Lemma 4.1.6 the above 18 equal to n q-H )(m F100”) q+n— —1 Z Z (9:1 —x)"+"‘t‘(q+l)!1_ Z: (_1)q+n—l-j (q+n-1)ja. l=n-l s=q+n—l Applying Lemma 4.1.6 once more it is equal to q+n—1 2 tq+””F[,,_1](x) + tq+"’1(x1 — x)F[,,](x) + t"""’ F[,,](x). This completes the proof. [I Proof of Theorem 4.1.12: The proof follows directly from Lemma 4.1.13 and Lemma 4.1.14. 1:: 66 Finally we are ready to prove the main result of this Chapter. Theorem 4.1.15 Suppose for each a: E R F is n-th generalized Peano diflerentiable at 1:. Then F[n—1] is diflerentiable on 75., = 75.,(£,6) relative to 75., with Fin-III'FJI’) = Fm”)- Proof: LetzEFq. Thereisap>qsothathAp. Let 1 >c’>0begiven. There 13 0 < 17 < 6 such that |£p+n($, h)| < 6’ whenever |h| < 17. Let {mm} be a sequence in Pq converging to a, so that [am — ml <-—"— p+ By Theorem 4.1.11, there is n a constant C so that |F[,,](:rm)| _<_ C, for every m E N. Let t = (mm — “6'71“". Then ljtl < 5 and Iain. -$+jt| < 17, forj = O,1,...,q+n — 1. Therefore Ielflna 2m — x +jt)| < e (11) and by Lemma 4.1.3, leyln(xm,jt)| < e for everyj = 0,1 .. . ,p+ n — 1 and for every m E N. (12) Since 1:". E A,,, the formula from Theorem 4.1.12 gives Fin-11(xm) - Fin—ll($) x _ x — was) s gear—31.11“”), + ”fl (”‘2' )(1 72:1”: “”1 (x mm — x+jt)| + j=o c p+n 65+» P—+’; 1'14me +2: (-1)’+"- _,( +"' )1?“ Human . =0 By (12) and (11) together with Theorem 4.1.11 the above is s 'avflLllnn] )1 + ,Zz; (”*2“)(1+je';+‘-)r+- + £131:{P+_n___—IC+pEl(P+n—1)Jp+n6}. i=0 67 Since 6’ was arbitrary we have Flu—1]($m) - Flu—1](x) — F[n](a:) —» 0 as mm 6 Pq, mm —> 2:. zm—a: Now for the general case let {2cm} be a sequence in P, such that 3:", —-> 2:. Let gm 6 R, be such that Iym — zml _<_ Tiilxm -— xl and that F(n—1](ym) — F[n_1](l'm) ym"mm — F[n](xm) S 1- (13) By what was just proved, there is such a sequence ym. By Theorem 4.1.11, there is a constant C such that |F[n1(:1:m)| S C for every m E N. This and (13) give Flu—l](ym) — F[n—1]($m) S. 0 +1. (14) gm '- 3m Now Flu—1]($:) — flu—H(x) _ F[n](x) = Flu—11(xm) '— Flu-1](ym) mm - ym+ m "' 3m — 3117: 3m "" 33 Fin-11(ym) - F[n_11(x) ym - 1: mm - ym { ym _ 33 FM”) mm _ x - Fln]($) mm _ x 80 by (14) Flu—11(3) fin-11”) ..W s 1 Fn— (ym) — Fri—1(a)) l 1 C 1 —— l ‘1 l l — Fn — —. ( + )m+ ym_z [1($)(1+m)+0m (15) Finally since arm —-) 2:, ym -+ 1'. But gm 6 Pg, and hence by the first part F[n-l](ym) "' Flu-110‘?) ym-x - F[n1(17) —* 0. (16) Therefore by (15) and (16) F[n-1]($m) ’ Flu-11(3) zm‘x -F[,,](:c) —+0as :cm 67’}, mm —* :c. This completes the proof. D 68 Lemma 4.1.16 For each 6 > 0, U°° °° Pq(e,1/m) = R. q=0 m=1 Proof: The assertion follows from Definition 4.1.2. D Corollary 4.1.17 Suppose for each a: E R F is n-th generalized Peano differentiable at 2:. Then Fla] is a composite derivative of F[n_1]. Corollary 4.1.18 Suppose for each a: 6 R F is n-th generalized Peano difi'erentiable at 1:. Then Fin] is an approximate derivative of F[n_1] a.e.. Corollary 4.1.19 Suppose for each a: E R F is n-th generalized Peano differentiable at 2:. Then Fin] 6 [A]’. Corollary 4.1.20 Suppose for each a: E R F is n-th generalized Peano difl'erentiable at at. Then F[n] is a Baire 1 function. That FM is a Baire 1 function, was proved in [9]. The proof in that paper is not as simple as the proof for Peano derivatives. Corollary 4.1.20 gives another proof of this assertion. 4.2 Generalized Peano, path and selective derivatives Next we will show that the following analogy of Theorem 2.1.9, holds for generalized Peano derivatives. Theorem 4.2.1 Let I E N with l S n — 1. Assume for each function g defined on a closed interval I having an l-th generalized Peano derivative, gm, on I, 9U] is a Darboux function and if 9U] 2 0 on I, then 9U] = g“) on I. Suppose F[n] exists on R. Then there is a bilateral nonporous system of paths E = {E, : a: E R} satisfying the LC. condition such that FM is the E-derivative of F[n__1]. 69 We will need some lemmas before we prove this theorem. Lemma 4.2.2 Under the assumptions of Theorem 4.2.1 for every 6 > 0 and n > 0 there is a 6 > 0 such that ifI is a closed subinterval of (x — 6,x + 6), a: is not in I with F[n—1](y) - Flu—1K9?) y — a: for all y e I, then m(I) S nd(:r,I). — Flatt) 2 6 (17) Proof: Let 6 be chosen according to Theorem 2.1.5, applied with 17 replaced by 171 = n/(1+ 1)) and withj = n — 1. Let I be as above, and let g(y) = F(y) — yn‘lffi—Z’i—g? — (y — wring-:2. Then g has an (n — 1)-th generalized Peano derivative and g[,,_1](y) = F[,,_1](y) — F[,,_1](:r) —- (y — x)F[n](z). So by assumptions g[,,_1] is Darboux. By (17) |g[,,_1](y)| 2 cly — 2| on I. Since a: is not in I, |g[,,_1](y)| > 0 for y 6 I and since g[n-l] is Darboux, we have either gln-” > 0 or —g[,,_1] > 0 on I. Hence by the assumptions, g[,,_1] is an (n - 1)-th ordinary derivative of g on I. Therefore F is (n — 1) times ordinarily differentiable on I and by the uniqueness of generalized Peano derivatives, F ("‘1) = F[,,__1] on I. Now we can apply Theorem 2.1.5, with n = k and j = n -— 1, which gives m(I) _<_ m . (m(I) + d(x, 1)) Hence m(I) S Ud($a1)- [j The statement and the proof of a next lemma follow line by line the corresponding Lemma 2.1.11 for Peano derivatives. So we will only state the lemma and omit the proof. Lemma 4.2.3 Under the assumptions of Theorem 4.2.1, for each point :1: E I there is a path E, leading to a: and nonporous at a: so that lim Flu-1101)" Fin—1K3) z Fin](x)- yEExm-w y — a: 70 Lemma 4.2.4 Let m _<_ l be two positive integers and let 6 > 0. Then 1 1 Pm(€a —) C H(Ca—) ' m 1 Proof: By Lemma 4.1.3, Pm(e, -1-) C P1(£, —1-), and by the Definition 4.1.2 m m H(é’r—ii) C H(Cail- [:1 Now we are ready to prove Theorem 4.2.1. Proof: For each a: E R let E; be a path satisfying the assertions of Lemma 4.1.9. We will define the system of paths E = {E,, : a: E R} as follows: For a: E R let E3 = E; U Pm(1,1/m) where m is a positive integer such that a: E Pm(1,1/m). That E is nonporous (therefore bilateral) follows directly from Lemma 4.2.3. Also Lemma 4.2.3 and Theorem 4.1.15, assure us that F[n—1] is E differentiable with F[,,_1]|’E(z) = F[,,1(a:) for every 1: E R. It remains only to prove that E satisfies the intersection condition I.C.. We will prove that for any two different points a: and y, E3 0 EV n [:c, y] at 0 which is stronger than the LC. condition. Let x E Pm(1,1/m) and y E P1(1,1/l) be any two distinct points. If m S I, then by Lemma 4.2.4 Pm(1,1/m) C Pz(1,1/l) and hence a: E E”. Similarly if m 2 I, then y E E3. Therefore E, 0 E3, 0 [x,y] 3‘5 0. Hence E satisfies the LC. condition. This completes the proof of Theorem . E] Next we indicate that we can drop the assumption concerning the arbitrary func- tion from Theorem 4.2.1. The proof of that assertion follows line by line the proof of Theorem 2.1.15, which is a corresponding result for Peano derivatives. So we will only state the result. 71 Theorem 4.2.5 Let F be a continuous function defined on R so that Flu] exists on R. Then there is a bilateral nonporous system of paths E = {Err : a: E R} satisfying the I. C. condition such that Flu] is the E-derivative of Flu—1]- Now using properties of path derivatives we get the following corollaries: Corollary 4.2.6 Under the assumptions of Theorem 4.2.5 Fin] is Darboux. Corollary 4.2.7 Let F be as in Theorem 4.2.5, let [a,b] be an interval and a E R. IfFln] Z a (or F[n] _<_ a) then a) F[n—1]($) - are (0:13 — F[,,-1](:r) ) is nondecreasing and continuous on [a, b] b) F(") exists and F(") = F[n] on [a, b]. Corollary 4.2.8 Under the assumptions of Theorem 4.2.5 Flu] has the Denjoy prop- erty. Corollary 4.2.9 Suppose F[,,](a:) exists for all a: in Io and let M 2 0. If Flu] attains both M and —M on Io, then there is a subinterval I of Io on which Flu] = F ("l and F (“l attains both M and —M on I. We end this chapter showing that every generalized Peano derivative Flu] is a selective derivative of F[,,_1]. The idea is very similar to the one that we used for Peano derivatives. 72 Definition 4.2.10 Let P1, be a set containing y so that y is a bilateral point of ac- cumulation of P1, , lim Flu-11(2) "' Flu-11W) zEPy.z-w z — y and Pin—11(2) "' Fin-1](y) __ Z — ll Fln](y) S 1 for every 2 E P,. Theorem 4.2.5 assures the existence of P,. To define the system {Ex : :c E R} of paths with respect to which a given n-th generalized Peano derivative, Fin], is the path derivative of F[,,_1], we begin with some notation. Notation For 2,3, e R let 6(x,y) = min{1, Ltgil}. For a: e R and M e N let R, = U{Py F1 [y,y + 62(x,y)) : y E PM(1,1/M) and y is right isolated from PN(1,1/N) for N 6 N} and let L; = LJ{Py n (y — 62(z,y),y] : y E PM(1,1/M) and y is left isolated from PN(1,1/N) for N E N}. Definition 4.2.11 Let a: 6 R. If there is an M, E N such that x is a bilateral point of accumulation ofPM,(1,1/M,), then let E, = FM,(1,1/M,) u R. u L... If a: is a right isolated point of 7540, 1 / M ) for every positive constant M but there is an M,, so that a: is a left point of accumulation OfFM,(1,1/M,), or ifx is a left isolated point ofPM(1, l/M) for every positive constant M but there is an M, so that a: is a right point of accumulation ome,(1,1/Mx), let E, = FM,(1,1/M,,.) u P, u R, U L... 73 Finally if: is an isolated point ofPM(1, l/M) for every positive constant M then let MJr = 1 and let E, =P’M,(1,1/M,)UP.UR.UL,. Definition 4.2.12 Let E be the system ofpaths {Ex : a: 6 R}. Lemma 4.2.13 Let n E N and let F be a function defined on R such that F[,,1(:r) exists V1: 6 R. Then E is bilateral and satisfies 1.1. C. condition. Proof: Clearly E is bilateral. We will prove a stronger condition than I.I.C. . Namely we will prove that for any two points :1: and y E: n E, D (x,y) 75 0. Let a: < y be any two points. Suppose Mr S M,,. If a: is a right point of accumulation of FM,(1,1/M,) c Emu, 1/M,), then E, n Ey n (x,y) ;£ 0. If a: is a right isolated point of PM,(1,1/M,,), then by choice of M,, :r is a right isolated point of P540, l/M) for every M E N and a: E PM,(1,l/M,,). Thus H P. n [x,x + 62m» n (x,y) c E. n E. n (x,y). If M,,. > M,, and ify is a left point of accumulation ofPM,(1,1/My) C PMJI, l/MI) then E: D Ey fl (x,y) at 9. If y is a left isolated point of PM,(1,1/My), then by an argument similar to the above E, n Ey n (9:, y) ;£ 0. Therefore E satisfies the I.I.C. condition. El Theorem 4.2.14 Let F be a continuous function defined on R so that Flu] exists at every point 1: 6 R. Then F[,,_1] is E difl'erentiable with F};_,]E(:r) = F[,,](:r). Proof: Let :c E R, and e > 0 be given. Then there is an e > 17 > 0 such that 74 F[n—1](y) - F[n—l]($) y — 2 whenever Iy — :c| < n where y E PM,(1,1/M,) or y 6 P1,. Let 2 6 E, be such — PM“) < C (18) that Iz — 2:] < 321. If 2 E Py for some y E PM,(1,l/M,) such that y is an isolated point of PN(1, l/N) from either left or right, and for every positive constant N, then 321 > '2 — 2:] 2 la: — yl — Iy — 2| 2 26(x,y) — 6(x,y) = 6(x,y). Therefore |y — ml 5 |y — zI + In: — 2| < 6(x,y) + 17/2 < 7]. Hence by (18) F“- -F,,_ a: [ 11(y)_ [ 1]( )—F[,,](:c) y z < e. (19) Thus Fin—11(2) - F[n-1](-’B) z—a: — Flam Z—IB = “Flu-11h!) - Fin-11(3) y — (I? y _ 3 — Fin]($)) + Z Pin—11(2) - Fin—1K3!) Z — y " y _ — < ( z _ y my) 2 _ x + Z _ m(m(m m(m» _ FINN”) " 1'1"“1”) — m(m) 1 — z - y + y—x z—z Fin-11(2) "' Fin—11(9) Z — y - F[n](y) + z ' y I (|F[,,](:c)l + |F1n1(y)l) z—y z—z z—x By (19), Theorem 4.1.11 and the relationship between points x, y and 2 we get the above inequality Waxy) 506.31) 62W) + may) 3 c(1 + 50M!) 6(x,y) )+1- 4M_<_ 2: + 6(x,y)(1+ 4M) 3 2.: + £0 + 4M) where M is a constant from Theorem 4.1.11. Since 6 was arbitrary we have that FI’,,_1]E at :1: exists and equals to Fw(x) E] 75 Corollary 4.2.15 Let F be a continuous function defined on R so that Fin] exists at every point x E R. Then Flu] is a selective derivative of Fin-1]- Proof: Let a selection p(x, y) be defined as follows: If x < 3] let p(x,y) = z, where z is any point in E, 0 Eu 0 (x,y), if x = y, let p(x, x) = x. Then for fixed point x0 we have lim F[::-1](P(1'o,y)) - F[n-1]($o) = lim Flu-11(2) — F[n-1]($O). V“’° P(ico, y) - 30 ""° 2 - 550 Since 2 E E, we have that the above limit exists and equals F[,,](xo). CJ BIBLIOGRAPHY BIBLIOGRAPHY 1. S. Agronsky, R. Biskner, A. Bruckner and J. Marik, Representations of functions by derivatives, Trans. Amer. Math. Soc. 263 (1981), 493-500. 2. A. M. Bruckner, Differentiation of real functions, Lecture Notes in Math., vol. 659, Springer, Berlin and New York, 1978. 3. A. M. Bruckner, R. J. O’Malley and B. S. Thompson, Path derivatives: A unified view of certain generalized derivatives, Trans. Amer. Math. Soc. 283 (1984), 97-125. 4. M. E. Corominas, Contribution a la the’orie de la derivation d ’order supe’rieur, Bulletin de la Société Mathématique de France 81 (1953), 176-222. 5. A. Denjoy, Sur l’integration des coefficients difl'erentiels d’order supe’rieur, Funda- menta Mathematicae 25 (1935), 273-326. 6. M. Evans and C. E. Weil, Peano derivatives: A Survey, Real Analysis Exchange 7 (1981-1982), 5-23. 7. M. Laczkovich, 0n the absolute Peano derivatives, Ann. Univ. Sci. Budapest, Ebtvbs Sect. Math. 21 (1978), 83-97. 8. C. M. Lee, 0n Absolute Peano Derivatives, Real Analysis Exchange 8 (1982-1983), 228-243. 9. C. M. Lee, 0n generalized Peano derivatives, Trans. Amer. Math. Soc. 275 (1983), 381-396. 10. H. Oliver, The exact Peano derivative, Trans. Amer. Math. Soc. 76 (1954), 444-456. 11. R. J. O’Malley, Decomposition of approximate derivatives, Proc. Amer. Math. Soc. 69 (1978), 243-247. 12. R. J. O’Malley and C. E. Weil, The oscillatory behavior of certain derivatives, Trans. Amer. Math. Soc. 234 (1977), 467-481. 13. J. Mafik, 0n generalized derivatives, Real Analysis Exchange 3 (1977-78), 87-92. 14. J. Marik, Derivatives and closed sets, Acta Math. Hung. 43 (1-2) (1984), 25-29. 76 77 15. S. Verblunsky, 0n the Peano derivatives, Proc. London Math. Soc. (3) 22 (1971), 313-324. 16. C. E. Weil, 0n properties of derivatives, Trans. Amer. Math. Soc. 114 (1965), 363-376. 17. C. E. Weil, 0n approximate and Peano derivatives, Proc. Amer. Math. Soc. 20 (1969), 487-490. 18. C. E. Weil, A property for certain derivatives, Indiana Univ. Math. J. 23 (1973/74), 527-536. 19. C. E. Weil, The Peano derivative: What ’3 known and what isn’t, Real Analysis Exchange 9 No. 2 (1983-1984), 354-365. 20. A. Zygmund, Trigonometric series, 2’“ edition, Cambridge, (1959). MICHIGAN sran UNIV. LIBRARIES 1M11H)1|)HIWMIllIHIIIHWIWHllllllllllll 31293007914751