Text
                    HARMONIC ANALYSIS AND THE
THEORY OF PROBABILITY
BY
SALOMON BOCHNER
UNIVERSITY OF CALIFORNIA PRESS
BERKELEY AND LOS ANGELES
1955


PREFACE This is a tract on some topics in Fourier analysis of finitely and infinitely many variables and on some topics in the theory of probability and the connection between the two is a very intimate one on the whole. Although drafted in part earlier, more than half of the tract was actually written while the author was visiting, February-August 1953, the Statistical Laboratory at the University of California, Berkeley, of which Dr Jerzy Neyman is the Director, and a most delightful and profitable visit it was. Special thanks are due to Dr Loeve for listening patiently to expoundings of half-ready results, and to Mrs Julia Rubalcava, also of the Laboratory, for preparing the typed copy of the entire manuscript. S.B.
CONTENTS Chapter 1. Approximations page 1.1. Approximation of functions at points 1 1.2. Translation functions 7 1.3. Approximation in norm 9 1.4. Vector-valued functions 11 1.5. Additive set functions 12 1.6. Periodic additive set functions 18 Chapter 2. Fourier Expansions 2.1. Fourier integrals 22 2.2. Positive transforms. Plancherel transforms 25 2.3. Fourier series 28 2.4. Poisson summation formula 30 2.5. Summability. Heat and Laplace equations 33 2.6. Theta relations with spherical harmonics 36 2.7. Expansions in spherical harmonies 40 2.8. Zeta integrals 44 2.9. Zeta series 47 Chapter 3. Closure Properties of Fourier Transforms 3.1. Pseudo-characters and Poisson characters 52 3.2. Pseudo-transforms and positive definite functions 55 3.3. Poisson transforms 59 3.4. Infinitely subdivisible processes ^5 3.5. Absolut^ moments 70 3.6. Locally compact Abelian groups 73 3.7. Random variables 76 3.8. General positivity 80
Vlll CONTENTS Chapter 4. Laplace and Mellin Transforms page 4.1. Completely monotone functions in one variable 82 4.2. Completely monotone functions in several variables 86 4.3. Subordination of infinitely subdivisible processes 91 4.4. Subordination of Markoff processes 95 4.5. A theorem of Hardy, Littlewood and Paley 99 4.6. Functions of the Laplace operator 102 4.7. Multidimensional time variable 106 4.8. Riemann's functional equation for zeta functions 108 4.9. Summation formulas and Bessel functions in one and several variables 112 Chapter 5. Stochastic Processes and Characteristic functionals 5.L Directed sets of probability spaces 118 5.2. Markoff processes 122 5.3. Length of random paths in homogeneous spaces 127 5.4. Euclidean stochastic processes and their characteristic functionals 133 5.5. Random functions 137 5.6. Generating functionals 143 Chapter 6. Analysis of Stochastic Processes 6.1. Basic operations with characteristic functionals 145 6.2. Convergence in probability and integration 150 6.3. Convergence in norm 153 6.4. Expansion in series and integrals 156 6.5. Stationarity and orthogonality 160 6.6. Further statements 163 Notes and References 168 Indexes 173
CHAPTER 1 APPROXIMATIONS 1.1. Approximation of functions at points In ordinary Euclidean space Ek: (£1} ...,£&)> -oo<g,<oo, j=l,...,&, for any dimension &^1, the ordinary Lebesgue measure element dgx...dgfc will usually be denoted by dvg. A function/(gl9...,§7c) will also be written briefly asf(£) or/(£$), and we will also put m=(£i+...+8)*. We take a family of functions {**&,...,&)}, (i.i.i) also called 'kernels', subject to the following assumptions. The index B ranges over 0<R<co and has continuous and occasionally only integer values. For each R, KR(^) is defined and Lebesgue integrable over Ekt so that the integrals J Ek J Eh exist, and we have Kr(£j) &vg — 1 (!•!•-) for alii?, and f | X^g,) | etof^Z0 (1.1.8) J Eh with jST0 independent of J?; and, what is decisive, for each 8>Q, no matter how small, we have Jim f !**(&) | <fog=0. (1.1.4) We note that for KEfo) £ 0; (1.1.2) implies (1.1.3) with Z0= 1. Starting from an integrable function K(£l9 ...,£*;) with L if* if we put Z^g1,...,&=2t*Z(2&,...)2gl), (1.1.5)
2 APPROXIMATIONS then this is a family as just described, since by the change of variables B£j-^£j, J = 1,..., Jc, we obtain f KB(^)dv^[ Z(&)d»£=l, J Eh J Bjc J Ek J Ek J I £15* JlflSltt and for fixed £> 0 the point set {| £ | 2> .&£} converges to the empty set as B-^co. Sometimes a statement will be intended only for such a special family of kernels, as will be indicated by the context. For a measurable function/(#) =f(#1} ...,a?fc) in Ek we introduce, if definable, the approximating functions *a(«)= /(^i-^.'-.^-Q^^.-.,^)^; (1.1.6) J Ek and since (1.1.2) implies f(x)=\ f(x)KE(gl9...,£k)dvi$ J Ek we obtain sB{x) ~-f(x) = (f(x - g) -/(a;)) JTB(£) dtof, and our first statement is as follows: Theorem 1.1.1. Iff(x) is bounded in Ek \m\^M9 (1.1.7) then sRix)^~f(x)> i^-^oo at every point x at which f(x) is continuous Also, iff(x) is continuous in an open set A, then the convergence is uniform in every compact subset A0. Proof. We have I sR(x)-f(x) 1 £ f \f(x-i)-f(x) I. I KB(i) | dve JJB* =Ii{R,x)+Iz{R,x). -f +f £1^* Now, 11^.8,^) | £ sup |/(a:—g)—/(a?) | . Z0, I.SI<*
APPROXIMATIONS 3 and by continuity of f(x) at x this is small for 8 sufficiently small. However, for S fixed (small) we have | I2(R3 £) | £ f (|/(*-£) | + |/(a:) |). | KJ£) I <% J\z\>* and by (1.1.7) this is g 2M \ EB(g) \ dve, which is small for large R by exphcit assumption (1.1.4), q.e.d. The global requirement (1.1.7) was only needed for obtaining lim f |/(a-£)M*a(g)|<tof=0, (1.1.8) and it can be relaxed if we correspondingly tighten the assumptions on KR(£). For any measurable set A in Ek we can introduce the L^(A)-norm, #^1, ll/ll* = sup(f \f{x-!i)\HvXl\ (1.L9) and for A=- Eh this simply is vl/0 {Sjm]Pdv$ Also, if-4 is the set -V-i£6<i i«i,...,*, (1.1.10) and if/(Si,...,£&) is (multi)-periodic with period 1 in each variable, then the LP(Tfc)-norm is the Z^-norm of f(x) over a fundamental domain of periodicity. Now, it follows from the Holder inequality that if, for a given KR(E,), (1.1.8) holds for every function for which sup f !/(*-£) | tto{<oo, (1.1.11) then it also holds for every function with finite Lv(Ek) or Z^(Tfc)~norm. Now, (1.1.11) means that \f(x) | becomes bounded after having been averaged over a ^-neighbourhood of each point, and for such an f(x), the integral . /(&)*&)*>
4 APPROXIMATIONS is definable, whenever we have 2 sup I K(£1+mv ..., £,+mfc) | <oo, (1.1.12) the summation extending over all lattice points (ml3..., mk) ,m^Q. Next, (1.1 J 2) holds in particular if we have for all i;eEk I *<&■-..& I s-qjjs* d.1.18) for some 0 > 0, no matter how large, and some p > 0, no matter how small. Also, if we form the special family (1.1.5), then the estimate I K*&) I ^i+jR&+p|g|*+p implies |ia(g)|^_^- for | £ | ;> £, E> 1, and this secures relation (1.1.8) under the assumption (1.1.13). We do not claim that the mere condition (1.1.12) would secure (1.1.8), but it could be shown that the condition S2*» sup \K(il9...9^\<^9 (1.1,14) which falls between (1.1.12) and (1.1.13) would already suffice, and hence the following theorem: Theorem 1.1.2. For a family of the form (1.1.5), if the kernel K(g) satisfies (1.1.13), or only (1.1.14), then theorem 1.1.1 also applies if, globally, f(x) has a finite norm Lp(Ek) or only Lv(Tk). Ifweput (I for £inTfc, J-&,.-.,&)-{0 for £notinrfc> d-1-16) 1 fft P then ^(^jyu *'* ^a?i + ^-'a;*+^efof' (1.1.16) and in particular for the simple exponential X(oc;x)~e-w>*K {a,x)&alzl + ... + *hzk (1.1.17) v Jt sin7rha,- , , ,^ it is xteaO.n-jj^. (1.1.18) The kernel (1.1.15) is a product kernel, in the sense that we have mi U=K°(Q...K<>(£k), where iT°(§) is a kernel in Ev
APPBOXIMATIONS 5 Another product kernel is the (nonperiodic) Fejer kernel ft (*£)•. ,1.1.19) which, however, although it satisfies (1.1,12), does not satisfy (1.1.14) and thus could not be used in theorem 2. With a kernel K°(£) we may also form the multi-index kernel and most statements would be valid if Bli...iBk tend to oo independently of each other, but we will not pursue this possibility. Of paramount importance is the Gaussian kernel e-^*--+fi»se-iriei,> (1.1.20) which in addition to being a product kernel is also, antithetically, a radial function, meaning that there is a function H(u) in 0 ^u < oo, such that Kfa)=H{\£\). (1.1.21) We will take as known the formula f e-7r^-27iiax^a==e-7rx^ (1.1.22) which implies e_„(lM=RJ e_ffE^|2+2„-(a>a^ (LL23) and this time we obtain for the function (1.1.17) the approximation *.&(<*; a) =X(<*'> x) <r*(IalW). (1.1.24) For any radial kernel (1.1.21) it is profitable to introduce in **(*)« f /(*+g)#(*|£|)2**efof J Eh polar coordinates g,=%, 0^t=|g|<co, 7|+...+7}=l, in which case the volume element dvg is the product of t^"1 dt with the volume element do)^ on ^ 2 , , 0 _ ,, _ _„. * #*-i^i + ---+7l=l, (1.1.26) 2^**> the total volume of /S^-a being ^fe,x= . We then obtain 1 (f k) sR{x) = o)kT fx(t)H(Rt)RH^dt J*=o = w*_iJ^/» (|) HW **-1^ <LL26)
6 APPBOXIMATIONS where /«,(*) s ffa+tyv ...,£&+%)<K (1.1.27) is the spherical average of our function at distance t from the given point x. By Fubini's theorem, /^(t) exists for almost all t, and we are always permitted to put fx(Q) =/(%), and a glance at the term ii(.R; x) in the proof to theorem 1.1.1 leads to the following conclusion: Theobem 1.1.3. In theorems 1.1.1 and 1.1.2, if K(^) is a radial function, then locally it suffices to assume that the spherical average fx(t) ->/a;(0) as t->0, which is a weaker assumption than continuity proper. For h=2 we have ^x = 27r and 1 f2 /«(*)« ^ f{xx + t cos 0, z2+t sin 0) dd. However, for &= 1 we have g>0=2 and/a;(t) = |[/(a:-ht)+/(a; — £)], and radiality means evenness, K{ — £)=JT(£). For & = 1, a function is even if it is invariant with respect to the (then only nontrivial) orthogonal transformation g' = — £ which leaves the origin fixed. Now, for k g: 2, radiality means invariance with respect to the entire group of such orthogonal transformations and/^t) was an average over this group. However, if JBT(^) is invariant with respect to a subgroup only, then the function /(#+£) may be averaged correspondingly. Thus if Jl(£1} ..., £fc) is even with respect to each £,5 separately, then it suffices to assume in theorems 1.1.1 and 1.1.2 that the averaged function 2jtXf(x1±£1,...,xk±£k) shall be continuous at £ = 0. Turning for a moment to the smoothing operation (1.1.16) we note that by iterating it (or by some such procedure) we obtain the following result: Lemma 1.1.1. A continuous function f(x^) in Ek which is 0 outside a compact set is a limit, uniformly in Eh, of suchlike functions each of which is of class C^ (that is, has continuous partial derivatives of order g r) for any fixed r. The conclusion also holds more precisely in the class O00, but of this we will not make use in primary contexts.
APPBOXIMATIONS 7 1.2. Translation functions In Ek we take a family J5" of functions {/(#)} with the following properties: (i) it is a group of addition, meaning that if/, g€#" then /—g€j^; (ii) it is invariant with regard to translations, that is, if f(x) € ^, and if for any u = (ul3..., uk) in Eki we define fu(x) Es/fo+mx, ..., xk+uk), then /"(a;) € J5"; and (iii) it is endowed with a norm ||/1| such that 110H = 0, 0£||/||<oo, (1.2.1) 11-/11=11/11, ll/+0ll^ll/ll+ll<7ll, (L2.2) and, what is important, this norm is invariant, that is, 11/" 11=11/11- d-2.3) With any/eJ^ we associate a certain non-negative function in Ek:(%,..., uk), namely, the function r,(w) = ||/«-/||, (1.2.4) and we call it the translation function of /. It has the following properties. First, tM = q (12 5) by (1.2.1). Next, due to \\f"-f\\£ Uu 11 +11-/11 = II/H+ ll/H=211/11=-^. we have 0^rf(u) <M=21|/||. (1.2.6) Next, we have ||/ll+.-/. ||= „ (ftt_f)v ||= ||/«_/||, the last by (1.2.3), and on putting v = — w we obtain rf(~u)~Tf(u). (1.2.7) Next, we have \\f+* -/1| g ||/w+» -/* || + ||/"-f || and hence Tf(u+v)^Tf(u)+Tf(v), (1.2.8) and finally for /, g € IF we have ll/w+^-/-^ll^ll/w-/ll + llsrw^ll and hence Tf+g(u)=Tf(u)+Ta(u)* (1.2.9) Now, (1.2.8) implies Tf(u+v)-rf(u)£rf(v)9 (1.2.10) but we also have
APPROXIMATIONS and if herein we replace —ubyu+vwe obtain -t^St^+d)-^). (1.2.11) Combining (1.2.10) and (1.2.11) and also using (1.2.5) we obtain | Tf(U +1>) - Tf(u) | ^ Tf(v) = Tf(v) - Tf(0) and hence the following conclusion: Lemma 1.2.1. If a translation function is continuous at the origin it is uniformly continuous throughout. Next, by the use of (1.2.6) and (1.2.8) we now obtain by a familiar reasoning the following conclusion: Lemma 1,2.2. If<F0 is a dense (in norm) subset of IF and ifrf(u) is continuous in u for fin !F§ it is continuous for f in IF. But if W is a normed vector space, more can be stated. Lemma 1.2.3. If ^ is a normed vector space and ifrf(u) is continuous in u for a set JFQ whose linear combinations are dense in IF, then it is continuous for all of IF. This follows from ^Vi+.-.+W^) S\cx\ rfl{u) + - + K | Tfn(u)- Now, for all finite multi-intervals Ia'.ajgXjKbj, j=l,...,& (1.2.12) we introduce the ' characteristic functionsJ , . (1 for xmlabi »a*(v)**L r , T L2-13) (0 ior x not in Iab, and it is a basic fact of the Lebesgue theory that their linear combinations are for every 1 <p < oo dense in the Z^iSy-space with the norm / /• \i/p (Jj/B !•**)' On the other hand, we have {u)=(L,' wo>(^+m) ~a*®|s>^)1 and it is easy to verify that this tends to 0 as | u | -> oo. Similarly, if we introduce for the periodic functions f(x± +m1}...,xk+mk) =f(xl3...,xk)
APPROXIMATIONS 9 the Zr35(TA.)-norm (L m |6^)1/3,=(Ll/(*+g) '^F' then linear combinations of periodic functions of the form (1.2.13) are again dense in norm. Hence the following conclusion: Theorem 1.2.1. For functions in L^(E^) and (periodic) functions f in Ly(Tk), l^p<co, the translation functions are (bounded) and continuous. We note that the general norm as defined by formula (1.L9) is invariant with respect to translations, but we do not at all claim that every function with a finite norm of this kind has a continuous rf(u). However, if we take any set of functions {#"'} each of which is bounded and uniformly continuous, and then form their smallest Banach closure with respect to the norm for a set A of finite Lebesgue measure, then their translation functions are continuous. If we choose for {^"'} the simple exponentials and for A the set Tfc, then the smallest closure is composed of the almost periodic functions of the Stepanoff class Lp, to which we will sometimes refer incidentally. 1.3. Approximation in norm We will now state a certain proposition first in a general version heuristically and then in a specific version precisely. Theorem 1.3.1 (heuristic). If Tf(u) is continuous then the approximating function r Sr(x)=\ f(x-e)Zs®dve (1.3.1) J Eh converges to f(x) in norm: ||sR —/1| ->-0 as R->oo. Reasoning. For a finite discrete sum we have II Sym(/(*+gJ-/(*)) U 2 |y» I • 11/^-/11 m m = S \7m | Tf(£tn), m and this suggests for (1.3.1) the estimate II«*-f II = 11 I" (/(*-£)-/(*)) **(£)dvi11 IIJ Eh II <;f ||/f-/||.|^(l)|^=f rM)-\KR(i)\dvt J Eie J Eh
10 APPROXIMATIONS But the last term is the value of {Tf{x-i)-rf{x)).\KB{i)\dvi J E 1 Ek for x = 0, and for bounded continuous 77(g) this tends to 0 as R ~» 00 as in theorem 1.1.1. Assume now specifically that f{x) is in L-^E^). The function #(*,g) =■/(*-!) *a(£) is measurable in (x, E), and we have f (f |/(*-0|.|tfa(£)|tfoeW J ifo \J ■#& / = (* (|JTa(£)|.f |/(*-£)|<WU£ J Ek \ J Ek I I ^©|^.||/||= ||/||.Z0, JjB and since the last term is finite, it follows by Fubini's theorem that the integral (1.3.1) exists for almost all x and is an integrable function in x. This being so, we now obtain f \*R(*)-M\dvm£{ (f \f{x-£)-f{x)\.KE{Z)\dv^\dvx J Ek J Ek \J Ek J =£ (J l/(*-SW(*)l*.)-l*a(£)l*e = f TM)\K*(£)\dv£, J Ek and this time rigorously. This argument also works for (periodic) feLx(Tk), if we replace one of the two symbols Ek by Tki and thus we obtain the following theorem, at first for p = 1: Theorem 1.3.2. Iff(x) belongs to Lp(Ek) or to periodic or Stepanoff almost periodic Lv(Tk)9 then the integral (1.3.1) exists for almost all x, is a function of the same class with ll«B IIS 11/11. *0 and we have lim ||^—/|| = 0. (1.3.2) R-*-co For p > 1 it is necessary to apply the Holder-Minkowski inequality {\a{\bH{X' £) ^P dV^lP ~LilAH{X' ®****)"**"*
APPROXIMATIONS II for H(x,g) being first |/(#—£) J. | KR{^) | and then \f(*-$-f{z)\.\KR<&\ and B=Ek and A =j£fc or 2V 1.4. Vector-valued functions Theorem 1.1 on convergence at a point and theorem 1.3.2 on convergence in strong average can be brought together by a third theorem embracing them both. We define in Ek a function f{xs) whose values / are not ordinary complex numbers but more generally elements of a Banach space 5 the norm of which will be denoted by || ||. For f(x) we employ the concept of (strong) measurability and (strong) integrability as introduced by this author (the so-called Bochner integral), and if f(x) is bounded in norm *, ... _- _ \\J{x)\\^M, xeEk, (1.4.1) then for numerical KR(^) in L±(Ek) there exist the approximating functions . **(*) = f{x~£)KR{mn (L4.2) as functions again with values in B. We have Pa(*)-/(*)ll = f ll/fc-fl-ftO II-!**(£) |«k{ and thus continuity in norm at a point x, lim ||/(#-£)-/(z)||=0 (1,4.3) implies convergence in norm lim m*)-fa) 11=0, (1.4.4) B->co and hence the following conclusion: Theorem 1.4.1. Theorems 1.1.1, 1.1.2 and 1.1.3 also apply to functions f(x) with values in a Banach space B. We now take mEk: (yv..., yk) a family W as in section 1.2, assuming that it is a Banach space and an element f(y) in & for which the translation function Tf(x) is continuous. If now we denote hjf(x) what in 1-2 we denoted by /*, then this /(#) is bounded and continuous in
12 APPROXIMATIONS norm and thus falls under theorem 1.4.1. In a certain formal sense we can write in (1.4.2) where sR(y+x) is sR(x), and if our Banach norm is LJJE^ or the periodic or almost periodic £p(jTfc) then this is rigorously so, for almost all y> for any given fixed x, as can be realized by assuming first that f(y) is a finitely valued function as in section 1.2 and then passing to a limit in norm. But if this is so then relation (1.4.4), if applied at x = 0, is simply \\sR(y)—f(y) ||->0 in the sense of theorem 1.3.2, and thus theorem 1.3.2 appears likewise subsumed under theorem 1.4.1. Now, the (heuristic) theorem 1.3.1 has also a (heuristic) converse to the effect that if sR(x) converges in norm to f(x) then rf(u) is continuous. But in the specific version in which we will establish this rigorously, we will not start out from a point function at all but from a (more general) set function, prove for it a ' weak' approximation by sR(x)9 and then show that if this approximation is also a strong one then the set function is the indefinite integral of a point function, and Tf(u) is continuous. 1,5. Additive set functions We denote by V{Ek) the vector space of set functions F(A)=F1(A) + iF2(A) which are defined and <r-additive on the a'-field of (ordinary) Borel sets in Ek, the norm being the supremum ||F|| = SupS,|F(A,)| < - Uv) for all partitions into disjoint sets. If F(A) is real and 2^0 then \\F II—jF(jB&), and the subset of such elements in V will be denoted by F+. Any F in V can be written as F(A)=Fl(A)-F,(A)+iF3(A)^iF,(A), (1.5.1) where F1,F2,FZ,F^V+, and there exists a GeV+ such that |F(A) \&Q(A) for aU A. Also, ||F\\£ \\G||-G(Ek). Now, among all such G(A) there exists a 'smallest' one, which we will sometimes denote by P(A) and call the 'absolute value' of F(A), and it is characterized by j^,,^,, „,„_„*„. (L5jJ
APPROXIMATIONS 13 We will say that F is zero on a set B if F(B) — 0, and this is equivalent to stating that F(B0) = 0 for any subset B0 of B. Also if we are given a cr-additive set function F(A) only for the subsets of a set A0, then there is another set function in V which coincides with F for A C A0, and is zero on Eh—A°. If/(*)el^)then F(A)=j/{x)dVx (L5.3) defines an element of F, the function f(x) being defined up to sets of measure zero. We will call F(A) an (indefinite) integral of f(x) and f(x) a 'derivative' of F(A). If F is the integral of/(#) then F is the integral of \f(x) |. The elements in V which are integrals are a closed subset of V and constitute by themselves a Banach space which we will denote by A C(Ek). the letters A 0 standing for' absolutely continuous', and the meaning of this is that F in V belongs to AC if and only if F(A0) = 0 whenever v(A0) = 0. Any Fe V+ defines a bounded Lebesgue measure on the Borel sets of Ek9 and every bounded Baire function b(Xj) is integrable, and the integral will be denoted somewhat ambiguously by f bfrddFfa), (1.5.4) JEk the employment of the symbol F(Xj) instead of F(A) being somewhat arbitrary on the whole. For any FeV we take any decomposition (1.5.1) and define (1.5.4) by fbdF!- [bdFz + ihdFz-i[bdF^ (1.5.5) and thisintegralis uniquely defined andhas many customary properties. If FeAC, then (1.5.4) has the same value as the ordinary integral b(xi)f{xj) dvx, and the familiar estimates JJSjc \jbfdvx\sj\b\.\f\dvx<suy\b(x)\j\f(x)\dvx can be generalized to I fbdFUf|b|.|dF|^sup|&^)|f|^F|, where | dF(x) \ =dff(x), with P{A) being the absolute value of F(A), and we also have f \dF\= f dff=P(E!c)=\\F\\. J Eh J Ek
H APPROXIMATIONS The following properties are of considerable importance: Lemma 1.5.1. If F19 F2e 7, then Fi=F2. that is F1(A)=F2(A) for all A, whenever * * c(x) dFx(x) = c(x) dF2(x) (1.5.6) J Eh J Ei for every continuous c(x) which vanishes outside a compact set, and (by lemma 1.1,1) it suffices to assume that c(x) belongs to class C^\ for some fixed r. Furthermore, if we are given F in V and a measurable function f0(x) which is integrable (with regard to ordinary measure) over any compact set and if we have » » c(x)dF(x)=\ c(x)f0(x)dx J Eh J Eh for all such c(x), then FeAG andfQ(x) is its derivative. Next, if Fv F2 in 7 have the same value in all' octants' Ia:-cQ<Xj<aj, j=l,...,k, (1.5.7) then Fx-sFg. On denoting F(Ia) by F(av ...,ak) we obtain a certain point function F(x$) which is representative of F(A), and it is this interpretation of the symbol F(x§) which may be read into the formula (1.5.4). We will simply put F(xs) € 7, and we note, what is much used in the theory of probability, that a function F(x$) is in 7+ if and only if F(x1,...9xk)^F(y1,...,yk) for xx<>yXi...,xkSyk, (1.5.8) and also £ (-l)t^'+t^F(yj+tj(x3—yj))^0) F(xx-0,...,xk-0)^F(xx,...,xk), (1.5.9) ]imF(xv...,xk) = 0 if a?,--> —oo (1.5.10) for a single index j, and F( + co,..., +oo)s ||.F || <oo. (1.5.11) We will now take two elements F, G in V and make statements about them which we will explicitly discuss only if they are both in 7+, relying on a decomposition (1.5.1) for both of them if they are not. If we interpret F(A) as a measure in space Ek, and 0(B) as a measure in a space E\> then on the Borel sets of the product space E^ElxEl (1.5.12) there is a product measure /i(C) such, that fi(AxB)=F(A).G{B). (1.5.13)
APPROXIMATIONS 15 The functions ^ = £5+%.J==l.....& are Baire functions in E2k} and thus with any octant (1.5.7) in Ek there is associable the Baire set Ca:-co<^+Vt<^ i = l,...,* (1.5.14) in E2k. 1^now we Pu^ H(av...,ak) =/i(Ca), then this defines a point function in H(Xj) in V as previously described; and thus there is an element H(G) in V such that for every bounded Baire function b(x^ in Ek we have f b{z)d9H{x)=\ f Hi+V)diF(S)dvG(7j)9 (1.5.15) J Ek J EkJ Ek the second integral being taken as a double integral or repeated integral, indifferently. Hence the following statement: Theorem 1.5.1. With any two elements F, G in V(Ek) there is associable a third, H=F*G (their convolution) such that (1.5.15) holds The properties F*G=G*F and (F1^F2)*F3=Fi^(F2^F3) are obvious, but others will be stated formally although taken as known. Theorem 1.5.2. For H=F*Gwe have H{A)=[ F(A-V)dvG(V) (1.5.16) J Ek for every set A. IfF has a derivative f then H has a derivative h, and iff is a bounded Baire function then » h(x)=\ f(x-V)dnG{V). J Ek If all three functions have derivatives, then h(x)=\ f(z-i))g{7i)di] J Ek for almost all x. Defection 1.5.1. A sequence of elements {Fn} in V will be called Bernoulli convergent if their norms are jointly bounded and if there is an element F in V such that lim f c(x)dFn{x)^\ c(x)dF(x) (1.5.17) n-*<x> J Ek J Ek for every bounded continuous c(x).
i6 APPROXIMATIONS It will be called weakly convergent if (1.5.17) holds for every continuous c(x) which vanishes outside a compact set, and (by lemma 1.1.1) it may be even assumed that c(x) belongs to C(r) for a fixed r. Certain decisive properties will be taken as known, and they are being stated as they will be needed. Lemma 1.5.2. (i) Any infinite sequence {Fn} in V(Ek) with || Fn\\<>M contains an infinite subsequence which is weakly convergent, and if the entire sequence is not weakly convergent then it contains two weakly convergent subsequences whose limits are not identically equal. (ii) For Fn € V+(Ek), if {Fn} is weakly convergent to F, then lim Fn(Ek)>F(Ek), (1.5.18) and equality holds if and only if the weak convergence is Bernoulli convergence. (iii) If {Fn} is weakly convergent and ifallFn are zero on an open set A0 then the limit is also zero on A0. (iv) If{Fn} is Bernoulli convergent and if a function X(oc; x) is bounded and continuous for {ocl9..., cck) in Ek and (xx,..., xk) in Ek then the limit relations » * lim X(a;z)dFn(x)=\ X(*;x)dF(z) (1.5.19) rc-->oo J Ex J Ek holds uniformly in every compact set \ a | ^ ocQ, 0 < oc0 < oo. (v) IfFn-+F weakly, and \Fn\^M, then (1.5.17) holds for every continuous c(x)' which is zero at infinity \ that is, for which lim c(x) = 0. |a |->oo All this will be needed later on, and for the present we are stating a theorem on approximation. Theorem 1.5.3. (i) IfFe V(Ek), and {KB{£)} is as before, then ^(A)=f F(A-^)KEa)dvi (1.5.20) is Bernoulli convergent to F(A), as JR~>co. (ii) 8R(A) is absolutely continuous, and if KR(£) is a continuous function then its derivative is **(*)=[ K1Az-g)deF®m f K^d^x-i). (1.5.21) J Ek J Ek (iii) IfSR(A) converges in norm to F(A), lim ||SM-F || = 0, (1.5.22) then F is absolutely continuous.
APPROXIMATIONS 17 (iv) Finally, if K(^) is as in theorem 1.1.2, then the convergence of $R(x) at a point xisa local property, meaning that ifF is zero in a neighborhood of x then sR(x) -> 0 at the point. Proof. We have Z, I SR(A,) I £ f £„ I F{Ap-g) I. I KM) I d*e J Eh gf llPll.l^gJittoj, and thus l|SJ,||g||J,||.Z'0l (1.5.23) and for a bounded Bake function b{x) we have f 6(*)d(B/Ss(»)=r (f &(*)«*-£))£*(£)<% (1-5.24) J JET* J Ek\J Ek J by Fubini's theorem. Now, the inside integral can also be written as b(x + £)dxF(x)^M), J E I Ek and if b(x) is continuous then /?(£) is (bounded) and continuous, and /?(£) KR{i) dvg therefore converges to /?(0) s b(#) dxF(x), which J $* J #*; proves part (i) of the theorem. Part (ii) is taken directly from theorem 1.5.2, and part (iii) follows from the fact that AC is a closed subset of V; and part (iv) states that under the assumptions of theorem 1.1.2 we have . lim |i^(£)|.|<%F(z-£)|=0 for F(#) € V{Eh), which can be easily verified. Next, for F in V we introduce the translated element Fu(A)~F(A + u), and the translation function r*(w)=||Fw--F||, which, if F(A) has a derivative f(x) is our previous rf(u). Theorem 1.5.4. For F in V(Ek), ifTF(u) is continuous, then FeAC. Proof. We have S,| 8n{Av)-F{Av) IS f 2,| F{Av-i)-F{Av) \. \ KR{£) \ dvs, J Eh
i8 APPROXIMATIONS and hence \\SS-F\\ ^ I t*{1-).\Kr{1-) \dv£, and for tf{E) continuous this tends to 0 as E->co. Now apply part (iii) of theorem 1.5.3. Theorem 1.5.4 also holds for a periodic function Fe V(Tk) which we will introduce next, but we want to point out that it also holds on compact groups 67, as we have shown elsewhere; that is, if we take a <7-additive set function F(A) on the Borel sets in G and introduce the right translation function rF(u) = || F{A) -F(uA) || say, then F(A) is absolutely continuous with regard to the Haar measure v(A) if (and only if) tf(u) is continuous in u. An extension to rioncommutative locally compact groups would be of some interest. Furthermore, if a function/(#) in ( — oo, oo) is integrable over every finite interval and if (I (*T+a r-T+a \ r\ + \m\dx)=o, 1 J T J -T I —oo < a < oo, then the function — l rT t(u) = lim - | f(x + u) -/(a) | dx, if finite, has all the properties of a translation function, and it would be of some interest to study the implications of the assumption that r(u) is continuous without being identically zero, and the study ought to be extended to more general means of the form ^ -L\ !/(*+«)-/(*) I **. say,#>0. 1.6. Periodic additive set functions Periodic set functions F(A), like all other periodic functions, are to be thought of as being first introduced on the multitorus Tk: — \ ^ Xj < | viewed as a compact space, and they may then be transplanted onto the entire Ek, as covering space of Tk, by periodic repetition F(A+m)=F(A), m=K,...,mfc), so that they may be used in the formulas of theorem 1.5.3 which, as in the case of point functions, we wish to retain as they are, with integrations in them extending over the entire Ek.
APPROXIMATIONS If now we introduce the symbol V(Tk) to designate the periodic cr-additive set functions (we will also employ the symbols AC(Tk), V+(Tk), etc), then we can first of all state as follows: Theorem 1.6.1. Theorem 1.5.3 remains literally in force for F e V(Tk), provided Bernoulli convergence is now defined by lim c(x)dFn(x)^\ c(x)dF(x) (1.61) n->co J Tk J Tk for every continuous periodic function c(x). On the torus, due to its compactness, there is no difference between weak and Bernoulli convergence, and for instance any sequence in V+(Tk) for which Fn(Tk) ^ M contains a subsequence for which (1.6.1) holds. Thus, in this respect, the study of joint distribution functions of random positions on a closed wire is somewhat less sophisticated than for positions on the open infinite wire, on which the theory of statistics operates traditionally. If/(#) is periodic then for the integral **(*)= f Kx-QK^dv^ (1.6.2) J Eh we can write formally 2 f /(* - g) **(£) dvs=2 f /(* - i) KR(i+m) d»e, (m)J Tk+(m) (m)J Tk and thus we have sR(x) = f(#—£) S.R(E) dvg, (1.6.3) J Tk where we have put J?it(.;) = S K^+m). (1.6.4) On) Now, by Lebesgue theory we have 2 f \KR(£+m)\dv^( S|^(g+m)|d^=f \KR(£)\dvi} (m) J Tk J Tk (m) J Mk and since, by assumption on KB(^), the last number is finite, the entire reasoning is rigorous in the following sense. The sum (1.6.4) is absolutely majorizedly convergent at almost all points x in Tki and, as can be easily seen, in every compact subset of Ek, and the resulting sum function is independent of the order of the terms. Therefore %r(£j) is a periodic element of Lx(Tk) as seen from (m) (w)—(p) 2-2
20 APPROXIMATIONS and it has the following properties: (i^(£)|^£o> (1-6-5) /. Tk f KR£)dvz=l, (1.6.6) lim [ &R(g)dvi=0. (1.6.7) Furthermore, for feL!P(TJC)) (1.6.2) has indeed the value (1.6,3) a.e., but if we start from some kernels RE(£) on Tk with the properties stated, and if we introduce the approximating sums (1.6.3) and for FeF+the sums . SR(A)~\ F(A~£)KRtt)dvE, (1.6.8) then most of the previous theorems can be established likewise. Theobem 1.6.2. For periodic point and set functions, theorems 1.1.1, 1.3.2 and 1.5.3 can also be established for the partial sums (1.6.3) and (1.6.8). As a rule we will represent periodic functions by the previous integrals over Ek9 which are formally the same as for nonperiodic functions, but at one stage theorem 1.6.2 will be made use of, and for this utilization of it we are going to supplement it by a lemma in which jR will have integer values n only. Lemma 1.6.1. The (periodic) Fejer kernel ft /ex- A (sin^TT^)2 ^^^/A^sin^-)2 Wj= —n 5—1 \ n J = SA»e2*^, (1.6.9) (m) where A» = n l-1—— I if |% | <£n,..., \mk \ £n, and An(m) = 0 if for some j we have | % | > n, has the properties (1.6.5), (1.6.6), (1.6.7) needed. Proof. Since jS^g) ;> 0, (1.6.5) is implied by (1.6.6). Now we have JU&)«#«&)...#»(&), where (sinnflf)2 /ift1A\ H-(0-J^gji' (L6-10)
APPBOXIMATIONS 21 and for 8< | £ | ^ \ we have sothatajfbrtion lim .ffn(g)eZ£=0. (1.6.12) We also have j Hn(£) dg = 1, (1.6.13) and if in (1.6.7) we replace the exterior of the sphere | g | <S by the exterior of the cube \£j\<S,j — l9...,k9 as we may, then it suffices to estimate a certain number of terms of the form f P f ffn(fl) #»(&) • • • Hn{i-k) d£± ... d£7c. Now, by (1.6.13) this reduces to Hn(£i) d£v an(i this tends indeed to 0 as n->oo by (1.6.12). It should be noted that the sharper estimate (1.6.12) does not imply the same estimate for the product kernel (1.6.9), and the latter would again not be eligible for application in a theorem analogous to (1.1.2) on Tk instead of on Ek. Finally, we ought to mention that the periodic Fejer kernel (1.6.9) is linked to the nonperiodic Fejer kernel (1.1.19) by relation (1.6.4) as could be deduced from theorem 2.4.1.
22 CHAPTER 2 FOURIER EXPANSIONS 2.1. Fourier integrals For the present we will introduce Fourier integrals only for functions Lx{Ek) and more generally V(Ek), but not for Lj,(Ek)9 p>l, except that we will summarize statements on Plancherel transforms for functions L2(Ek) whose theory we will take as known. However, when introducing Fourier series, we will envisage Lp-classes in general because it takes very little effort to do so. If Fe V(Ek), we can introduce the Fourier transform ${a)s<f>F(as)=\ e-^^dFix), (2.1.1) J Eh and it is trivial that it is a bounded function, H{*)U\\F\\, (2.1.2) and it is also very easy to see that it is uniformly continuous for (a5) in Ek. If F has a derivative / we will also write ^(ct) = ^/(ct.)==r e-*"*«>*)f(x)dvx. (2.1.3) J Ek Theorem 2.1.1. IfF9 67e V(Eh) and H=F*G, then ^K-) = ^K).^(%), (2.1.4) that is, the transform of a convolution is the product of the transforms. In particular, iff, geLv then &(*) = &(«) 0,(a), (2.1.5) where h(x)=\ f{x~y)g{y)dvyi (2.1.6) J Ek the integral existing almost everywhere. Proof. Put b(x)=e-2ni^^ in formula (1.5.15). Theorem 2.1.2. IfF,Qe V, and 0(a,)= (V2™^>dF(£), f(Q= te*««£>*>dG(*), (2.1.7) then \<j>((x)eW>*»dG(a)« fVfc-f)dF(£). (2.1.8)
FOURIER EXPANSIONS 23 Proof. Since jje**«*>-£>dG{*)dF(g) U jjl«?(«) | -1 dF® | = || 01|. ||F || <oo, it follows by Fubini's theorem that the integral eW*.»-&dG(<x,)dF(!z) !!• exists, and relation (2.1.8) expresses the equality of the two repeated integrals by which it can be evaluated. In the next theorem a function 8(0.3) will be called a convergence factor if . \$fa)\dva<co, (2.1.9) J Ek and if for its (anti) -transform #(£.)-= f e^^a)d(a)dva J Ek we have | | £(£,) | <fof < oo, \ K^dv^l. (2.1.10) J E J Ek Now, the corresponding transform of S(oc^B) is BkK(Bi;1>...,Bi;k), and if in theorem 2.L2 we put for 67(A) the indefinite integral of Sfa), then the theorems of chapter 1 imply as follows: Theorem 2.1.3. The integrals (2.1.3) and (2.LI) can be inverted to f(x)~\ e27ri^^^>f(oc)dva (2.1.11) JEh and F(A)-f dJ"T e**^>a>0*(a)dfoj (2.1.12) in the following sense: (i) For any convergence factor 8(a) the approximating function sR(x)=( $(^eW*>*)<f>f(oc)dva (2.1.13) converges in norm tof(x), Km f \f(x)-sR(x)\dvx-0, (2.1.14) and for the approximating function **(£)=[ *(^)ea^*^(a)efoa (2.1.15)
24 EOURIER EXPANSIONS the indefinite integral SR(A)= sR(x)dvx is Bernoulli convergent J a to F(A). (ii) For f(x) bounded, the convergence of sR(x) at a point x is a local property, and for anyfeLv or even FeV, it is so if \K(g)\£C(l + (£)"rp)-1l and in either case, sR(x) ->f(x) iff(%) is continuous at x, and for a radial K(E) only the spherical average off(^) around x need be continuous at x. For us the dominant convergence factor will be e~nl a|2 (and not the Abelian factor eHal) and we are going to utilize the corresponding sums /• sR(x)= e-"«aVR2hW*'«tyF(oc)dva (2.1.16) JEk for several purposes. If Fls F2eV have equal transforms, then for F=F1 — F2we have <j>F(cx) — 0, and hence sR(x)=0. But the Bernoulli limit of a null function is a null function, and thus part (i) of theorem 2.1.3 implies the following uniqueness theorem: Theorem 2.1.4. If <j>Fi{cx) = (j)F^(x), then FxsF2. If <f>fl(cc)=s$f%{a) thenfx(x) =/2(^) a.e. Next, assume that for F in V it so happens that I \ </>*{(*) \dva<co. (2.1.17) JEk Since | e-*«a '2^2) e27**a> | ^ 1 and e-*<"a ' W ~> 1 as R -> oo, the functions (2.1.16) are then convergent, locally uniformly, towards the function f0(x) = j eW- a> <f>F{a) dvai (2.1.18) JEk and for any continuous c(x) which vanishes outside a compact set we have _ c(x)sR(x)dva->\ c(x)f0(x)dvx. J Elc J Ek However, again by theorem 2.1.3 we have c(x) sR(x) dva -* c(x) dF(x), J Ek J Ek and by the second half of lemma 1.5.1 we obtain the following conclusion: Theorem 2.1.5. If for F in V, the transform <j>F(<x) is in 1^(18 m), then
FOURIER EXPANSIONS 25 F(A) is the indefinite integral of the continuous function fQ(x) as given by the inversion formula (2.LI8). In particular, iff^Lx and ^f(oc)eLv thenf is a.e. equal to the continuous function f0(x)=\ e2^>«>^(a)dv (2.L19) J 32* As an incidental application of this theorem we note that a convergence factor 8(a) as previously introduced may be assumed to be continuous to start with, in which case it is given by 8(a) =f &-*«**>&K(&)dvs, J Ek so that in particular £(0) = L This being so, we will now give the actual formal definition of a convergence factor. Definition 2. LI. A convergence factor 8(a) as was used in theorem 2.1.3 is a function with the following properties: 8(a) is continuous and 8(0) = 1, and 8(cc)eLv and its Fourier transform is in Lx. 2.2. Positive transforms. Plancherel transforms Theorem 2.2.1. If for feL± we have fif(oc)^0, and \f(x) \ ^M for \x\Sto (that is, if f has a positive transform and is bounded in the neighborhood of the origin) then <pf^a)eLx andf(x) is equal to the function /0(s) (see 2.1.19). This theorem is obviously contained in the following one. Theorem 2.2.2. IfforFe V we have <f>F(a) ^ — #(a), where #(a) £> 0, X(oc) eLv and if we have kcJi dm) <M19 Q<t<ao, ' (2.2.1) or only LB* e~^^dF(i Jltfl-Si SM2i (2.2.2) then <pF(a)eLli and F(A) is the indefinite integral of the function f0(x) givenby (2.1.18). Proof. We note that due to ||F || <oo, (2.2.1) is satisfied automatically for J t J £ 1. Next, (2.2.2) is indeed at least as general as (2.2.1), this being an abelian theorem, since, if we put G(t) = dF(£), so that
26 EOUR.IER EXPANSIONS <?(t)^0(tfc),wehave W f e-"W^dF(£)=Bk P e~"RH*dG(t) =*R*e~*#G(l) + 2tt^+2f te~"RH20(t)kdt = 0(l)+0(l)fC0^+2e~^2i2tft+1dt-0(l) as i?-> oo. On the other hand, if it is known that F(A) ^ 0, then (2.2.2) is not more general than (2.2.1), this being the classical Tauberian theorem, and (2.2.1) is the 'natural'way of putting the condition then. Now, for the function (2.LI6) we have 8B(0)=R*{ e-^i2dF(£)=f +f , J Ek J |*|^1 Jl*l>l l r l but ^JRfce-7rJR2||F||->0,andtherefore(2.2.2)isequivalentwith |J|*l>i| \sR(0)\£Mz. Therefore 0 £ f (<f>(oc) + #(a)) 6-^>a iWdva=sR(0) + f Z(a) e-»«aI W> dva, and because of the assumptions on %(a) this is SMZ+M^=MS. Therefore, letting B~+oo, we obtain (j> (a) -f #(a) dva < oo, and hence ^(a) € 1^, and now apply theorem 2.1.5. Definition 2.2.1. (In the theory of probability) a (joint) distribution function F(A) or F(%) in Ek is an element FeV+ which is normalized by \\F\\=F(Ek) — li and its Fourier transform <f> (<x,3) is called a characteristic function. For such data we obtain the following special conclusion: Theoeem 2.2.3. If for h random variables Xv ...,Xfc the joint distribution function F(xj) has a non-negative characteristic function and if ltoip{|Z1| + ... + |Xj5|gr}<oo, ra*o r then F(A) is the integral of the function /to)=f e~w>*><f>(a)dvai this expression converging absolutely.
FOURIER EXPANSIONS 27 If 0(a) is the transform offiaL^Ej,), then 0(a) 0(a) is the transform ofg(#)= /(ic + g)/(g)d^(seetheorem2.1.1), If, by chance,/belongs J Ek also to L2(Ek), then by Schwarz's inequality we have I g(x) -g(y) \ S j |/(*+g) -/(y+© |: |/(g) [ A% ^ (J|/(a,+fl-/(jf+£) |2d^ (J|/(g) |«)* =T/(*-y).||/ll. where r/w) is the translation function in the L2-norm. Therefore, g(x) is continuous, and by theorem 2.2.1 we have !f(x+^)M)dvi=U(oc)W)e^i^Mv^ (2.2.3) If fl9 /2 are two such functions, then J/i(* + g)jWl)*= f &(*) 0^2^">d^ (2.2.4) and in particular A(g)^(|) (tog = &(a) 02(a) dva (2.2.5) J|/(g)|aetof-J|0(a)|8efoa. (2.2.6) Now, by taking limits in L2-norm the restriction that /, fl3 f2 shall be not only in L2 but also in Lx can be eliminated and the following statements can be obtained: Theorem 2.2.4. With eachf(x) in L2(Ek) there is associable another function 0(a) = 0/(a) likewise in L2(Ek), both determined a.e., such that (i) relations (2.2.4), (2.2.5) and (2.2.6) hold, (ii) each 0(a) in L2 is the transform of some fin L2, (iii) for any Borel set A the inversion formulas hold: « f / f \ I f(x)dvm=\ 0y(a)( e»^«)(toBW, (2.2.7) J A J Ek \J A J f f(x)({ e~W>*)dv^)dvx~f 0y(a)aX, (2.2.8) J Ek \J A I J A and (iv) iff(x) is also in Lx then 0y(a) is the ordinary transform. If it also happens that the integral g(z)=\ <pf(oc)e*"i(*>°*dva J Ek
28 FOURIER EXPANSIONS is boundedly convergent, say, and if we integrate both sides with respect to a set A and compare with (2.2.7), we obtain f(x)dva= g(x)dvx, J A J A so that/(a;) ~g(x) ax. 2,3. Fourier series We are now turning to periodic point and set functions. For any F € V(Tk) we are introducing the system of Fourier coefficients 0(m)s0*(m)=f e-w^adFig) J Th for all lattice points m=(mv ...,mk), and if F has a derivative f(x) then this is of course 0(m)==^(m)== f e-tottoafffidvg. (2.3.1) J Th As before, we now have | <pF(m) \S\\F\\ and for H=F*G the convolution rule <p3S(m) = <pF(m).<f>G(m), and as an analogue to theorem 2.1.2 we are making the following statement. Given FzV(Tk), if a function \[r{£) is given by an absolutely convergent series M) = S 7M e2**^, £ | y(m) \ < oo, (w) then (by a direct substitution of this expansion) we obtain f ^(cr-^dF^^SrW^^)^^^ (2.3.2) J Th (m) If, in particular, we put for ijf(x) the periodic Fejer kernel of lemma 1.6.1 j then we obtain a sequence of approximating functions sn(x) = f Sn(x-g) dP(£) = £ Ajm) 0(m) a*****, where J i* (m) (i) | A„(m) | g 1, lim A„(m) = 1 (2.3.3) tt->00 and (ii) for every n only a finite number of' multipliers' Xn{m) are =f= 0. Hence, by theorem 1.6.2, the following conclusion: Theorem 2.3.1. (i) Every Fe V(Tk) is the weak limit of indefinite integrals of certain exponential polynomials, and any feL^, I<p<co is the limit in norm, and every continuous f is the uniform limit of
FOURIER EXPANSIONS 20 exponential polynomials, and the polynomials arise from the Fourier series formally by the insertion of certain multipliers Xn{m). (ii) If Fv F2e V{Tk) have the same Fourier series then they are equal, (iii) if S I <f>F(m) I < 00, then F is the integral of /(#) = 20(m) e27ri^x\ (2.3.4) On the basis of this theorem we can now introduce kernels in general. Theorem 2.3.2. Ifd(oc) is a convergence factor (definition 2.1.1)/or whichalso |Z(fl|SO(l + |fr^ (2.3.5) then for any Fe V(Th), the approximating function «*(*)= f KR(cc-£)dF(£) (2.3.6) J Eh has the Fourier series (m\ -U(m)e2^*>. (2.3.7) KJ Proof. We have already stated in section 1.1 that (2.3.5) implies S sup \K^-£i+^>->->Vk-£k+™k)\=MB<co. (m) x,£<=Tk For fixed E if we start out with & finite sum fn(x)=E0w(m) e2*'<w' *> (2.3.8) we obtain s\(x) = | KR(x-£)fn(£) dv JEk (m) W (m)e2;rz(m,cr) (2.3.9) by direct substitution. If now we take a general feL^T^ with a Fourier series ... „ ,, , 9.. ^ then by the preceding theorem there is a sequence of polynomials (2.3.8) with ||/n -/1| -> 0, so that automatically also lim <j>n{m) = 0(m). Now J-Eft therefore the Fourier series of sR(x) is the formal limit of the series (2.3.9) and hence the conclusion. For a general F in V, there is a sequence Fn of integrals of polynomials which converges Bernoulli to F on Tk, and this does not allow
30 FOURIER EXPANSIONS us to complete the last step of this reasoning quite so directly, but it suggests itself to make the following modification. We introduce the periodic kernel «) = S^+m), (m) as in section 1-6, which, because of (2.3.5), is also continuous, and the uniform convergence of J Tk towards sR(x) = KR(x - g) AF(£) J Tk was formally treated in lemma 1.5.2. However, the original reasoning on Eh itself had the advantage of applying step-by-step to the Stepanoff almost periodic functions as well; whereas, on the other hand, the description of a Stepanoff class of set functions F(A) which would correspond precisely to the periodic class F(Tfc) has apparently not been given. Theorem 2.3.2 has great technical advantages. For instance, if we put x = 0 in the formula J Bk (w) we obtain the following analogue to theorem 2.2.2: Theorem 2.3.3. IfforFe V we have <f>(m) £ — x(w>), where #(m) ;> 0, S^(rn) < oo, and ifF(£) satisfies the same boundedness condition at # = 0 as in theorem 2.2,2, then 2 | <fi(m) | < oo and F is the indefinite integral of Finally, forf€L2(Tfc), this implies again J, Tk (m) and in particular | /(£) |2 dv$=S | ^ (m) | 2> but this and the Riesz-Fischer theorem will be taken as known anyway. 2.4. Poisson summation formula We have seen in section 1-6 that forf(x) in L^E^) the series /fc) = Z/(*+*0 (2A1) (m)
FOURIER EXPANSIONS 31 converges in L^T^-norm, and that for any continuous periodic c(x) we have - f{x) c(x) dvx = f(x) c(x) dvx. J Tk J Bk Hence the following conclusion: Theorem2.4.1. Iff{x)eL1{Ek)iand 0(a) = ^a)= L e~27TZMdvx J Eh is its Fourier transform, then we have Itfix+m) ~ £ 0(m) e2**^ (2.4.2) (m) (m) in the sense that the series on the right is the Fourier series of the function (2.4.1). Thus in particular 2 e-" l™W20(m) e2^™'*) (m) converges in L^Tj^-norm tof(x), and it converges literally tof(x) at every point where f(x) is continuous or at least where its spherical meanfjf) is continuous. For practical purposes the following special conclusion is important : Theorem 2.4.2. Iff{x) in Lx(Ek) is continuous and the series (2.4T) is uniformly convergent and S | $(m) \ < 00. then relation (2.4.2) is a true equality at all points x. For any (y^), the Fourier transform of/(^)e""27r^»^ is $(a3+y3-), and thus (2.4.2) formally generalizes to S/(m+x) <r^m+*> y)=%<f)(m+y) e27ri(m> x\ (2.4.3) (m) (m) Furthermore, for any nonsingular afline transformation x' = Tx, k «-l <f>(oc) = (e-W^ftz')d%= J detT J \ e^^^fiTx)dvx. (2.4.4) But (<x, Tx) = (T'a, x), where T is transposed to T, and thus 0(a) = |detT|^(T'a), where ifr(a) is the transform of f(Tx). Therefore, \detT\f(a) = <f>((T')-1*)>
32 FOURIER EXPANSIONS and hence we obtain the following statement: Theorem 2.4.3. If <f>(a) is the transform off(x) then for any point x, y and any affine transformation T we have | det T | %f(Tm + Tx) e-^*<**+*.v> = 2; ^{{T')-1 (m+y)) e27ri(™>*l (2.4.5) (m) (m) The most renowned application of this is _1 j^eHnlt)\m+x\*~^e--7Tt\m\2+27Ti(m,x) (2.4.6) ^fc (to) (m) and the full affine version of this is V e-nQ(m+x)+27Ti(m+x, y) _ (^et Q}-\ N£ e-»Q'(m+i/)-2jri(w. x)^ (2.4.7) (w) (m) where Q(£) is a nonsingular real symmetric quadratic form and Q'(g) is its inverse5. If we take as known the formula J JE7 g-27rt|al+2jr«(t5t,aB)^ _. . !a a (t*+1»|*)«*+« then we obtain y p—tttt | m l+2rri(m, x) tt^+D S(**+|ro + a,|«)«*+» & for any & ^ 1, and especially 1 °° t °° _ y _ y p—27rt\m\+2nimx 7r^t2 + (m + ^ -^J for k=l. However, in this case the right side can be 'summed', the sum being l_e-4^ 1-2 e-27r'cos 2nx+e-*nt' and on this point we will still comment. Finally, we ought to mention the case of the ' discrete' i^-space in Ek which arises by taking as underlying space the set of all lattice points Mh -= {m} in Eh and making it into a measure space by assigning the measure 1 to each point singly. The Lv{Mk)-space thus arising is the space of sequences {f(m)} with ||/|| = (S|/(m)|^<oo, and the analogue to the set functions V(Mk) need not be analyzed
FOURIER EXPANSIONS 33 separately, every such set function being absolutely continuous now. Any element / with finite Lx-norm 2|/(*»)|<oo has a transform £ /(™) e~27^a> m>=<f> (a) (m) which is continuous and periodic, the inversion formula being /(rn)= e27r*w>a^(a)dv J Tk Uniqueness, convolution, and the theorem on 'positive transforms' are trivial now, and the analogue to the Plancherel theorem for L2{M) is part of the Riesz-Fischer theorem already mentioned. The approximating sums for convergence factors are now 8dm)=B*XKR(m-n)f(n) = [ s(^) <P(oc)e^^dvaf (m) J Eh W and, in particular, we have f e~^la|2^V(a)e2^(w'a)^^ J Eh (n) which is worth stating perhaps, for the sake of analogy* 2.5. SummabHity. Heat and Laplace equations Some of the findings of the preceding sections may be summarized as follows: Theorem 2.5.1. Whether we are dealing with a Fourier integral or a Fourier series / /• ^a^e^^dv^ * S0(w)e27r*(m,aj) if we take a convergence factor £(%) in Ek with 8(0) = 1 and put J E J Eh then the approximating functions e*"i&*)jfa)dva
34 FOURIER EXPANSIONS exist and they have the 'identical' value sr(x)-b4 K{B{x1^^)i...yR{^-U)Mo)dvt JEk This also applies to set functions F(A) instead of point functions f(x), and it also applies to Fourier series of Bohr and Stepanojf almost periodic functions. As regards Cesaro-Riesz summability we have found that its natural'spherical' version arises if we put \ 0 |a|>l [and not 8(a) = (1 — | a |)* for | a | ^ 1, which for &= 1 and k — 1 happens to correspond to Fejer's kernel], and in this case we have Z&)=Z»&)=ff*(|£|), where -(«)=J_!?£^> (2.5.2) where Jp(u) is the Bessel function. For large u this is 0((u)-s-^k+1^), and thus for , 8>~ (2.5.3) we have \KS{& | ^(l + l^l**')"1 with p = 8-%(k~l). Therefore, if 8 exceeds %(lc~ 1) then all our theorems apply. The partial sum s$B(x) exists at all points and, incidentally, has the value 4=W\M u^-^ Js+ik(uR) du, (2.5.4) and the behavior of sdR(x) as i2->oo is a local property only. For the limiting exponent £=|(& — 1)—we called it the 'critical5 exponent— it was shown by Riemann for k = 1 that the localization property holds even then, although mere continuity of fx(u) as u = 0 no longer suffices. However, for k >; 2, the situation is more complex and more interesting, and although localization still holds for f€Lx(Ek) and even/e V(Ek), it no longer holds for/eL1(Tfc). In fact, as we have demonstrated, for each k> 1 there is a periodic ^-function which is 0 in a neighborhood of x=03 and for which sR(Q) is unbounded as R->cq. Furthermore, we showed by a subsequent argument that for L2(Tk) localization does hold, and we raised the problem for L^T^), 1 <p < 2, and this problem is as yet unsolved.
FOURIER EXPANSIONS 35 We are now turning to the convergence factors e"n! a >2, e~2n'a' for an analysis of a different kind. In the case of the first factor we put t = R~h, so that 0<t<oo, and R->co is to be replaced by t~>0, and we denote the corresponding sR(x) by s(t; x), so that by theorem 2.5.1 for the function s(t; x)=± j e-<*W*-i\*f(g)dvs (2.5.5) we have the expansion e~^cc\2+27Ti(atx)^a^Va Qr 26-ar*|in|»+2ffi(«i»a0^(w) (2.5.6) J Ek (m) respectively. In the second factor, however, we put t=R~13 and for the function we then have e-tvtlal+torHchtiififa)^ or 2 e-a*« iwi+2» *(«,«) 0(m) (2.5.8) J J2& (m) respectively, and we recall that, for h = 1, for the last series we can also *** r d-^)/(g)^ J _j l-2e2^cos27r(^-g)4-e-47r<5 ^ ; alternately. Now, if in (2.5.6), say in the integral, we take the (formal) derivative with respect to t, then this introduces the factor -7r|a|2=-7r(a?+... + 4), and, except for a constant, the same factor can be obtained by forrning the negativeLaplacean .-% ^. A-—te+-+a?) (2'5-10) with respect to xl9...,xk. Also, since | <j>(a) \ ^M it is very easy to verify that these differentiations are legitimate, and thus the function (2.5.5) is a solution of the heat equation 3<? 1 a7=-_AS, (2.5.11) of which the given function /(£) constitutes boundary values as t->0 in a manner described in our theorems. Next, in the case of the integral (2.5.8), one differentiation with respect to t introduces the factor — 27r|a| which cannot be compensated by differentiations 3-2
36 FOURIER EXPANSIONS with respect to the variables x3> directly, but two differentiations with respect to t can be so compensated, and in fact we obtain the equation ¥2 = A,s (2.5.12) which may be interpreted as a Laplacean in h +1 variables first of all. For &=1, formula (2.5.7) is the familiar solution for the boundary-value problem of harmonic functions in (y, t) in the half- plane t>0, and formula (2.5.9) is the even more familiar solution in the unit circle | z | < 1, for the complex variable % = e-27T^t+ix\ However, from another approach, it is indicated to keep in (2.5.12) the variables (t, x) apart, and to take the operational square root on both sides of (2.5.12) writing it thus !=-Ak (2.5.13) as we will still discuss. This done, (2.5.13) falls under the general form of an operational equation ^ i=-A*f> (2-s-14) in which Ax is a linear operator with certain positivity features as possessed by the Laplacean, primarily but not at all exclusively, and all such equations (2.5.14) will be interpreted as diffusion equations, as we will still see. 2.6. Theta relations with spherical harmonics We now denote a Fourier transform in Eh by gfe)=f e'^y^f(^)dvXi (2.6.1) J M and in both spaces we introduce spherical coordinates which we will denote by . , r ~0 „„ , Actually, in such coordinates the point of origin (x5) = (0) is a singular one, and, strictly taken, it must then be excepted in virtually all statements; however, we will do so explicitly only whenever its exceptional status is a material and not only a formal issue.
FOURIER EXPANSIONS 37 A function f(Xj) is radial if there is a function f0(u) in 0 < u < oo such that f(^)—f0(\x\), and it turns out that for such a function the transform (2.6.1.) is likewise a radial one, g(^)=g0(| y|)} and the functions of the radius are connected by a so-called Hankel transform 2tt f00 (u\ 9o(v) = - J o /o I - J Jjm (2t») «** *» -,-3-i /o(«)-7**-i (a™") «***«• (2.6.4) "" Jo If, however, the 'angular' variables ^ do occur, then the dependence on them is best expressed by means of spherical harmonics which we are going to introduce now. Definition 2.6.1. (Somewhat ambiguously) a (spherical harmonic) is a function either in Ek, or on the unitsphere Sk: E>\-f... + £f=1. If in Ek, a harmonic of order n, Pn(xi) *s a homogeneous polynomial of order n which is a solution of the Laplacean /a2 a2 \ -A,Pn(^)^^ + ...+--|jPn(a;,) = 05 and if on Sk it is a function Pn(^) which is connected with a Pn(Xj) by means of Pn(&) =Pn(av/| a? |) or, equivalently, Pn(ay,) = | a; |nPw(£y), such a function satisfying the equation AzPn&)=-~n(n + Jc-2)Pn(Q, where Ag is the e curvilinear' Laplacean on 8k. Lemma 2.6.1. If Hn(t; £,-) is continuous in 0<t<co, ^eSki and is for each t an n4h harmonic in (^) and if the integral rCOHn(t;^)dt=Hn(^) f o is absolutely uniformly convergent then Hn(^) is again an n-th harmonic. In fact, for given n, the nth harmonics are a vector space with finite basis, and a uniform limit of polynomials in real variables of the same degree is again a polynomial, and all (mixed) partial derivatives are uniformly convergent likewise. Lemma 2.6.2. For any harmonic Pn(£) the formula e-2"«Pn(£) du)^2ni-Pn{i) Jn+i^Z) (2.6.5) Sk holds. r. L
38 FOURIER EXPANSIONS This formula is equivalent to a general version of the addition theorem for Bessel functions and will be taken as known. Theorem 2.6.1. Ifffa)£L1{Elc) is of the form /(*,)=A(|*|)Pn(g), (2.6.6) then its transform (2,6.1) is of the form g(yi)=M\y\)Pn(v), (2.6.7) where Pn(Q is the same harmonic in both formulas and ^v)ss'WJQ Hv) Jn+i^i(2nu)u^du. (2.6.8) Proof. For spherical coordinates x3- = tg3- we have dv^^^^dtdco^ and hence f «> * / f \ 9(y>)=\o X{t)tk\]s e-2nit]MPn®da)zJ dt, and if we substitute (2.6.5) the assertion follows. If we replace A([#|) by A(|#|) |#|nand fi(\y\) by /*(|y|) \y\n, we obtain the following variant: Theorem 2.6.2. Iff fa) in Lx{Ek) is of the form ffa) = \(\x\)Pnfa), then g(yi)=M\y\)Pn(vi) 2ffin f00 [u\ and aw=-^ \QHv) J^-i(27m>unHk du> <2-6-9) and therefore, except for a factor in> the connection between \(u) and /i(v) in the case (k; n) is the same as in the case (2n + k; 0). Now, in any dimension, the transform of e-*"'x >2 is e~n'y '2} and hence the following conclusion: Theorem 2.6.3. For any harmonic we have I Pnfa) e-n >»i2 e~^^^ dvx=inPn{y3) er"ly |2, that is, Pnfa)e~lx^ is an eigenfunction, with eigenvalue in, for the Fourier transform operation in Ek. If we now apply theorem 2.4.3 we obtain the following general theta relation: Theorem 2.6.4. For any spherical harmonic P^,...,^), r = 0,l,2,...,
FOURIER EXPANSIONS 39 we "have SPrK, ...,m7c) e-^K+-+m!)=^^-^ S ^-(^i, ...,mfc)e-(^K+-"+™!>, (w) (m) (2.6.10) /or 0 < t < oo. More generally, for all x, y we have £ Rr(m + X) e-^(™+a)+2™2j(mj+ai) Vj (m) -=i^-^7c(det Q)-4 ^ i^(m-f.y) e-(7r'^'(m+2/)-2?ri-^i+^)a;j} (2.6.11) (m) where Q(m) is any positive definite quadratic form in (mv ...,mfc) and Qf(m) is the inverse form, and where Rr(n) arises from a harmonic polynomial Pr(£) by a substitution £=An which transforms 2 m| into Q(m), and where R'r(w) arises by the reciprocal transposed substitution For our next application the following fact will be taken as known: Lemma2.6.3. If/i, v are any complex numbers with Re (ji + v)> 0, then J-oJo lim e-etJv(t)W-1dt=- for all /i, v. If v—ju, + 2=—m) in which case the gamma factor in the denominator is oo, then th,e limit has the value 0 accordingly. Using this, if in theorem 2.6.1 we put X(u) = ((27r)* u)y e~eu then we obtain the significant relation lim J e~e i»i ( (2tt)* | a; | )r Pw(&) e-2^> *) dvx 2'P»(7) 2*>^rft(y+t+»)} ((27r)i|y|)*+r r{i(n-y)} * ' ' ' for | y | =[= 0. The constant y may be complex, and if we want the integral to be absolutely convergent we must put Re y > — Jc. However, the relation also holds for Rey> — k—n, with the integral existing at the origin as a Cauchy principal value (spherical). A remarkable situation arises for y= — \h which we will describe separately. Theorem 2.6.5. For any harmonic Pn(£) we have lim f ^@e^i^e-2^^)d^-i-^-^} and thus Pn{£>j) \x\-^k is an eigenfunction of the Fourier transform operation, for the eigenvalue in.
40 FOURIER EXPANSIONS 2.7. Expansions in spherical harmonics We now introduce the coordinates (2.6.2), (2.6.3) into the general relation (2.6.1), and if we put /(|a?|;6)3/(2,,), g(\y\;y^g(yi), (2.7.1) then this relation becomes g{v; 7,.) =1 M e~27rivu^^f(u; &) do) A uh"1du. (2.7.2) In order to utilize this we now take it as known that any function 0(6) on Sk_x which is continuous, or only belongs to Li($&), has a 'Fourier expansion' in spherical harmonics, #&)-:£<*«&), (2.7.3) n—0 by which it is uniquely determined, and we introduce these expansions for our functions (2.7.1), denoting them thus: /(";&)~S/»(w;&), (2.7.4) 71=0 00 ?(«;%)~Z ?.»(«; ft)- (2.7.5) If now we substitute (2.7.4) in (2.7.2), and apply the uniqueness property, then we obtain the following conclusion, which can be established rigorously for any class of functions for which (2.6.1) can be defined: Theorem 2.7.1. In spherical coordinates, the Fourier transformation (2.6.1) becomes the set of relations 2jrin r°° fu \ 9n(vi ^) = -^r /nb; VA Jn+ik-i^^^du, n = 0,1,2, .... (2.7,6) Our first application of this will be a kind of converse to theorem 2.6.1. Theorem 2.7.2. If a measurable function X(u)inO<u<oo is such that |»oo |A(^)|e^wdu<oo (2.7.7) /*oo for all A>Q, and if X^u^du+O (2.7.8)
FOURIER EXPANSIONS ^ for allcr>0 (thus, for instance, if \(u) > 0); and if for a function 0(g,) in L^Sjc) the function ... f(*i) = M\x\)M£i) (2.7.9) is such that its Fourier transform g(yj) is likewise of the form 9(y^M\y\)^(vj); (2.7.10) then <f>(£s) must be a 'monomial* spherical harmonic Pn(£), and g(yj) is as in theorem 2.6.1. Proof. If we introduce the expansion (2.7.3) then we have and by (2.7.6) we obtain ffn(v'>Vj)=Mv)$n{Vj)> ra=0,l,2,..., (2.7.11) 27Tin f °° where pn(v) = ----_ X(u) Jn+ik^ (2ttuv) u%kdu. (2,7.12) But by (2.7.10) we have alternate values gJviy^Mtffnivj)* n=o,i,2,..., (2,7.13) and by comparing this with (2.7.11) we find that if for an index n-=0,1,2,... we do not have &»foi)«0, (2.7.14) then there is a constant cn such that /tnW88^). (2.7.15) Now, by assumption (2.7.7) we may substitute the expansion in the integral (2.7.12) and integrate term-by-term. This gives rise to a convergent expansion and by assumptions (2.7.8) we have aM>0 + 0 so tnat tne expansion begins with vn effectively. Therefore, we cannot have (2.7.13) for two different indices n, and thus we must have (2.7.14) for all except one index n, and this is precisely what the theorem claims.
42 FOURIER EXPANSIONS Now, the function \{u)=*urer™*> r^O, satisfies (2.7.7) and (2.7.8) and hence the following very.special inference: Theorem 2.7.3. If we have for two homogeneous polynomials PT{x5), Qs(yj) of some degrees r, s, then Pr{Xj) is a spherical harmonic, and we have s=r, Qs(y3) = inPT(yj). In particular, if for a homogeneous polynomial P^) we have I Pr(xj) e-»i*iV2*<(v.») dv^cPrfa) e~"^\ (2.7.16) Eh then Pr{xj) is a harmonic and c = in. We now recall that the (inhomogeneous) Hermite polynomials in one variable En(x), n=0,1,..., after replacing x by (Ztt^x, have the property Hn(x) e-™2 e-27riv* dx = inHn(y) e-w2. However, any polynomial Pn(^) can be expanded into a finite series S a^^Hnfa) ...Hnk(zk), (2.7.17) m + .-. + njc^n and uniquely so. Therefore, the transform of Pn(x}) e~"|a!|!! is where QB(y,) = S W"^-^^...^^^...^^), and since Pn(<ty) is a homogeneous harmonic if and only if we obtain the following theorem: Theorem 2.7.4. A homogeneous polynomial Pn{xs) is a spherical harmonic if and only if, on representing it by an expansion (2.7.17), only such coefficients animtt njt are =f= 0 for which n1+...+nk=n—4cr, r = 0,l,2,.... Turning now to a somewhat different topic we note that relation (2.6.9) implies as follows: Theorem 2.7.5. Subject to assumptions, if we are given on Sk a
FOUBIEB EXPANSIONS 43 function*})^) with the expansion (2.7.Z), and iffor someywithB>ey> -k we consider on 8% the function with the expansion then the function ffa) = ((27r)* \x\)y $(£;) has for fa) 4= (0) the transform ^•) = ((27r)4|y|)-r-^r(7;,)3 in the sense that we have g(y,) = lim e-« I * i/fo.) e~2™to ») &>„, e\0JEjc the limit existing. Now, as for specific assumptions that are sufficient, we quote the following theorem: Theorem 2.7.6. For given y, if 2n is an even integer for which 2n>Rey + p — 1 and if 0(gy) belongs to differentiability class C®n) on $£, then the function (2.7.18) exists and is continuous and the conclusions of theorem 2.7.5 hold true. Also, if q°^Q is an integer for which 2n>Rey+p~ 1+g0, then pfa) e C<«°>, and thus if(j>{^) e C(°°> then also pfa) € C&\for Re y > - h. Finally, if 0(^) is {real) analytic on Sk in its 'natural' local coordinates, then ifry(£j) is also analytic. The last-made statement on analyticity is non-obvious. The theorem applies in particular to where T2h(^j) is a homogeneous polynomial which is > 0 for | x | =(= 0, and s is a complex number for which Res<~. (2.7.20) It suggests itself that the function g(y; s) might also be holomorphie (that is, complex-analytic) in the parameter s in the half-plane (2.7.20), and this is not only correct but easy to prove. Altogether we will have in the next section a rather general theorem from which we will deduce at the end of section 2.9 that the function (2.7,19) is g(y,;s)=limf e-*\*\\m *]^, (2.7.19) e |0 J E
44 FOURIER EXPANSIONS holomorphic in s and C(c0) in (y5-), for (y5) =|= 0. But the actual analyticity in the real variables y3- is a much more specific proposition and the proof will not be reproduced here. 2.8. Zeta integrals Definition 2.8.1. We say that a function f(xf, s) is an adjusted Zeta kernel if it has the following structure: (i) It is defined for (x^) in Ek> and for complex $ in a certain domain D. (ii) It is 0 in a fixed neighborhood | x \ < x° of the origin in Ek. (iii) It is C(00) in the variables x^ and for any combination of integers ^. . A /rt n _ v 6 ft^O, ...,#*£(> (2.8.1) the function D*i~*tf(x; *)s d^x^^[ (2.8.2) is continuous in (x; s) and holomorphic in s3 and what is decisive (iv) corresponding to each s° in D there exists a neighborhood N of s° and a real number y=y(N)y y^O, such that for each combination (2.8.1) we have as | x | ~> oo, uniformly for s in N. Definition 2.8.2. In the present context a convergence factor is a function d(t) of the following description: (i) it is defined and continuous in 0 ^ t < oo, and 8(0) — 1; (ii) (t) is C(oo) in 0 < t < oo; (iii) we have dpS(t) ~^ = 0{t-«) as t-^oo (2.8.4) for every p ^ 0, q ^ 0, so that in particular S(t) = 0(t-«) as t~>oo (2.8.5) for every g;>0; and (iv) we have dp8(t) —±L = 0(1rv) as t->0 (2.8.6) for every p > 0 (but not p = 0). The function £(t) = e~*P is such a convergence factor for every p > 0. Definitions 2.8.1 and 2.8.2 are such that for any e>0 the product fe(xj;s)~8(e\x\)f(xj;s) (2.8.7)
FOURIER EXPANSIONS 45 we obtain the relation is again an adjusted Zeta kernel; and, in consequence of (2.8.3) and (2.8.5), this new function and each mixed partial derivative of it is small at infinity in such a manner that for the transform gjly* *H f er^^v)fe(xf, s)dvx (2.8.8) J Ek ition (27riy1)vi...(2iTiyk)^ge(y,s)= f e-*«*-*D*v'**fe(x;8)dva (2.8.9) J Ek for any integers (2.8.1), by means of partial integrations which are readily justified. But now we claim as follows: Theorem 2.8.1. If N and y-y(N) are fixed, then for Pi+-+Pk>Y + k> (2.8.10) we have lim I e-*ir*t*.v)])91..-Pkfe(x',8)dvtt = \ e-2nil*>»D*i'"**f{x;8)dval9 e 10 J Ek J Ek (2.8.11) uniformly in the neighborhood of every point (y, s). Proof. If we write D® for Dri—n and put r~rx+... + rfc, then we have DWfe(x; s) =8(e | x \) DWf(x; s) + £ DM6(e\x\)D®f(x; s). (2.8.12) r>0 Now, by (2.8.3), we have DMf(x; s) = 0 (| x |t-»), | x | ~>oo, and by (2.8.10) this function thus belongs to Li(i7&), uniformly in N. Also, 8(e \x |) converges boundedly to 1 as e->0, and thus our theorem will follow if we verify lim f | JDW*(e | x |) |. | W>f(x; s) \ dva=Q (2.8.13) e 10 J Ek '* ' \ & * J uniformly in 5, for r+s =p,r>0. Now, since/(»; s) — 0 for ] a; | <a;0 it is not hard to verify, by the use of (2.8.3), that the integral (2.8.13) is majorized by | _ 8(et) . tys+K-1 dt=e*-y-* \ \ &r)(t) | tr-s+^-i dt. J a* \dtr I J ea!o However, by assumption (2.8.10) the quantity p=p-y—k is >0, /•oo and thus e*>-y+fc tends to 0, as e -> 0, for any fixed 9/, no matter how Jv
46 FOURIER EXPANSIONS small. Now, if r>05 then, in ex°^t^n, \ 8®(t) | is ^cr(7})t-r, where cr{rj) ~> 0 as tj -> 0, and thus we only have to verify that stays bounded as e~>0. But, due to r+s=*p, this is a trivial fact, and hence the theorem. The theorem implies in particular that for 2n>y + k (2.8.14) the function (yj+... +y\)nge{y; s)=\y\*nge(y; s) (2.8.15) has a limit, as e~>05 uniformly locally in (y,s)3 and this limit is independent of the special factor 8(t) employed. However, we can divide the function (2.8.15) by the factor | y |2n whenever | y | +0, and we thus obtain part (i) of the following theorem: Theorem 2.8.2. (i) For any adjusted Zeta kernel the transform g{yf, *)~\ e-^^f(xj; s) dv* (2.8.16) J Eh exists for | y 14=0, as a limit g{yj;s) = limge(yf,s), (2.8.17) e 4-0 uniformly in the neighborhood of every point (y,s)i and the limit is independent of the particular convergence factor used in (2.8.7). (ii) The function g(yj; s) is holomorphic in s and C(00) in (y^), and the derivations with respect to the variables y3- can be performed under the integral sign in (2.8 J6). (iii) We have . «,. , „, . . /ft«,«v QAyi\*)=*0{\y\-*), as \y\-+co (2.8.18) for every q>0, uniformly say in 0<e<l and in the neighborhood of every point s. Proof. Ad (ii). The holomorphy in s follows from the fact, which is a consequence of Weierstrass's limit theorem, that a definite integral h(x; s) dvx is holomorphic in s, whenever h(x; s) is holomorphic in s /■ and the integral is approximate by finite Riemann sums locally uniformly in s. Next, for given e>0, relation (2.8.8) implies tyeiyf, *) dy. 3—=\ e-27ri^^S(6\x\)(-2nix1)f(xj;s)dvx, (2.8.19) i J Bh
FOURIER EXPANSIONS 47 say. But xj(xf, s) is again an adjusted Zeta kernel, and thus, by part (i), the function (2.8.19) must have a limit, as e ~>0, locally uniformly, and it now follows that this limit must be this derivative existing. We can now iterate this procedure, and the assertion follows. Ad (iii). The function (2.8.15) has, except for a constant, the value f e-^>v)&%f(x3.',s)dvx, where A is the Laplacean. It follows from our proof to theorem 2.8.1 that, for 2n>y + k, this function is bounded in yj9 uniformly for 0 < e < 1 and $ in N, but n can be chosen arbitrarily large whence the conclusion. 2.9. Zeta series The unnatural restriction incorporated in definition 2.8.1 that f(x; s) shall be 0 in a neighborhood of x = 0 was meant to be only transitory, and we are now going to allow, to the contrary, that it shall even be undefined and singular at x=0. Definition 2.9.1. We say that a function/^; s) is a (proper) Zeta kernel if it is defined for xhxEk except at the origin x—0 and for $ in a domain D, and if it has properties (iii) and (iv) of definition 2.8.1. Theorem 2.9.1. (i) Iff(xf, s) is a Zeta kernel then the corresponding 'Zeta fusion' ^. s) = s,e_«/(%.,} (2.9J) (m) exists for 2/4=0 on the torus Tk:—\-£yj<\ and for s in D as a limit Urn %'e~^m>ti8(e | m |)/(%; s) (2.9.2) e 10 (w) for any convergence factor 8(t), and has a value independent of it, uniformly in the neighborhood of every point (y}s), and it is C(c0) in y and holomorphic in s. (ii) Also, corresponding toy—Q, the limit ]im[zfd(e\m\)f(m3-;s)- f 8(e\x\)f(x,;s)dv^\ (2.9.3) ejoLw J\x\^l J exists, continuously for s in D, and is holomorphic there.
48 FOURIER EXPANSIONS Proof. Ad (i). We take it as known that it is possible to construct a continuous function #(t) in 0 < t < oo which is C(00) in 0 < t < oo and for which we have %(t) = 0 for 0 ^ t < | and #(t) = 1 for f < x(t) < oo. If now, with the given function f(xf3 $), we construct the new function ^"'H o, *~o, then/(#,-; s) is on the one hand an adjusted Zeta kernel, and on the other hand we have /(^; s) —/(#,•; s) for | # | ;> § so that in particular we have ^,^^3,)^(m.. 5) = j e-2*i(«.s/)/e(%; 5). (w) (w) However, by theorem 2.4.1 we have S^K-+^) = S^^(m,.)/e(m.; 5), (2.9.4) On) where ge is the Fourier transform of/, and if we apply all properties stated for ge in theorem 2.8.2 then our assertion follows. Ad (ii). For y=*0 the term ge(0) need not have a limit as e->0. However, if we write S/^(wi)=S7«(%;*)-^(0), (m) Cm) then the left side does have a limit, and a holomorphic one too, again by theorem 2.8.2. Now and if we write for this J|*|Sl J*£|0l.Sl then the second integral is holomorphic in $ for $ in D, so that we need subtract only the first integral, and this was done so in formula (2.9.3). We now take a homogeneous polynomial T2h(xj) > 0, for #4= 0, as in section 2.7, and with it we set up any (nonhomogeneous) polynomial Tfa) ~ T^Xj) + (lower powers), and we assume that we have T(xi) + 0 for #4=0. For &§:3 it then follows by monodromy that we can define (^(x^)8 as a holomorphic function in s for #4=0 and all complex s, and for h—2 we make this into an additional assumption (which is needed only if T(xs) is non-
FOURIER EXPANSIONS 49 homogeneous). We take some further polynomial Qr{x/j9 say homogeneous of degree r, and if we put/(a?,; s)^Qr{xi) T{x^s, then this is a Zeta kernel with D being the entire s-plane. Therefore, theorem 2.9.1 implies immediately parts (i) and (ii) of the following theorem: Theorem 2.9.2. (i) The function £(y,.; s) =lim£'e-27r^»m>8(e\m\) Qrfai) e|0(m) T(m5-)" exists, for y=£0 on Tk, and is an entire function in s. (ii) The Dirichlet series and the associated integral I Qri*j) dvx^H(s), (2.9.5) being both absolutely convergent in the right half-plane Bes>^-, (2.9.6) the holomorphic functions defined by them in this half-plane are such that their difference has an analytic continuation into the entire plane, ' £(s)—E(s) = entire function. (iii) Taken separately, the functions £($), H(s) have meromorphic continuations into the entire s-plane. If T(mj) is homogeneous, then H(s) and £(s) have (at most) one simple pole at the obvious point If T(mj) is nonhomogeneous, there is first of all again a simple pole at the same point (2.9.7) with Hhe same' residue (2.9.8), where T2h is the highest part of T; and there may also appear other poles, all simple ones, at the points , ^r + k-m m==l,2,3,4,..., (2.9.9) 2h which are placed at equal distances to the left of the first one.
50 FOURIER EXPANSIONS Proof. Ad (iii). In the homogeneous case we have in the half-plane (2'9'6) I™ r ***** f &® *, and in this product the first factor is (2hs —r—Jc)"1 and the second is an entire function, which gives the conclusion. If T is not homogeneous, we put T — Tm + Pm-I " ^2fc( 1 + ~m I > and if we substitute this in the integral (2.9.5) we obtain, at first formally, = 2 (7) f, , ^-1<to- n*0\ W / JlalSl i2fe Now, the nth term in this sum is of the kind just discussed, with s replaced by s+n, and Qr replaced by QrP^-.lf and since the last expression is an inhomogeneous polynomial of degree ^(2h —l)n, we see that for n > 0 the term has at most simple poles at the points s+n= , 0£v£(2h-l)n9 and is holomorphic otherwise. Finally, corresponding to any bounded domain D of the 5-plane we can find an M such that on putting the remainder r ° BM(xj; s)dvx J ! x |^l J-2U will be absolutely uniformly convergent for s in D and thus will be holomorphic. From all this our assertion follows. Finally, returning to integrals instead of series, we take a function <f>(^) on Sk as in theorem 2.7.6, say in C(00)? and we put f(xj;s) = \x\^*<f>(£j) with any real numbers a, b+0, and we consider this in the half-plane a + b. Re s > — h for which the transform j e-27ri&>y°)e-eWf(x3-;s)dvx J Eh
JOUKIEE EXPANSIONS Si is definable. With our previous function x(t) we now put fix,; «) = x(| *|)/C«v, «) + (l -X(\ x \)f(xf, «)=/*(*,; a)+fH&; s), and then put ^. s)=hm f e-^iix,y)e-ela!lj>r{Xj; s)^ (2.9.10) e 10 J Eh for r=l,2. Now, g%i; s) falls under theorem 2.7.6, and is therefore C(°°) in y and holomorphic in s. For the function g2(^-; s) this is, however, likewise so, simply because for this function the integration in (2.9.10) extends effectively only over the unitsphere | x \ ^ 1. Therefore, the function gx+g2 is likewise so, and this proves a statement made about the function (2.7.9) near the end of section 2.7. 4-2
52 CHAPTER 3 CLOSURE PROPERTIES OF FOURIER TRANSFORMS We will now present an analysis of the principal closure properties of characteristic functions in general, and for those of (infinitely) subdivisible processes in particular, and doing the latter immediately in Ek for general h will require a rather elaborate setting indeed; and such simple functions as are ei(et*^ and 1 —cos (ax) will be replaced by more general ones (which we will term 'pseudo-characters' and 1 Poisson characters' respectively) for the sake of greater clarity, we hope. 3.1. Pseudo-characters and Poisson characters Definition 3.1.1. A function X(oc; x)^x(ch> —><**; «i> —»«*) (3.1.1.) is a pseudo-character if (i) it is defined and continuous in E2k=E%xE% and is bounded there, | %^. x) | ^M^ (3JL2) (ii) for (a) — (0), %(<%; x) is a constant in x, but not==0, X(0;av) = Co=|=0, (3.1.3) and, what is decisive, (iii) we have lim I jrto)x(a,;a,)dfoa = 0 (3.1.4) I cc |—>oo J Eh for every KeL^Ejc). Theobem 3.1.1. The class of pseudo-characters is closed under uniform convergence in E21c, except for property (ii). In fact, if (3.1.4) holds for a sequence xn having a uniform limit Xo> then it also holds for #0. On the other hand, for fixed bounded #, if (3.1.4) holds for a sequence Kn, for which || Kn—K || -> 0, then it also holds for K. Now, step functions on the finite intervals Iab: aj^xi<b5 are dense in Lx(Eh), and hence the following statement: Theorem 3.L2. A bounded continuous function #(a; x) fulfils (3.L4) if we have r lim x(x'>%)dv^0 (3.1.5) 1 x |->oo J lab for all finite multi-intervals.
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 53 In particular, we have L e**>xHva=z n i-l ix* and it follows from gibx giax %X <2|6-g| ~ l + \x\ that (3.1.5) holds, whence the following statement: Theorem 3.1.3. The function 6%**$ is a pseudo-character (Riemann- Lebesgue lemma). Since e*^***) is also a pseudo-character, for any A, we may also state as follows: Theorem 3.1.4. IfB(t) in — co<t<ao with 5(0)4=0, is a sum Scwe<Aw* (3.1.6) (m) with Am + 0, E|cw|<oo, On) or a uniform limit of such sums (Bohr almost periodic function with Am 4= 0), then the function X(a;x)=B(a1x1 + ... + cckxk) (3.1.7) is a pseudo-character. For instance, 2 cos (a, #) =e*<a»*> + e~iCa»fl0 is a pseudo-character. Definition 3.1.2. A Poisson character is a function Q(a; a?) in Elc x jE^ of the following description: (i) Q(^x) = c1{l-x^\x)), Cl>0, (3.1.8) where ^(a; a;) is a pseudo-character, and #(a; a) is real valued, and Q(oc;x)>0 (3.1.9) and also Q(0; a;) = 0for alia;, and <9(a; O) = 0for alia. (3.1.10) (ii) The function q(x) = \ Q(jS; x) dvfi (3.1.11) is strictly >0 for x4=0 (and not only ^0 as (3.1.9) implies) and O(oc x) (iii) the quotient P(a;^) = ^Vr, #4=0, (3.1.12) q(x) is bounded and uniformly continuous in the variable a in the point set 0<|a|<a0, 0<|a|gl (3.1.13) for every a0 > 0.
54 CLOSURE PROPERTIES OP POURIER TRANSFORMS Important remark. We do not assume that (3.1.12) is uiiiformly continuous in (3.1.13) as a function in (oc, x) but only that it is uniformly continuous in a, uniformly in x. For & = 1, for the classical function Q(oc; x) = (sin ocx)2 and for similar ones the quotient does happen to be continuous in (a, a?) simultaneously, and all proofs, known to this author, of the structure theorem for subdivisible processes make full use of this simplification; but for h > 2 this definitely ceases to be so and a more systematic approach similar to ours must be sought. We note that in (3.1.11) and (3.1.13) the unit spheres could be replaced by others with fixed radii, equal or not, and that, on the other hand, the constant cx in (3.1.8) could be put equal to 1, which in the reasoning we will frequently do anyway. Next, if o)s(fi) is the' indicator function' (characteristic set function) of the unit sphere | fi | % 1, and if, on putting c1 — 1, we write q{x)=\ o)s(/3)[l-X(P;x)]da)ji, then we obtain q(x) = c2—p(x), (3.1.14) where c2 > 0, and the function J Ek tends to 0 as | a; | ->oo by definition 3.1.1. Therefore, in connection with property (ii) of definition 3.L2 we obtain the following further property automatically: (iv) For any fixed n > 0 we have M1<q{x)<Mz (3.1.15) for |a?[^??> withO<M1<M2<oo. Theorem 3.L5. // B{t) in — oo<t<oo is periodic or Bohr almost periodic, if it is non-negative and even, B(t)^0, B(-t)=B(t)t (3.1.16) and if in a neighborhood of t—Owe have B(t)^c2\t\™+o(\t\™) (3.1.17) with c2 > 0, m> 0 (m not necessarily integer), then Q(cc; x)=B(a1x1+... + cijtxk) (3.1.18) is a Poisson character.
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 55 Also we then have q(x) = c31 x \m+o(\ x \m) (3.1.19) for \x\->0and P(a;a) = - \x\ + "t+CCkjt\\ +P*(a^)' (3'L2°) where P*(a; x) is continuous in (a,x)for oceE^, a?4=0 and lim P*(ct;a;)-=0 (3.L21) uniformly in \ oc | ^ a0 for a?M/ a0 > 0. Proof. For an almost periodic B(t) there exists the mean value i r B{t)dt 3ssKm7T I T-^oo ■* J 0 which in our case must be >0, and if we put P(t)=c(l —B0(t)) then X(cc; x)~=B0(oc1xl + ...+ctkxk) is a pseudo-character, first of all. Next, if in the /?-space we introduce polar coordinates (spherical), and if for given vectors (x§) and (/?,) we denote the angle between them by 6, then we have \{j3,x)\mdvo^\ < |yff|w.|al".(cos0)mefo, Jl^lSi J\fi\zi = |#|m \fi\m.\cosd\mdv0=c5\x\m, and therefore also, if (#)-»0, \o{\(fi,x)\)\"dvfi=o(\x\<*), I '- o(\(cc x) \m) which proves (3.1.19). Finally, an easy discussion of -, 9. -— will |#| prove (3.1.21). 3.2. Pseudo-transforms and positive definite functions We take the set of all continuous complex-valued functions in Ek: (oCj) and we define as follows: Definition 3.2.1. A sequence{ifrn(<x)} is called P-convergent (towards its P-limit), in symbols ? !M*)-*J0r(a), (3-2.1) if we have lim ^w(a) = f{ct) (3.2.2) uniformly in every compact sphere 0^|a|S<x0, a0>0, (3.2.3)
56 CLOSURE PROPERTIES OF FOURIER TRANSFORMS A set {i]r(cx)} will be called P-closed if it contains the P-limits of its sequences. The set of all continuous functions is P-closed and therefore to any subset $={^(a)} there corresponds a smallest P-closed set S containing it ('the P-closure of $'), and it is easily seen that S arises by adding to S all the P-limits of its sequences. Definition 3.2.2. Given a, fixed pseudo-character, a function 0(t%) will be called a pseudo-transform if it can be represented with some F in F+ in the form * <f>~<f>F(a)=\ X^\x)dF{x). (3.2.4) Obviously, <f>(cc) is bounded, |^(a)|^0.||P||, (3.2.5) and is continuous in a although perhaps not uniformly continuous mEk. If we are given a sequence 0»(*)=f x(^;x)dFn(x), (3.2.6) J Ek then ^(0) = c0PM(^) (3.2.7) implies that we have | Fn(Ek) \£MX (3.2.8) whenever | <fin(Q) | <iM; (3.2.9) and also if a sequence (3.2.$) converges a.e. and (3.2.8) holds, then the sequence converges boundedly a.e. If the sequence Fn is weakly convergent, then the sequence <f>n(a) need not converge at all, but if Fn is Bernoulli convergent to P0> then $n(a) is P-convergent to &>(*)= f x("',x)dF0(x) (3.2.10) J Ek (see lemma 1.5.1). Theorem 3.2.L (i) The set of pseudo-transforms is P-closed. Also, if a sequence (3.2.6) is convergent at all points, 0w(a)-^(a), (8.2.11) and if <fi(oc) is continuous at the origin, then <f>n(*)^<f>(cc) (3.2.12) and Fn is Bernoulli convergent. (ii) // (3.2.11) holds a.e. and also (3.2.9) holds, then Fn converges weakly to F0 and <p{<x) — <fi0(ot) a.e.
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 57 Proof. Due to (3.2.9), and hence (3.2.8), the sequence Fn has a weakly convergent subsequence. We denote the latter by F'n and its limit by F'0i and we introduce the transforms &(*)=] Xtotfffln&h ^=0,1,2,3,.... (3.2.13) With a family of kernels {KR(oc^)} in Lx we introduce the functions which are connected by $'rd*)=\XRfax)dFtix)' Now, by the very definition of a pseudo-character, ^(a; z)~*0 as | x | -»oo, uniformly in (3.2.3), and therefore p However, if (3.2.11) holds, then lim 0;E(a) = ^(a)= KR(<£-fi)<f>(J3)dvfi9 n->oo J and thus we have 0oE(a) = 0u(a)- Letting B-^co, we hence obtain by theorem IT 0j(a) = 0(a) at all points where both functions are continuous. However, 0 '0 (a) is continuous by construction and <fi(a) is continuous at oc = 0 by assumption, so that we have ^q(0)= lim ^(0). By (3.2.7) we thus have -^o(-'fc) — l*111 F'n(Ek), and thus Tn is even Bernoulli convergent and p • therefore 0n(°O -> 0o(a) = 0(a) • Now, by explicit assumption, the entire sequence <j>n{ot) is convergent to (f>(cc), and by what we have just proven, any subsequence of <j>n(a) must contain a sub-subsequence which is P-convergent and this implies that the entire sequence must be P-convergent to begin with. For the second part of the theorem we take any continuous function C(fl) which is zero outside a compact set and obtain and thus 0o(/O = 0(/?)? a*e- as claimed.
58 CLOSURE PROPERTIES OF FOURIER TRANSFORMS Theorem 3.2.1 applies in particular to transforms proper <p{a) = <f>F(ac)=\ e~^i^HF{x), (3.2.14) J Eh and for these the following property is supplementary: Theorem 3.2.2. ForFe V+ we have | <f>F(a+h)-<?>F(a) |2^2 ||F||. Real part (<f>F(0)-<f>F(h)), (3.2.15) and, more generally, for FeV,ifF is the absolute value of F and $ is its transform then we have \<f>{a+h)-${oc)\2S2\\F\\. Real part ($(0)-$(h)). (3.2.16) Proof. | 0(a+A)~0(a) \2s(\ \ er*"W>>*)~\ \ dFj -^f |sin7r(h3a;)|d^2 and by Schwarz's inequality this is ^f {&m7r(h,x))2d$.{ dF = 2||F||f (l-cos27r(h,a;))dF, J Eh J Eh J Eh as claimed. Theorem 3.2.3. In order that a continuous function 0(%) be presentable in the form (3.2.L4), withF € V+, it is necessary and sufficient that it be positive definite in the sense that we have 2 ^"^PvTa^O, (3.2.17) #,a=i for any finitely many points a1,a2,...icxF in Ek, and any complex numbers pv..., pN. Proof. For an integral (3.2.14) the left side in (3.2.17) has the value I N |2 %pP6-*"**.4\ dF(x), which is indeed > 0. Conversely, if we start from (3.2T7), then using it for two points (oc, 0) and arbitrary complex numbers p, cr, it follows that | 0(a) \ <; 0(0), so that 0(a) is bounded, and this being so, the 'discrete5 condition (3.2.17) implies its 'integrated' counterpart if J Eh J J ftarfl) p(a) p(fi) dvadv# £ 0 (3.2.18) Eh
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 59 for any p(oc) e L-^E^. For fixed x and e > 0 we put p(oc)=e~2e J a |2 e27ri(a> x\ so that J |e-2^«>2+i^i2)^(a--^)e2^^^^)dvad^>0, and if we make the change of variables (in vectors) then this implies J Ek\J Bh J This means that the function <fie(y) — e~e]yl2(fi(y)> which for e>0 belongs to Lv has a non-negative (anti)-transform, and theorem 2.2.1 implies that <fie(&) has a representation (3.2.14). Letting e^>Q, this then also applies to {5(a) itself by theorem 3.2.1. The proof just completed also implies the following statement: Theorem 3.2.4. If a bounded measurable function <j)(a) satisfies condition (3.2.18), then it differs from a continuous positive definite function on a set of measure zero. 3.3. Poisson transforms Definition 3.3.L Given a fixed Poisson character Q(oc;x) we call a function ^(^y) a Poisson transform if it is continuous and can be represented by an integral ^(ay)=f Q(*i;Xi)dF(zs)> -WISO, (3.8.1) where E'% is the set 0 < | x \ < oo in Ek, that is, the set which arises from Eh by deletion of the origin. (For a description of the integral see section 4.1.) Theorem 3.3.1. (i) If an integral (3.3.1) is finite on an a-set of positive measure then we have /» dF(x) <oo (3.3.2) for some, and hence every, rQ > 0. (ii) If the integral (3.3.1) is bounded in \fi\£l,or only if f f(fi)d(fi) <ao, (3.3.3) J\fi\£l then f q(xj)dF(x)<oo3 (3.3.4)
60 CLOSURE PROPERTIES OF FOURIER TRANSFORMS or what is equivalent with it, [ f q(x) dF{x) + f dF(x) < oo. (3.3.5) J s'k J \x\>l whenS'k={0<\x\^l}. (iii) Conversely, if (3.3.5) holds then ijf(a,) is finite and continuous and thus a Poisson transform. Proof. If (3.3.1) is finite on a set of positive measure then there is an N>0 and a Borel set A with 0<v(A)<oo such that ifr(<x)^N for (xeA. Now, by definition 3.1.1 the function X(x) = <oA(fi) xifil a?) dvfi= x(fc x) <h>A J Etc J A tends to 0 as | x | ~> oo, and thus we have Nv(A) £ J^ftf)&>,=J^ (JjiM x)dv#)dFW =£,[»W-X(*)]dvx J !aj[^r0 J |as|^r0 for r0 sufficiently large,which proves part (i). Part (ii) follows obviously from . . tfifi)dvfi=\ q(x)dF(x) (3.3.6) J\j$\£i J si by part (ii) of definition 3.1.2. Finally, if in E'h we introduce the set function 0(A) = q(x) dF(x)9 then (3.3.4) means that 0(E'^ < oo, and, on the other hand, we can write ^(yff)— P(J5; x)dG(x), and part (iii) follows from the properties stipulated for P(ft\ x) in definition 3.1.2. We are now going to make statements on (3.3.1) in case F(A) is zero either inO<|a?|<lorin|a;|>l. Theorem 3.3.2. The Poisson transforms of the form ^2(a3)=f Q(oc;x)dF(x) (3.3.7) are a P-chsed set. Proof. If a sequence lft(a)=f Q(oc;x)dFn(x) (3.3.8) |a>|S.l
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 6l is P-convergent towards a function rfr(a), then by (3.3.6) we have Fn(Ek)^M, and since it suffices to show that the limit of a subsequence of (3.3.8) is of the form (3.3.7) we may immediately add the assumption that Fn is weakly convergent towards some F(A) and that the numbers yn—Fn(Ek) are convergent towards some number y. Thus the functions Yn-fi(*)=\ x(oc;x)dFn(x) JI«IS1 are a P-convergent sequence, and by theorem 3.2.1 we therefore have y-^(a)=f X(^x)dF(x), J | x I SI and Fn(Ek)~^F(Ek)=y, which proves the theorem. Definition 3.3.2. A set {^(a)} will be called P-finite if every sequence {^n(a)} in it for which sup liMfl.SJfn (3-3.9) l^isi is equi-uniformly continuous in | a | ^ oc0 for every ocQ > 0. Now, equality (3.3.6) and our assumptions on P(a; x) easily imply as follows: Theorem 3.3.3. The set ofPoisson transforms of the form ^i(a)=f Q(a;x)dF(x) (8.3.10) J si is P-finite, and for a sequence !#(<*)= f Q(a;x)dFn(x) (8.3.11) J a* the relation (3.3.9) is equivalent with q(x)dFn(x)SM2. (3.8.12) L Definition 3.3.3. We call a function in i?fc: (%) pseudo-Gaussian, and we denote it by B(cc), if it is a P-limit of a sequence of Poisson transforms p #K«H Q(*;x)dFn(x) (3.8.18) with rn->0. Any function (3.3.13) is a special case of (3.3.10), and it is not hard to establish the following theorem: Theorem 3.3.4. The set of pseudo-Gaussian functions {E(a)} is P-finite and P-closed.
6Z CLOSURE PROPERTIES OP FOURIER TRANSFORMS But now we are coming to a key theorem. Theorem 3.3.5. The set of functions 5(a)+ ^(a), (3.3.14) where B(<x) is pseudo-Gaussian and xjr1^) is of the form (3.3.10) is the (smallest) P-closure of the set of functions {^(oc)}. Proof. It is quite easy to argue that the P-closure of (3.3.10) must include all functions (3.3.14), but we must show conversely that if a sequence (3.3.11) is P-convergent towards a function ^(a), then there exists a function (3.3.10) and a function R(oc) such that ftlioc) =*R(a) + ^(a). (3.315) Now, the P-convergence of (3.3.11) implies (3.3.12), and, after passing to a subsequence if necessary, we may assume that there is a weak limit F(x) in the sense that, again, q(x) dF(x) <* M2 and I c(x)dFn{x)-*\ c(x)dF(x), (3.3.16) J si Js'k for every continuous c(x) in S'k which vanishes in some neighborhood of the origin. Now, the function in r, L dF(x), 0<r<l, is monotonely increasing as r decreases. Therefore, it is continuous for all but countably many values r, and if we keep those excluded then (3.3.16) implies f Q(oi;x)dFn(x)->( Q(a;x)dF(x). (3.3.17) We now form ^x(a) with this F(x), and our theorem will be proven if we show that the difference fi(a)«^(a)-^(a) is pseudo-Gaussian. Tor each p = 1,2,..., we can pick a value rv < \jp such that Q(a;z)dF(x)g- for\<z\^p. (3.3.18) 0<\x\grp V But, for fixed r^ we can find an index n^ such that .1 J. I ***(«)-#L(«) I + f «(«,*)d(F -F) <- (3.3.19) P
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 63 p for I a I £p, and from (3.3.18) and (3.3.19) we can put together the estimate 1 /• \B(a)-\ Q(a;x)dFnp I JO<\x\^rp y for I a I <p, which shows indeed that E(a) is a P-limit of a sequence (3.8.13). Theorem 3.3.6. For the set ofPoisson transforms {ft(oc)} the P-closure is the set of all functions ^a) + ^a^ (3.3.20) Proof. Obviously all functions (3.3.20) are in the P-closure. Take now conversely a P-convergent sequence ofPoisson transforms i]fn(oc) and put yjrjct) = ^J(a) + V^(a)> where ^i(a)=f Q(*,x)dFn(x), ^(*)=f Q(oc;x)dFn(x). J S'k J.a|>l By what we already know, after passing to a subsequence, we may assume that the sequence {^(a)} is P-convergent, its P-limit being of the form r 22(a) + Q(a; x) dF\x). (3.3.21) J 4 The differences ^(a) = ^(a) — V^(a) are therefore also P-convergent, the limit being of the form I Q(ot;x)dF*(x). (3.3.22) l«lisi But the sum of functions (3.3.21) and (3.3.22) is of the form (3.3.20), q.e.d. Theorem 3.3.7. (i). If Q(a-,x) is as in theorem 3.1.5, Q(a; x)=zB(oc1x1 + ...+ccJcxk), (3.3.23) then the set of pseudo-Gaussian functions {22(a)} is the P-closure of the finite linear combinations with positive coefficients of the 'monomials9, |al£i + ... + a*&|m (3-3.24) with ZI + ... +£1=1, say. (ii) For m = 2, {22(a)} are all non-negative quadratic forms Je S ^a^a^O. 3>,a—l (iii) In the representation of a closure element iJr0(ot) as a sum R(a) + i/r(a) the two addends 22(a), ^(a) are uniquely determined. Proof. Ad (i). In general, if we can put Q(oc; x) = Ql(cc; x) + Q2(a; x), ta 2teJ8_o
64 OLOSUKE PROPERTIES OE EOURIER TRANSFORMS uniformly in | oc \ ^oc0, for every a0> 0, then in the construction of an E(a) as a limit of a sequence (3.3.13) we may replace each element of the sequence by - (P(a;x)dFn(z), J 0<| x\^rn without altering the fact that it is P-convergent, and towards M(oc). Now, in our present case this applies with Q1(oc;x) = \ (a, x) \m, but every integral r \oc1x1+...+akxk\mdFn(x), (3.3.25) J0<|sc[^rn is a P-limit of approximating Riemann expressions each of which can be written as a linear combination, with positive coefficients, of terms (3.3.24). Conversely, the pseudo-Gaussian functions are a (semi)-vector field with positive coefficients, but each term (3.3.24) is a pseudo-Gaussian, since it is a P-limit of expressions (3.3.25), all having identical values, and corresponding to set functions Fn(A) having the value 1/r™ at the point rn£1? ...,rn£&5 and being zero everywhere else. Ad (ii). For m=2 these functions are precisely all quadratic forms ^0. Ad (iii). If we envisage two representations ^i(a) + ^i(a) = -B8(a) + ^a(a), then on putting R-^ct)—J?2(a) = ^2(a) — ^(a) we obtain an identity .8(0!,...,^)= [ B{alz1 + ...+akzk)dF(z)t (3.3.26) J K in which -R(t%) is a homogeneous function of weight m, and F(A)=F1(A)-Fi(A) is a set function of mixed sign for which we have L \x\m\ dF(x) I + I dF(x) I < 00. (3.3.27) Jl»l>i If we divide (3.3.26) by | a |m we obtain \\a\' '|al/~Jo<|!C|s*|a|m-|^|ml ' ' Wl \a\ J\x\>$
CLOSURE PROPERTIES 03? FOURIER TRANSFORMS 65 Now, the quotient B{(oc, x))j \ oc \m | x |m is bounded in [a, x), and therefore the first integral is ^ e for 8^ 8(e) for all oc. But, for fixed 8, since B((oc,x)) is bounded, the second integral tends to 0 as | oc | ->oo, and therefore the left side tends to 0 as | oc | ~> 00. But the left side depends only on the ratios ax: a2:...: ock, and must be identically 0 therefore. 3.4. Infinitely subdivisible processes We take the same Poisson character Q(oc; x) as before and the function q(x) obtained from it, and in addition to that a real-valued function Q(oc; x), not necessarily a character, which is continuous in (a, x)9 and bounded in x for | oc | <; a0, any oc0 > 0, and for which we have Km %lLo, (3.4.1) uniformly in | a | g a0. Theorem 3.4.1. The 'transform1 ^(a) + *#(a)=f (Q(a;^) + iQ(a;^))dF(a;) (3.4.2) exists for F(A)^0, \ q{x)dF(x)<oo, (3.4.3) M and for the set of functions (3.4.2) the F-closure is the set of functions B(oc) + ir(oc) + ift(a), (3.4.4) where {B(cc)} are the same pseudo-Gaussian functions as before. Proof. In other words, the imaginary parts \jr(oc) do not create pseudo-Gaussian functions of their own, and this, as we will see, is simply due to assumption (3.4.1), which, as can be readily seen, first of all secures the existence and continuity of the imaginary parts $*(a) whenever F(A) satisfies the requirement (3.4.3), although this requirement was arrived at to suit the needs of the real parts ifr(a)f originally. Now, if we are given a P-convergent sequence frn+iftn* 'fcnen we again dissect it into the parts J si lft+*#t=f (Q+iQ)dFn, (3.4.6) J\x\>l and the sequence (3.4.5) is P-finite again. Therefore, after passing to
66 CLOSURE PROPERTIES OF FOURIER TRANSFORMS a subsequence, we may assume that in S'k the sequence Fn is weakly convergent, and we now obtain ^+*&^i?(a)+f (Q+iQ)dF*(A), because formula (3.1.16) implies that we have [ QdFnX[QdF\ on account of (3.4.1). This being so, the sequence (3.4.6) is now likewise P-convergent, and so are therefore the real parts ^*. Therefore, the sequence Fn(A) if envisaged on the closed set | x | ;> 1 is Bernoulli convergent there, towards a function F2(A), and thus tt+i$*n+\ (Q + iQ)dF*(A), J\afel which completes the proof. In the classical case we are given, to start with, the combination Q + iQ' = 1 - #-**%*. x)=2(sin tt((x, x))*+i sin 2tt(oc, x) in which, however, Q' does not satisfy requirement (3.4.1). But at the expense of adding a linear term in the a's outside the integral, this can be corrected by putting Q(oc; x) = 2smn(<x,x) — 2n(a,\(%)), (3.4.7) where in (ociA) = cclX1(x) + ...+ockAk(x) we may take for X3(x) any continuous function in Ek, for which i()~\0(l) as \x\ ^oo holds. In the theory of probability it has become customary to put Xj(x) = x3-(l + \x\2)"1, and in the theory of generalized Fourier integrals another normalization is being used but no choice has a stochastic preference over any other, and in order to emphasize this we will leave the functions X3-(x) unspecified, although we assume that they have been chosen fixed. We are introducing the linear form ^(«)=f;<V*, (SA8) for any real cP (B for Bernoulli), the general real symmetric form ^(a)= S c^a^^O, (3.4.9)
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 67 (67 for Gauss) and the general function ijr^oc) = (1 -e-2^*>-27ri(a, X(x)) dF(x) (3.4.10) for any F(A)^0, f \x\*dF(x)+{ dF(x)<oo (3.4.11) J S'k J|0!|>1 (P for Poisson) and we claim as follows: Theorem 3.4.2. If we start out with the set of functions ^°(a) = (1 - e-27Ti(a,s)) ffi(x) (3.4.12) J Eh JF(A)£0, %)<«), then its P-closure is the set of functions ir{a) = ^(a) + ^(a) + ^p(a), (3.4.13) with all Bernoulli, Gauss and Poisson addends used. Also.wehave ijrp(a) — o(\ ol |2) as |a|->oo, (3A14) and the quantities c^, cm, F(A) in the decomposition (3,4,13) are uniquely determined. Proof. If a sequence ^(a)=f (l-e-2^«^))dF^) (3.4.15) JJSk is P-convergent then so are its real parts by themselves. Therefore Js'k J\ x\>l dFn£M, and if we put each function (3.4.15) in the form *E<V**+f {l-e^i^^-2iri(<x,X))dF(x), (3.4.16) then it follows that the imaginary parts of the integral and the linear parts outside are both P-finite. Therefore, in a subsequence, the linear parts and the imaginary parts will be P-convergent, and by using our previous results it is not hard to see that the P-limits of (3.4.15) will be the functions (3.4.13) with all possible fa{&), and ^p(a) and some i/rB(oc). But actually, we can obtain all possible linear parts i/rB(oc), since any term iocf is the P-limit of l/n(l — e~"*ai/n) as n~>co. 5-2
68 CLOSURE PROPERTIES OP FOURIER TRANSFORMS Next we note that for | x \ < 1, | a | ^ 1 we have \*\*.\x\* = and (3.4.14) follows, by putting (Q+iQ)dF(x) and letting | a | ~>oo for 8 small though fixed. As regards uniqueness we note that if we consider the real part only, then the addend i/rG(oc) is uniquely determined by part (iii) of theorem 3.3.7, and we will have to show that if a function ^(a) is of the form (3.4.16) and F(A) satisfies the general assumption (3.4.11) then F(A) is uniquely determined by this. With any vector we form h = (hx, ...,hft) 1 Jc TS[^(ai>...Ja,.+h/,...J^)-2^(a1,...,ai,...,afc) *3=1 + f(a1)..,ar)iir.,afc)] = I ^e-2ni(a,x)( 2 (Bm7rhjXj)2\dF(x) = e-%7Ti(«>xHFh{x), where Fh(A) = f ( S (sinnh^A dFfa), and by (3.4.11) we have Fh(A)^0, Fn{Ek)«x>. If now we extend Fh(A) from Ek to Ek by assigning the value 0 to the origin, then by theorem 2.L4, Fh(A) is determined uniquely in Ek and hence in E'k. Now, for any finite number of vectors hP = (h£, ...,h£), p = 1,..., r, this then determines uniquely the sum £ J?™(.4) = f (£ (sin^,)2V(z), />~1 J^\/)=-l / but for an appropriate choice of the vectors h? the integrand on the right side will be 4=0 for x4=0, and thus F(A) itself is uniquely determined in E'k, as claimed.
CLOSURE PROPERTIES OP FOURIER TRANSFORMS 69 Definition 3.4.L A family of characteristic functions $(t;<xj)=\ e~27Ti^>*>dxF(t;x) (3.417) F(t;A)^0, F{t;Ek) = l9 (3.4.18) 0 < t < oo, will be called an (infinitely) subdivisible process (or law) if there is a function fifa) such that we have 00; a,) = 6-*^) (3.4.19) inallt>0, <xeEJc. Theorem 3.4.3. A family (3.4.17) is of the form (3.4.19) if and only if we have ,, v ,, ,, v ._ . .„, J 0(r;a)0(s;a) = 0(r + s;a) (3.4.20) p /orr>0,s>Oand 0(t;a)~>l as HO; (3.4.21) or, what is equivalent with it, if we have F(r;A)*F(s;A) = F(r+s;A), that is F(r;A-y)dyF(s; y)=F(r+s; A)3 J En and if, furthermore, for t^Q, F(t;A) is Bernoulli convergent to the 'identity' distribution J which is concentrated all at #=0, an<d which can also be characterized by J*G=G,for GeV (or V+). Proof. It is obvious that (3.4.19) does all this. Conversely, if <f>(t; a) is as in the theorem, then, on account of <f>(r+s; cc)-~<f>(r; a) = 0(r; cc)(<f>(s; a)-l), there is an open set 0 < t < t0(cc) in which we can define a continuous function log^(t; a) which satisfies the functional equation log0(r+s; a)=log0(r; oc) -flog0(s; oc) for 0 < r, s < %tQ(oc). By known properties of this functional equation we obtain indeed log$(t; oc) = — ti/r(oc) for all t, a as claimed. Theorem 3.4.4. A function ^(t%) describes a subdivisible law if and only if it belongs to the closure class (3.4.13) that is, if and only if, it is of the form iS^^+SvA+ f (I-e-*"^~2m(cz,\))dF(x), p p,q J %i where cp, cm, F(A) are as described before.
7o CLOSURE PROPERTIES OF FOURIER TRANSFORMS Proof. (3.4.21) implies p 1# l-6tu\a5T so that ^(flfy) belongs to the P-closure. Conversely, if ^(%) is a positive definite function, then oo und>(ot)n o nl is again positive definite, and thus 0(0) — 0(a) describes a subdivisible law. But the P-limit of subdivisible laws is again subdivisible, and hence the theorem. 3.5. Absolute moments /•oo For any #*)= e-Z7TiaxdF(x), FeV.we have 0(a+h)~0(a)_ f00 e-*****-! _ f00 * J -00 h / _ oo hx e-%7Ii**xdF{x), and since l/hx (e~27rihx— 1) is in absolute value ^ 1 and tends to 1 as h ~> 0, we see that if we have ;. \x\\dF{x)\<oo then 0(a) has a continuous first derivative which can be formed under the integral. More generally, for any m ^ 1 and any h ^ 1, if we have I \x\^m\dF(xj)\<(X)i JJSk then the Fourier transform 0-F(%) is in C(2m) and the partial derivatives of order ^ 2m can be formed under the integral; so that in particular, for k = 1 we have d2™0(O) (-l)-J' (-l)H a2™dF(#). (3.5.1) da2wl Now, it is notable that for F(A) ^ 0 these assertions can be significantly inverted and it will suffice to describe the one-dimensional situation only. If we put 2m /2m\ A2w0(a)=2 (-1)' ^ 0((m~r)a), so that in particular A20(a) = 0(a)-20(O) + 0(-a),
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 71 then we have JL y2m a* A2m^(a) = (2i)^r (E^)*m7**F(x)9 (3.5.2) J -_oo \ CCX J and if now we replace the integral by , for A fixed, let a->0 and then A ~> oo, then, for Fe V+, we obtain r. ^dF(^)<Hm-^^-, whether the quantities are finite or +00. Thus, if the function 0(a) has what is called a generalized symmetric derivative of order 2?n at the origin, or only if the corresponding difference quotient is bounded, then the (2m)th moment is bounded, and the function 0(a) is in C7(2m) (3.5.1) holds. Thus, for a characteristic function other than 0(a) ==1 the (2m)th derivative at the origin cannot be zero for m ^ 1, and this is the quickest way of verifying that for p > 2 the function e-l<*iy is not a characteristic function. All this is preliminary, and it suggests that for 0<p<2m the relation roo \x\vdF(z)<oo J -00 ought to be in some manner 'equivalentJ with A2TO0(a) = O(|a|*>), and we are going to show that such is indeed the case, although not literally so. We take in 0 < a g 1 a measurable function A(a) ^ 0 for which n a2wA(a)da<oo, /: /o and if we introduce the Poisson transform ji^x) = (sin 7T0cxfm A(a) da, multiply both sides of (3.5.2) with a2mA(a) and integrate, we obtain the following conclusion: Theorem 3.5.1. We have I A2m0(#) I A(a)da<oo /:
JZ CL0SUBE PBOPERTIES OF FOUBIEB TBANSFORMS if and only if fo(x) dF(x) < oo, J —00 or what (due to jlc0(x) ^ M0 in | x | ^ X0) is the same if and only if + /i0(x)dF(x)<cc. J -oo JXo From this we will conclude as follows: Theobem 3.5.2. If a measurable function A(a) ^ 0 in 0 < a £ 1 is swch that the function ji{x)=x2m\ VwA(a)da (3,5.3) is finite, and if X(a)da=0(ft(x))3 as #->oo, (3.5.4) Jl/as then we have + jjl(x) dF(x) < oo (3.5.5) if and only if | A2m <j)(a) | A(a) da < oo. Proof, If for ^ ^ 2 we put /£0(#) =/^i(x) +ft2(%), where fi1(x)=\ (smnocx)2m A(a) da, /£2(#) = (sin7ra#)2?nA(a)da, Jo J l/as then we have /£2(#)^ A(a)da, jn1(x)^M1/i(x) Jl/as nix and /^(&)^jtfaa,2w a2mA(a)da = M2^(a;)3 so that (3.5.4) implies Jf8/t(o;) g/^(a)+/ft2(a?) <Jf4/t(a?), and our theorem is a consequence of the preceding one. In particular, if we put A(a)=a~a""1L(l/a), where q<2m and L(x) in X0 ^ x < oo is a so-called function of slow growth, and is monotonely decreasing, then by the theory of such functions, ji(x) is bounded above and below by x?L(x), and condition (3.5.4) is fulfilled, so that the following specific conclusion ensues: Theobem 3.5.3. if L(x) in X0£#<oo is a monotonely decreasing function of slow growth, then, for q < 2m, relation f:a jLc(x)dF<oo
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 73 is equivalent with relation i ^"""lPuckco. o ^+1 V*, In particular, if for 0 <p < 2 we introduce fP(x) = 2 e-a2? cos 2ttx doc (probability density for a stable symmetric law), then we have f<oo I f1>(x)xpL(x)dx x »v' •• l=oo- depending on whether -LI-Ida] , J0a \a/ t = oo whenever L(x) is monotonely decreasing and of slow growth; and this follows from the fact that we now have -Aa#x)=2(l-e-«*)~2a» for small oc. Thus, for instance, f^xvQogxyL-tdx /; is finite for e>0 and infinite for e=0, and similarly for the entire logarithmic scale. 3.6. Locally compact Abelian groups If we replace the pair of dual spaces {Ek: {x),Ek: (a)} by the pair {Tk, Mk}} and if for Fe V+(Tk) we introduce the coefficients <j>(m)=[ e-27Tiim>x)dF(x)} (3.6.1) J Tk then an analogue to theorem 3.2.4 would be vacuous because on Mk every function is continuous and the analogue to theorem 3.2.3 is as follows: Theorem 3.6.1, A sequence {^(m)} is of the form (3.6.1) if and only if we have S <j>{m-n)p{m)p{n)^0 (3.6.2) (m\ (n) for (m) and (n) ranging over any finite set in Mk.
74 CLOSURE PROPERTIES OP POURIER TRANSFORMS Proof. As previously, (3.6.2) implies (m), («) and on putting m~n+p, the sum is V e-4e|«+*+iD|2e-i3|p|a^(-p)e-2jri(j»,a!)# On), (p) Now, by adding over a suitable number of points t, this will be Ce.fe(x), where Ce is a positive number and (») Thus we have/e(cr) ^0, and on letting e->0we obtain the conclusion of the theorem by the use of an analogue to the closure theorem 3.2.1, easily provable. Next, since Mh is discrete, P-convergence on it is ordinary convergence point-by-point, and the following theorem can be obtained by suitable adaptation of previous syllogisms: Theorem 3.6.2. IfafamilyF(t; A) e F+(Tfc) satisfies the assumptions of theorem 3.4.3, but on Tk, then for the Fourier coefficients 0(t;m)=j er^m>*HxF(t\x) (3.6.3) J Tic we again have <f>(t; m)^e-t^m\ (3.6.4) where the function ijr{m) on Mk is in the P-closure of the functions 7jr°(m) = (1 _6-a**<«»»)) dF(x)9 (3.6.5) JTjc Fe V+{Th)\ and conversely any element of the P-closure gives rise to such a family F(t; A). The closure elements are again of the form f{m) = i/rB(m) 4- fG{m) + fp(m), (3.6.6) where i/rB(m) is any form i 2 cpm99 ftG(m) is any P-limit of finite sums of expressions {a1m1 + ...+aJcmk)*, and ii*(m)=\ {l-e^^^^2iZrrivsm7Tx^dF{x), (3.6.7) where T'h*=*Th-(% and F(A) ^Ois defined on T'k and is such that 2 (sin 7Txvf dF{xj) < oo.
CLOSURE PROPERTIES OF FOURIER TRANSFORMS Also, for given \jr(m) everything in the decomposition (3.6.4) is uniquely determined. But now we consider the dual pair {Mk, Tk}, and in this case any element of V+(Mk) is a sequence {/(m)} for which /(m)£0, S/(m)<oo, (3.6.8) On) and its transform is 0(aj) = 2e-2w*fow)/(w). (3.6.9) <«) If we view this transform not as a function on Tk but as a periodic function on Ek, then it is a particular case of our previous transforms in Ek, for functions F(A) which have nonzero values only at the lattice points, and thus theorems 3.2.3 and 3.2.4 apply readily. We also note that P-convergence on Tk is uniform convergence on Tk9 due to its compactness. However, with regard to infinitely subdivisible laws a new situation arises. If we define them suitably, then the characteristic functions ^. a \ = ^ &-*"**>»>/(*; m) On) have the values e-W*fi, and a function ^(a5) is of this kind if and only if it belongs to the P-closure of functions representable in the form 2' (1 -e-27^ m>)/(m), (3.610) On) with (3.6.8) holding. Now, due to the fact that the origin (m) = (0) is ' isolatedJ in the topology of Mk it can be seen, by going over previous arguments, that the class of functions (3.6.10) is already P-closed as is, so that no Gaussian distribution emerges, at least not as a fimit of Poisson functions, and incidentally, no corrective term of Bernoulli term is needed either. Actually, if we view our function (3.6.9) as a periodic function on Ek then what we have just stated implies that if an exponent (3.4.13) of a general subdivisible law in Ek happens to be multiperiodic then it is of the pristine form (3.6.10) perforce, and this assertion could have been so proven directly, But the reasoning straight on {Mk, Tk} has the advantage that it can be generalized from Mh to any discrete Abelian group, the result being as follows. Let Q: (x) be any locally compact Abelian group and G: (a) its dual group, and #(a; x) the character from G to 67. If with each F(A) eV+(G) we associate the function #*)=(" xte*W(*)> (3-6-n)
CLOSURE PROPERTIES OF FOURIER TRANSFORMS then we take it as known frome abstract? harmonic analysis that these functions {^(a)} are continuous positive-definite and P-closed. Also, if we introduce an infinitely decomposable process as before, then the exponents {^(a)} are again the P-closure of the set of functions ^>(a)=f (l-x(a;x))dF{z)9 FeV+(G), (3.6.12) but the actual analytical representation of the closure class is not known to us, although various parts of our previous statements could be upheld by making suitable assumptions on the character function X(oc; x) axiomatically. However, the following statement goes easily through, without additional assumptions being needed. Theorem 3.6.3. If the Abelian group G is discrete, that is, equivalently, if § is compact then the class (3.6.12) is P-closed as is. In other words, if the values of random variables are the elements of a discrete commutative group then there are no (joint) Gaussian distributions for them possible. 3.7. Random variables Definition 3.7.1. A {Lebesgue) measure space is an ensemble {R; Sl\ v}, (3.7 J) in which R: (x) is a general point set, J*: (B) is a <r-field of sets in R (with R itself being an element) and v~v(B) is a cr-additive measure on a, Ogi?(jB)^+oo. We will caff (3.7.1) finite if v(R)<oo and <r-finite if there is in SI a sequence B1CB2CBdC ...~^R such that v(Bn)<oo, n,= l,2,3,.... If v(R) = l, then, in stochastic contexts, (3.7.1) is a probability space, and if so viewed we will also denote it by {n;^;P}, (3.7.2) with Q = Q: (o)), ^=^: (8), P=P(S), P(Q) = 1. Definition 3.7.2. We caU (3.71) topological if R is endowed with a HausdorfT topology and if {@\ v} are such that for each Be& and each rj > 0 there is a compact set Ov in S such that Ov C B and v(Cv)>v(B)~7j. (3.7.3) We call it strictly topological if every Borel set (relative to the given topology) belongs to ^.
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 77 The following fact will be taken as known. Theorem 3.7.1. IfR is the EuclideanEkJor anyk^l, and@*=$tfk is the cr-field of ordinary Borel sets A, then {Ek; */Jc; v} (3.7.4) is strictly topological for any cr-additive v(A) on sfk. If v(A) is a probability measure then in the Euclidean case we may also denote it by F(A) or F(:ty) as before, and the reason for then calling it a (joint) distribution function is as follows. Take any two spaces R':(x') and R: (x) and any mapping x=f(x') (3.7.5) from R' to all of R and for any BCR introduce its preimage B'~f-\B). (3.7.6) If now we are given a cr-field £8 in R then (3.7.6.) gives rise to a cr-field ^o=f,-i(^) in R' which we will call the generated field, and if a field £$' in R' was given to start with and ^° C £%' then we will call (3.7.5) a consistent mapping of {R\ 38'} into {R, SS). Furthermore, if there is given a c-measure v'(Br) on @t', then v(B)^v'{f~\B)) (3.7.7) defines a suchlike measure on ^, with v(R) = v'(R')t which we will call a generated measure, and if this measure v(B) was so defined originally on @i then we will call (3.7.5) a consistent mapping from (R';&';vf} to {R; @\ v}. Take in particular for {R'; £$'\ v'} the probability space (3.7.2) and on it h real-valued Baire functions f^co), ...,fk(o)), and consider the mapping from Q: (o)) to Ek: (o^). Since the functions are Baire functions this is a consistent mapping from {Q; £?} to {Ek; &?k}, and the probability measure P(S) then generates a probability measure F(A)=FV(A) on stfh. Definition 3.7.3. The probability measure F(A) thus generated is the (joint) distribution function of the Jc functions (3.7.8), the latter functions being now called random variables in this context.
y8 CLOSURE PROPERTIES OF FOURIER TRANSFORMS From Lebesgue theory we obtain the following fact: Theorem 3.7.2. For any bounded Baire function b(xv...,xk) in Fk we have I 6(/iH,...,A(w))iPH=f b(xv...3x7c)dFP(x^} (3.7.9) and, more generally, for an unbounded Baire function if one of the two integrals exists then so does the other and equality holds. Definition 3.7.4. The integral on the right-hand side in (3.7.9) is called the expected value of b(xl3 ...,xk) and is also denoted by then. E{b(xv...,xk)} or E{b(fi,...M (3.7.10) It suits stochastic purposes admirably that in this last notation the awareness of the original probability space (3.7.2) is kept in abeyance and that the substitute probability space (3.7.11) {Ek; j/j; F(A)} (3.7.11) is being put forward instead. For b(Xj)=6-*ff*(«.*), oceEk> the expected value ^O-^e-^^H I e-toHttih+'-'+akftidPia)) (3.7.12) is the characteristic function 0(o,)=f e-^^dF^x), (3.7.13) J Etc but in the notation (3.7.12) it pertains to the random vector (3.7.8) rather than the set function F(x). Definition 3.7.5. A sequence of random vectors *Z=ft(o>)9...9a%=fi((o), »-a,2,... (3.7.14) is called convergent in probability (or in measure) to the vector (3.7.8) if for every bounded continuous function we have E{b(xl ...3<%)}-+Eq>(xl9 ...,<**)}. (3.7.15) For the corresponding characteristic functions we obviously obtain 0n(a,)-*0te), and for the distribution functions we have F*{A)-+F{A)\ but this last relation does not conversely imply that the vectors (3.7.14) are convergent in probability, since distribution functions of different random vectors may be identical. What seemingly is true is that there
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 79 exists some (other) probability space and on it a sequence of vectors having the same distribution functions and being convergent in probability. But this converse will not be of consequence to us (although closely related statements will) and we need not dwell on it. Next, if we are given two probability spaces {Q1: (at); &*; P1}, {Q2: (w2); ^2; P2}, (3.7.16) and if (3.7.2) is their product in which a) = [y,u2], Q = Q}xQ\ S^^S^xS?*, P(S1xS^) ~PW) PHS2) then for any random variables J v " a^ =/>!),.•., **-/>*), yi=gi(w8),..., yi^gim (3.7.17) and for any two bounded Baire functions <j)(xx, ...,#&), i{r(yls ...,2/z) we have E{^xK)f(yA)}=E{4,(xK)}.E{ir(yx)} (3.7.18) in the sense that f WKW(0x)dP(<o)=\ <f>(fK)dPHa)i).\ ir{gx)dP*{o>\ and this gives rise to the following definition: Definition 3.7.6. On any probability space two sets of random variables {xK}, {yA} are (stochastically) independent if (3.7.18) holds; and more generally if a family of finite or infinite sets of random variables {xT} is given, where r signifies membership in the family, then we call the family independent if for any upper indices rlt ...,Tn, and after choosing those, for any finite subsets x{>>,....s&Jy, v= 1,...,n, we have ( n \ n E\ n^v..,*© = n w*iv..,*y} (3.7.19) U-l J v-1 for any bounded Baire functions <f>v(x). Given two sets of variables {xK},{yx}, if we introduce the characteristic function $((%, ...,ock) of the x\ the function ^(/?x,...,/??) of the y's and the function #(yl5..., yki yk+v ..., yk+l) of the (x, y)'s, then, in the case of independence, (3.7.18) implies ^K,...,afc)^1,...,A)=^K,...,a,,/?1,...,A)3 (3.7.20) and it is possible to deduce from theorem 2.L4 that the converse likewise holds. For l=h, if we denote by p(Sls..., Sk) the characteristic function of the random variables z1—x1+y1,...) zk=xk+yki then theorem 2.1.1 implies 0(aij ...j0ck)jjr(al3..., a*) «/>(*!, ....a*), (3.7.21) and this relation is of fundamental importance indeed.
80 CLOSUBB PROPERTIES OF FOURIER TRANSFORMS Also if we start from functions 0(o,)=f e-^^dF^x^iP^l ertoWrfiFHy,), and if in the product J Elk we put fa — a>j, then we obtain <pfa) fioij) = f er^«> *+*> dF1^) £Fa(&), and our previous statement admits therefore of the following converse. If three characteristic functions are as in (3.7.21), then they pertain to three random vectors x,y,x+yon some suitable probability space, and in fact on a 2&-dimensional Euclidean space, as it happens. Of this converse there are various generalizations to cases in which any (infinite) number of characteristic functions are involved, and, for instance, if we are given a set {e-W*fi} underlying a subdivisible process then on a suitable probability space there are corresponding random vectors x—x(t) for which x(r) +x{s)=x(r+s) is satisfied. And furthermore, if we have a decomposition ^(%) = ijra (fy) + i/rp (%) (compare (3.4.13)), then we may secure a corresponding decomposition of the random vectors, thus, x{t)—xa(t)+xp(t). All this is implied in a well-known theorem, but we will have a rather more general proposition of our own in which it will be contained. 3.8. General positivity Definition 3.8.1. We denote by i^ a vector space with real coefficients which is partially ordered in the following sense. There is denned a relation X>Y (or equivalently Y^X) for some pairs of elements X, Y of "T such that the following conditions are satisfied: (a) XZX, (b) X^ 7, Y^X imply X- Y, (c) X^ Y} Y^Z imply X>Z, (d) X^ Y implies X+Z^ Y+Z for any Z, (e) X^ Y implies aX^aY for any positive number a, and (/) every monotone non- decreasing sequence which is hounded from above has a least upper bound. That is, X^X^ ... ^Xn^... <; Y implies the existence of an element X'=supXn such that XngX'(n=l92,...) and that Xn£X'(n=l,2,...) implies X'^X".
CLOSURE PROPERTIES OF FOURIER TRANSFORMS 8l We further denote by if a vector space with complex coefficients for which to every element I of / there corresponds a unique element Z* of if such that (X*)*=X, (X+7)*=X*H-F*, (pZ)*=pX* for any complex number p, and if if denotes the subset of elements for which X*=X, then in if there is given a partial ordering as described. We emphasize that we do not require that 'V shall be a lattice in the sense that for any two elements a 'meet' and 'join5 ('sup' and 'inf') shall exist. A very significant nonlattice space if is the set of all bounded Hermitian operators in a Hilbert space of any dimension if we define the order relation H1^H% to mean that H2—H1 is positive semidefinite and if i* consist of all (non-Hermitian) bounded operators then. Theorem 3.6.1 can be generalized as follows: Theorem 3.8.1. If$(m) has values in i^ then the condition (3.6.2) is fulfilled for all finite sequences of complex numbers {p(m)} if and only if <p(m) can be represented by a suitably defined Eiemann integral (3.6.1) in which F(A) is a finitely additive interval function whose values are non-negative elements in if. It is a little harder, for general if, if, to establish the corresponding generalization of theorem 3.2.3 that a continuous function ^(%) from Ek to if can be represented by a Fourier transform 2.LI for a non- negative interval function F(A) with values in if if and only if it satisfies (3.2.17) or (3.2.18) respectively, the difiiculty being simply one of finding the right version of continuity for which to secure the statement without restricting it unduly. But for special if, if of importance the task of setting up the theorem is sometimes quite easily accomplished. For instance if if, if are operators in a finite dimensional Hilbert space then the elements are matrices, and thus 0(<fy) is a matrix {$uv(a>j)}, u, v— 1,..., M and it is then not at all restrictive to assume that each component ^MV(a^) is continuous in oc by itself. This was so stipulated by Cramer and he obtained the result that such a matrix function <f>(a0) is positive- definite, if and only if, component-wise, we have a representation where each FW(A) € V(Ek) and the matrix {FW(A)} is real symmetric positive semidefinite for each Borel set A. 6 BHA
82 CHAPTER 4 LAPLACE AND MELLIN TRANSFORMS 4.1. Completely monotone functions in one variable We will frequently encounter in Ek an integral h #,)*>&) (±-1.1) of the following description. B is a Borel set in Ehi ^(^) is a Baire function on B, usually continuous, and, what is important, 0(^)^0; and p(A) is denned on the Baire sets A C B, and is cr-additive and positive, p(A)^0. However, p(A) may also assume the value +oo, but if A is a compact subset of Ek, then p(A) is finite. The integral (4.LI) always 'exists' as a finite or infinite number ^ 0, but whenever we introduce it as Existing' without qualification, then we will tacitly imply that its value is a finite one. For Jc— 1, if we write down the integral rmdpit), then B is the set 0 <j t < oo, and p(A) is therefore finite for A:0^t^t0 (<oo), but might become infinite if t0 -> oo. If we represent p(A) by a mono- tonic point function p(t), then p{+oo)—p(0)=p(B) whether this is finite or not. But if we write down the integral P° <f>(t)dp(t), Jo+ then-BisO<t<oo, andp(A) is finite for a<t^b, 0<a<b<oo, but p(A) may ->oo if 6~>+oo, or a-> 0, or both. For the point function p(t) we may thus have p( + oo) -= 4- oo or p(0 +) = - oo, or both. We take the following theorem as known: Theobem 4.1.1. A function f(x) in0<x<cohas a representation /*00 e-^dpit), p(A)^0 (4.1.2)
LAPLACE AND MELLIN TRANSFORMS 83 if and only iff(x) is completely monotone in the following sense: it belongs to differentiability class C(°°> and (_1)n£-°' w=0>1'2—• (4-L3) The class of such functions will be denoted by CM. Actually, if we introduce the difference operator then it suffices to demand <-l)«AJkl>...,A]kfl/.>0 for any hx>0, ...,hn>0 without even adding continuity, but the description used in the theorem is more pertinent to our immediate context. Definition 4.1.1. We call a continuous mapping x=xjr(y) of 0 < y < 00 into 0 < x < 00 completely monotone if for every f(x) € CM, the function g(y) =f(ft(y)) is also in CM. Theorem 4.L2. If fr(y) > 0 and ifr(y) eC(00) and (-1)^^*0, n=l,2,3...., (4.1.4) then the mapping x=ijr(y) is completely monotone. Proof. We have g^O and dg/dy=df/dx, di/r/dy^O and this verifies (-l)"pL^0, n=0,l,2,... (4.1.5) for n = 0,1. Now, for n ^ 2 we obtain, by induction on n, an identity in which all C^ ;> 0 and the summation extends over p0 ^ 2, p± ^ 0,..., #^0, ^o+^)i + --*+^=:'?2'j wi^ v arbitrary. Relations (4.1.3) and (4.1.4) imply (4.1.5) as claimed. There is a converse theorem that a completely monotone mapping must satisfy (4.1,4) and we will prove it in the following version. Theorem 4.L3. Letf{x) be Cico) in 0<x<oo, let f'{x) < 0 in some interval 0 < x < x°, (4.1.7) andfom^2let /(n)(z)=0(/'(a)) as a:-*0. (4.1.8) 6-2
§4 LAPLACE AND MELLIN TRANSFORMS If x[r{y) e C(c0) and \jr(y) >0in0<y<co, and if for every u>0 the function gu{y)=f(ui/f(y)) belongs to CM, that is, (-l)n^M^09 n = 0,l,2,..., (4.1.9) dyn then x — ifr(y) is a completely monotone mapping Proof. We have by assumption (4.1.9) ay ax ay and by assumption (4.1.7) we have, for fixed y, ax for sufficiently small u and x=uijr(y). Therefore dft/dy^Q, which proves (4.1.4) for n — 1. For n ^ 2 we have =(_1) L &*r+S—1-" ^5i^sJ' and if for fixed y and small u we divide through by — df/dxu and then let u->0, then (4.1.7) and (4.1.8) will lead to (4.1.4) as claimed. Theorem 4.1.4. if \jf(y) satisfies (4.1.4) and tfr(y)>0, then ^(0 + ) exists, and we have a representation ^(y)-^(0+)=J —j—d<r(t), <r(A)Z0, (4.1.10) poo sothatalso f{y) = %+cy+\ (l-e-^d^t), a;(.4)£0, (4.1.11) Jo+ where Co=f(0 + ), c=cr(0 + )-cr(0), <r(t) - cr(0 +) = j %(t). (4.1.12) Conversely, if a function f{y) in0<y<bo can be represented in the form (4.L11) with c0^0, c>0, then y=i/r(y) is a completely monotone mapping. Also, the integral in (4.1.11) is o(y) as y->co, and the representation is unique. Proof. Relations (4.1.4) imply that ft'(y)eCM, and therefore we have /»oo ^(y)« e-**dcr(t), cr(A)ZQ,
LAPLACE AND MELLIN TRANSFORMS 85 and by Fubini's theorem we have /*oo e-et _ p-ty f(y)-Me)=\ dcr(t) (4.1.13) for 0 < e < y < 00. Now, the integrand increases as e decreases, therefore we can let € 4,0 in (4.1.13) which leads to (4.1.10). Conversely, if we are given (4.L11) and if we denote the integral by f0(y), then we put f0(y) = ft^y) + f^y), where Uy)=\1 (l-e-**)dX(t)] Jo+ ! Uy)~\™ (i-e-**)dX(t),\ Jl+ J and examine the two parts separately. We have Jo+ % eJ+o so that the last integral is finite. But then we can put ^i(y)-^i(0 + )=r dVF e-**tdX(t), which proves that r/r[(y) is in CM. Also if we write (4.1.14) fifa) ■tdX(t) y Jo+ ty and consider that g-1( 1—e_£) is bounded in 0 < £ < 00 we can let y -> 00 under the integral and we obtain ^(y) — o(y) as y ~> co. Next, owing to ^J^XW^ (1-e-*) dv(t) = ^2(l)<oo, it follows that we can put W3/) = c2~P° e~tydx{t), so that again iJr'2(y)<=CM, and i/r2(y) is not only o(y) but also 0(1) as ?/->oo. Finally, our reasoning implies that we have f'(y) = c+r e-*tdx(t), J +0 and by the uniqueness of the Laplace representation (4.1.2), which we take as known, c and %(A) are uniquely determined and so is therefore also c0.
86 LAPLACE AND MELLIN TRANSFORMS For actual applications the following conclusion from the preceding theorems will be needed: Theorem 4.1.5.7/ i/r(x) > 0, then eru^ belongs to CM for every u>03 if and only if t/r'(x) belongs to CM. In particular for 0<p^l the function e~uxP belongs to CM for every u > 0. 4.2. Completely monotone functions in several variables If k ^ 2 we denote by T the closed octant 0 S t3- < oo, j = 1,..., h, and by X the open octant 0 < x3 < oo, j = 1,..., Jc, and we are introducing the class of functions f(xv ...,xk) in X which can be represented in the form - m-j^-totdpit) p(A)Z09 (4.2.1) where, as always, (x,t) '£& + ...+zktk. Such a representation, if possible, is unique, and this can be deduced from the uniqueness of Fourier transforms as follows. If we introduce the complex variables Z5=Xj + iyj} j=l,...,Jc, then the integral f e-C*.*)dp(t) = f e-tM e-te *> dp(t) (4.2.2) J T J T is absolutely uniformly convergent in the £tube' (xv...,tk)eX9 -oo<y,<oo, j=l,...,.%, and there is a holomorphic function in Zl9...,Zk whose values are uniquely determined by those of the original function f(xl9 ...9xk). Now, if we choose a fixed point (#J,..., xfy in X and vary the y/s, then we can write for (4.2.2), e-W'QdFQ), (4.2.3) where F(A) = e-^*>dp(t), (4.2.4) for any AmEh. This F(A) belongs to V(Ek) and therefore it is uniquely determined, and since, for ACT, we have p{A)=i e&°>QdF(t)> it follows that p(A) is uniquely determined too.
LAPLACE AND MBLLIN TEANSEOBMS 87 Theorem 4.2.1. A function f(xj) in an octant X can be represented by an integral (4.2.1) if and only if it is CC°°>, and we have for all combinations nx^0, ...,n& = 0, the representation being unique then. This theorem will be taken as known, and we will proceed to further developments. If for any point (£,) in the point set closure X of X we introduce the operator then the set of conditions (4.2.5) is equivalent to the set of conditions (-l)wD^D|2...D^/^0, n=0,la2}... (4.2.6) for all possible points g1,..., §M in X, and this new phrasing of the conditions has the following advantage. If we take any nonsingular anine transformation and the transposed transforma- tion Uj-^Ha^Up—which together leave the inner product (x,t) invariant—and if we denote the corresponding images of X and T by the same letters, then the conditions (4.2.6) remain identically the same for representing a functionf(#) in X by an integral (4.2.1) over T. Also, the new point set X is an open ' affinely placed' octant, and T is its 'dual' octant, this one closed, and for X given T can be described as consisting of those points t=[tv .... tk) for which (x,t)^0 for xeX. (4.2.7) If we take two affinely placed octants X19 X2 whose intersection Xx f| X2 is nonempty and if in Xx U Z2 we are given a function f(x) which there satisfies conditions (4.2.6) for all g* in X± U X2> then it satisfies them separately in Xl9 X2> and thus there are two representations „ f(x)=\ e-^dPl(t), xeXv (4.2.8) jt. e-MdpS), aeX2, (4.2.9) where Tx is dual to Xt and T2 to Z2. However, in Xx n -^2 there is
88 LAPLACE AND MELLIN TBANSPOBMS some affinely placed octant X3, X3 C Xl9 X3 C X2, and for its dual we have T3 D Tl5 T3 D T2. Thus we have a third representation f(x)=\ e~l*>UPz(t), x<=X3i (4.2.10) but since p±(A) in Tx can be viewed as a set function in T3 which happens to be zero in T3 - Tv and p2(A) as one which is zero in T2 - Tv therefore, by the uniqueness property, a comparison of the three representations (4.2.8), (4.2.9), (4.2.10) implies that we actually have one'joint' representation /(a?)-=| e~Mdp(t) for x in Xx U X%. From this we can conclude that if we are given a family of affinely- placed octants {Xa} in which any two can be connected by a finite chain in which two successive ones overlap, and if in X= U aXa we are given a function/^) satisfying (4.2.6), then it can be represented by the integral (4.2.1) in which T is the intersection f\ aTa. Also, T can be defined purely in terms of the set X by (4.2.7), and it always contains at least the origin (0, ...5 0). Now, if the integral (4.2.1) converges in a point set X it also converges in its convex hull, and the point set T as denned by (4.2.7) remains the same. For our point set X, the hull is an open set, and, if it contains the origin it contains the entire Ek and the point set T must consist of the origin only, so that f(xj) is a constant perforce. If, however, the hull does not contain the origin, then it is itself a connected union of octants, and it is a £ cone' in the sense of the following definition: Definition 4.2.1. A cone is an open convex set not containing the origin which is a union of affinely-placed octants or equivalently a union of half-lines {pxv ..., pxk}, (xlt..., xk) e X, 0 < p < oo. We call it a proper cone if it is part of an affinely-placed octant, that is, if after an afrlne transformation centred at the origin it becomes part of the octant (c1>0,*..,xk>0. A cone is a proper cone if and only if its dual set T as defined by (4.2.7) contains a neighborhood,'in which case T is itself the closure of an (open) proper cone. Otherwise, T is contained in a linear sub- space through the origin, and has this structure there, or it reduces to the origin. Returning to the function f(x) we may now summarize as follows:
LAPLACE AND MELLIN TRANSFORMS 89 Theorem 4.2.2. A function f{x}) in a cone X has a representation (4.2.1) there if and only if it is C<°°) and satisfies relations (4.2.6) there. These functions will be called completely monotone and their class will be denoted by OM(X). Definition 4.2.2. If X is a cone in Ek: (xv) and F is a cone in i^:(yG), &g:l,l^l,&$l, and if ^ = ^2>(yi>-",^)> 2> = 1,...,& (4.2.11) is a transformation from F to X, then we call it a completely monotone mapping if i/rv e C{c0), and if for any points n1,..., rjn in F, and any y in F, the point in Eh with the coordinates ^p=(-i)w-V---£Vv^, i>=i,...,fc lies in the closure X of X. Theorem4.2,3. Iff(x)eCM(X),and{4t.2.11)isacompletelymonotone mapping, then g{y) =f(t/rl9..., r/rh) € OJf (F). Conversely, if (4.2.11) is a transformation C^ from Y to X, if X is a proper cone and if for every teT, which is the dual to X, the function r(^i(y)+-+W!/)) (4.2.12) is in CM{ Y), then (4.2.11) is a completely monotone mapping. The proofs, though outwardly more elaborate, are the same as for &= 1 and we will omit them. Theorem 4.2.4. A single function %='ijr{y1,.",yi) in Y is a completely monotone mapping from F to the one-dimensional cone 0 < x < 00 if and only if we can put ^0 = cO + 2Xyff+f (l-e-^)dcr(u) (4.2.13) a J & where c0 ^ 0, cq ^ 0, cr(A) ;> 0, and U' is the point set arising from the dual U of Yby deletion of the origin', the representation (4.2.13) being unique then. If X is a proper cone, then a transformation (4.2.11) from Y to X is a completely monotone mapping, if and only if we can put W^) = ^o + 2Xa^ + f (l~e-<^)d<r» (4-2.14) q J U' where cv0,cm are real numbers and each crp(A) is the difference of two
9o LAPLACE AND MELLIN TRANSFORMS non-negative set functions on U', cr!i>(A)~cr%{A) — o,p (A), such that for each (tv ...,4) in T the combination $ P &Q J XJ' \ S> J is as in the first part of the theorem. The second part of the theorem can be easily obtained from definition 4.2.2. As for the first part of the theorem we note that the cone Y contains an afnnely-placed octant Y0i so that correspondingly we have U C U0, and it is not hard to see that we need only continue with the proof for the case Y: {0<ya< oo}, U: {0 <uQ<oo}. Now, a function x—ifr(yQ) is a completely monotone mapping of this Y into 0 < x < oo if and only if the I functions %> S=1,...,J (4.2.15) exist and are a completely monotone function each, and we need only analyze functions of this kind. Now, if we are given (4.2.13), then it is not hard to generalize the reasoning used for I— 1 and show that fr(yj) has first partial derivatives, which can be obtained thus: dy« Cq Jtr e~(v>u)uqdcr(u), and this not only implies that the functions (4.2.15) are in 0M( Y) but that everything is uniquely determined too. The converse is less obvious. Since the functions (4.2.15) are in GM( Y) we have representations 57^+f <H^>do», 0=1,...,J °Vq J U' with certain set functions cra(A) ^ 0. Now, if we differentiate this with respect to yr and interchange (q, r) and compare, we obtain urdcrq(u) =uqdcrr(u), (4.2.16) g,r=l,..., l9 in the sense that urdorQ(%)« uqdcrr(u) (4.2.17) for every A C U'. For fixed r, if we denote by Ar the subset of U' for which ur=0, then (4.2.17) implies I uqdorr(u) = 09 g;=l,2,... ,1 Ar
LAPLACE AND MELLIN TRANSFORMS 91 Hence (%+.♦. +ul)darr(u)=*09 J Ar but in U' we have ux +... + u% > 0. Therefore we have dorr(u) = 0 for r = l,...,Z, L I Ar and from this we can conclude that in (4.2.16) we can divide out thus: daq(u) dcrr(u) in the sense that there exists a set function <r(A) ^ 0, such that I dorQ(u)=\ A for .A C U\ Therefore, we have ■^- = ca+\ e-^u^uQdcr(u), (4.2.18) cq ^ 0, or (A) j£ 0, and if for some fixed i}x > 0,..., 7}l > 0 we introduce the function lM!rt = ^(%y,.-.,7tf), then -Zl^JJc^+f e~y^^)(7],u)dor(u), and by a judicious application of theorem 4.L4 we now obtain lMyHlM0+) + XMrf+ f (l-e-^>u>)dcr(u). JET If now we put y= 1 and vary the 7/1}..., ?/z, denoting them by y1}..., yl3 we conclude that the function Ecrffe+f (l~e-(y>u>)dcr(u) is finite, and since its derivatives have the values (4.2.18) it differs from t/t by a constant c0> and since this constant has the value ^(0 +) it is ^ 0, as claimed. 4.3. Subordination of infinitely subdivisible processes If a completely monotone function in 0 < y < 00 m=\°e-»&y(t) (4.3.1) is bounded there, say /t(0 +) s y(+00) — y (0) = 1, and if we introduce the complex variable 7j = y+iy\ then the integral converges in the
92 LAPLACE AND MELLIN TRANSFORMS closed half-plane 0^y<oo, -oo<y'<oo, and uniformly in every compact set of it, and the resulting function which is holomorphic in the open half-plane will be denoted by Mv)- Likewise, a completely monotone mapping from y > 0 to z > 0, v(y)=cy+r (l-e*)dp(t) = cy+v0(y) (4.3.2) can be extended in the same manner into the closed half-plane, as can be seen from the decomposition (4.1.14). Now, for any of our 'exponents' i/r(oc3) in Ek we have Re^(ay) ^0, and thus we can form fe = ^(a,)), (4.3.3) for which again Re $*(ay) ^ 0. Since Re v0(t}) ^ 0 and v0(t}) is analytic in y>0, it follows that we cannot have Re v0(tj) = 0 at an interior point with y > 0 unless v0((t]) == 0; and if we have Re v0(n) = 0 for a boundary point 7) = iy', /»«> (l-costy')<2p(t)=0, Jo+ then this implies that the cr-additive function p(A) must be 0 on every Borel set A not containing appoint t=27rn/y'i and thus we must have Im v0(iy') = 0 likewise. In particular, Re *>0 Wai)) ~ 0 implies for any cceEk. uvr v °" Theorem 4.3.1. If{ert^aJ)} is any subdivisible process, and if we form (4.3.3) with any (4.3.2), then {e-^te)} is again such a process, and we call {ftioCj)} 'subordinate9 to {^(a^-)}. Afeo, a subordinate process of a subordinate process is again subordinate. Proof. Since /i(y; u)==e~uv(>v) is an element of CM for every u^0} we have /•<» My\ u) = e~rydry(r; u) (4.3.4) with dry(r; u)^Q, and hence /*00 e~*a>= e-r*Mdry(r; u). (4.3.5) Now, the left side in (4.3.5) is obviously continuous in a, and the integral obviously satisfies the condition J0 \3),g-l / if the integrand does, and thus Theorem 3.2.3 can be applied. Also, the 'chain rule' for subordination is a consequence of the corre-
LAPLACE AND MELLIN TRANSFORMS 93 sponding chain rule for completely monotone mappings, which in its turn follows immediately from definition 4.1.1. We now put where i/ft is the Gaussian addend as in (3.4.13) and i/r^ifr1 are the real and imaginary parts of fB + fp. Relation (4.3.2) leads then to ^ + ^=c(^-h^)4-Rei;0(^+^+i^)} (4.3.6) ^I=cfI + ImpQ(iJrG+fR + if^, (4.3.7) but since $R, i/rR and v0(ft) are all o(\ a, |2) as | oc | ->oo, by Theorems 3.4.2 and 4.1.4, relation (4.3.6) can be separated into ftG = aJrGy (4.3.8) $R = cfR+'RevQ{fG+fR+ifI). (4.3.9) All three terms in 4.3.9 are ;>0, and Re v0 = 0 implies also lmy0 = 0, and hence if we assume iJrR=Q, and then write ^(ql) instead of $*(#), we are led to the following major conclusion: Theorem 4.3.2. A subdivisible process with an exponent of the form i/r(cc) = fG{a) + i^J(a) (4,310) is not subordinate to any subdivisible process whatsoever but itself. The'stable laws' {e-*'*i2p}, 0<£><1 (4.3.11) are each subordinate to the Gaussian law {e~t|a|2}5 and quite generally if {e-^M} is any subdivisible process then so is {e~WaP} for any 0<#<1. Also, the relations (4.3.5) imply for the joint distribution functions the relations F(u; A) = P F(r; A) dry(r; u), (4.3.12) whose precise manner and range of validity will still be discussed, and they state that F(u; A) is a certain average over the set function F(r; A), the averaging weight being cdry(r; u)\ Also, the averaging process may be viewed either as a mixing of coexisting probabilities, or as a randomization of the parameter r in alternate probabilities. We might state though that the stable and similar subordinate laws are usually arrived at in the central fimit theory, and there the objects of study are families of random variables on a common ensemble {Q; £P\ P}, and not families of probability measures on a common ensemble {Q; S?}.
94 LAPLACE AND MELLIN TRANSFORMS Theorem 4.3.3. A process {frfa)} is subordinate to a Gaussian process Qfa) == £ cm ocP <xq if and only if we have ^(«,) = cG(o,) (4.3.13) and ^p(^)=f (l-cos(ot,x))f(Q'(x,))dvx, (4.3.14) where f(y) is a completely monotone function inO<y<co and Q'fa) is the quadratic form inverse to Q{ccs). Proof. By subordination we have #*,)-*<#(«,)+P (I-e-W^dpit), (4.3.15) and by the theorem 3.4.4 we have ^(ai) = ^(a/)+f (l-cos(a,a))dLF(;e), (4.3.16) J* and since by theorems 4.1.4 and 3.4.2 the two integrals are both o(\ oc |2) we hence obtain relation (4.3.13) and also ir*(a)=\ (l-cos(a,a))dF(z)=| (l-e-^(«j))dp(t). (4.3.17) We will now make use of the formula l_e~tQ(ocj)=hc f (l_cos(a,^))e-™Q'^d^ (4.3.18) where #'(#<,•) is the inverse to Q(ol5) except for a factor cx > 0, but since f{cxy) is completely monotone if f(y) is it will be justified to act as if we had cx=l. Now, if we substitute (4.3.18) in the second integral (4.3-17) we obtain (4.3.14) with /(y)=j^° «-•"§*>(*). f*co and for this we can write e~rvdcr{j), where d<r(r) = c0T^d —pi-1 . Conversely, if we start from (4.3.14) for any completely monotone f(y), we can gain back the second integral (4.3-17); q.e.d. Turning to spaces other than Ek9 we first of all note that the statements in section 3.6 can be supplemented as follows: Theorem 4.3.4. Our definition of subordination and theorem 4.3.1 also apply to locally compact Abelian groups.
LAPLACE AND MELLIN TRANSFORMS 95 Thus, in particular, if we are given any process e-r^m)^! e-27Ti^^dxF(r;x), (4.3.19) J Tic and if with any function (4.3.2) we form the exponents ft(m) = v(ft(m)) (4.3.20) then there is a process e-u$(m)= J e-27ri(m^dxF(u;x), (4.3.21) J Tic and in this particular case, theorems 4.3.2 and 4.3.3 likewise apply. Now, our 'subordination' can also be introduced on spaces in general provided we shift the emphasis from Fourier transformation (which might not even be definable) to (generalizations of) the distributions F(u; A), and we are going to do this now rather systematically. 4.4. Subordination of Markoff processes Definition 4.4.1. Any function y(r; u) in 0 ^ r < oo, 0 51 u < oo which occurs in a formula (4.3.4) will be called a subordinator, and it will be called a proper subordinator if y(0+;u)~y(Q;u)^Q for u>0. (4.4.1) Theorem 4.4.1. A function y(r; u) in r^O, u^Q is a subordinator if and only if (i) it is monotone in r and y(oo; u)~- y(0; u) = l, and y(0 +; 0) -y(0; 0) == 1, (ii) we have J r dry(r; u)dsy(s; v) = dty(t; u+v) (4.4.2) in the sense that y(t; u+v)-y(0; u + v)= (y(t-s; u)-y{0; u))dsy(s; v) ('additivity'), and (m)for each r0>0we have *dy(r;u)~0, (4.4.3) Hm p u \ 0 J rt i n that is, lim dy(r;w) = L (4.4.4) t* 4.0 Jo Also, if (4.4.1) holds for one u0>0it holds for all, and it so holds if and only if v(y)->co as y->oo. Proof. For a subordinator, (i) is obvious, (ii) follows from e~~uv(y) e~-vp{y) _ e-(u+v)v(y)
96 LAPLACE AND MELLIN TRANSFORMS and (iii) from /*oo foo 1 _ e-uv(llr0) ^ (1 _ e*/r0) dy(t; w) ^ (1 - e-*/ro) dy(t; w) Jo J ?*o dy(t; u). *hj: Conversely, if (i) holds we can set up (4.3.4), and if we write it as e-v(y,u) with v(y; u)^0, then (ii) implies v(y; u+v) = v{y; u) + v(y; v). Also r>ra [rQ ji(y; u) £ e-vtdy(t; u) ^e"^o dy(t; w), and (iii) implies limv(y; u) = 0 as u->0, and therefore v(y; u)=uv(y), as claimed. Finally, (4.4.1) is equivalent with lim /i(y; u)=Q, that is, uv(y)->co i/—>oo as y~>oo, and this completes the proof of the theorem. Theorem 4.4.2. In Ek (and perhaps in all locally compact commutative groups) the precise meaning of relation (4.3.12) is that we have f c(x)dj(u;x)=r (Y c(x)dxF(r;x))dry(r;u) (4.4.5) JEk JO \JEk 1 for all continuous c(x) which are 0 at infinity or any subclass as in lemma 1.5.1. Proof. If xMzL^Ek), then, as is easily seen, (4.3.5) implies f #(a)e-*">aX=r([ X^)e-^^dv\dry{r^u), (4.4.6) J Eh JO \jEk I and conversely if relation (4.4.6) holds for xia) variable and other entities fixed then (4.3.5) holds. Therefore, (4.3.5) is equivalent with relation (4.4.5) for functions c(x) which are Fourier transforms of functions in Lx(^fc). But any continuous c(x) which is 0 at mfinity is a uniform limit of such ones, and it is easily seen that (4.4.5) remains valid then. For r=0, F(r; A) is the 'identity', and not absolutely continuous, but for r > 0 we may have F(r;A)=j f(r;x)dvx, (4.4.7) f(r; x) g> 0; and if this is so then the right side in (4.4.5) can be written as Jo (Lfir; x)c{x)dv*)d*7{r; u)i (4A8)
LAPLACE AND MELLIN TRANSFORMS 97 provided y(r; u) is a proper subordinator. Also, if f(r; x) is a Baire function in (r, x) say, then F(u; x) is likewise an integral of a suitable function/(w; x) such that for every u we have f(u;z)=\ f(r;x)dry(r;u) (4.4.9) o for almost all x; but it does not follow that f(u, x) can be also assumed to be a Baire function in (u,x) automatically. Furthermore, apart from subordination, the continuity of F(r; x) at r = 0 can be expressed by the condition /• lim f(r;x-y)c(y)dvy=c(x) (4.4.10) t 10 J Eh as well. We are now going to envisage spaces in general but in order to avoid a number of complications we will generalize the densities f(r; x) only, and not also the set functions F(u; A) in general. Definition 4.4.2. Given a measure space {B; 01; v}3 (4.4.11) a Markoff (chain) density is a function/(r, s;x,y) which is defined for 0^r<s<oo; x,yeB and has the following properties. It is a Baire function in (r, s, x, y) and /(r,s; x,y)£0, f /(r,s; x,y)dvy^l (4.4.12) J B and f(r,s;x,y)f(s,t;y3z)dvy=f(r,t;x,z), (4.4.13) the integral existing without exceptions. We call the density continuous if we have lim f(r; s; x, y) c(y) dvy=c(x) (4.4.14) for c(y) € C, where C is a given set of bounded Baixe functions having the completeness property that for g(y) eL^R) we can have I g(y)c(y)dvy=0 R for all c(y) € G only if g(y) = 0 a.e. We call it time homogeneous or more frequently 'stationary' if there is a Baire function/(t; x, y) for 0 < t < oo; x, y eB, such that f(r,8;x,y)=f(8-r;z,y),
98 LAPLACE AND MELLIN TRANSFORMS the new function having the properties f(t;x,y)Z0, f f(t;x,y)d»v=l (4.415) J R f f(r\ x,y)f{8] y,z)dvy=f(r+s; x,z), (4.4.16) J B with (4.4.14) being now lim f f(t;x,y)c(y)dvy = c(x). (4.4.17) * 4-0 J R We call it space homogeneous if there is given on R a fixed transitive group of point transformations x' = Ux oiB onto itself such that U@=@, v(UB) = v(B), f(r,s;Ux,Uy)=f{r,s;x,y). (4.418) We call it (fully) homogeneous if it is homogeneous both in time and in space. The reader will easily verify the following statement: Theorem 4.4.3. Iff(t; x, y) is a stationary density, and if for a proper subordinator the integral r*co f(u; x,y)=\ f(r; x, y) dry(r; u) (4.4.19) is Baire in (u, x, y) then it is again such a density, and we call it subordinate to the original one. (A subordinate of a subordinate is again subordinate.) Also, if fit \ x,y) is space homogeneous or continuous then so is also f(u; x,y) respectively. Note that if (4.4.11) is the 'ordinary' Lebesgue space {E^, jtfk; v}, then (4.4.18) holds for the group of translations x'j—x^a^j—l,..., h, and in this case space homogeneity of a density means that there is a function f(r,s; z3) such that f{r,s; x,y)—f(r,s; yj—xj). Also, if stationarity is added then there exists a function f(t; zs) such that f(t;x,y)=f(t;yj—xj), and in the continuous case this is then the function occurring in (4.4.7) under the integral. Similarly for locally compact Abelian groups in general. Now. the group of translations is not the largest group for which (4.4.18) holds, not even among continuous transformations, the largest among the latter ones being the group of motions which includes orthogonal transformations as well. Now, for the latter group the functions f{r,$; z5) and/(t; %) depend on the distance | z | only, and the Fourier transforms of such continuous f(t; zs) are th6 functions
LAPLACE AND MELLIN TRANSFORMS 99 e-W*ft in which ^(o^) depends likewise on | a | only. Now, such a frfa) is real-valued since the imaginary part changes its algebraic sign under the orthogonal transformation aj= —a^, and, by uniqueness, the Gaussian and Poisson parts of frfa) must be radial separately. Here we have p ^(a,.) = c|a|2+ (l-coa(a,x))dF{x)9 (4.4.20) where F(UA) =F(A) for any rotation around the origin. Now, if we 'radialize' the integral, and denote by G{R) a function for which G(R) — G(r) is the value of F(A) for r^ \ x | <R, then we can write for it /»oo {l-Hw^(\a\.B)dQ(B))9 Jo+ where Hv(z) is the function cpJv(z)z~v, with cv being so chosen that Hv(0) = l. Therefore and if we let h -> oo, that is, v -> oo, then fl"„(2) -> e~^2. Therefore, if we put t = R2, G(R)=y{t), the formal limit of (4.4.20) is y(|a|2) = c|a|2 + f" (l-e-*"*'^)^) (4.4, J +o 21) and actually the following theorem can be proven: Theorem 4.4.4. If a function v(y) is such that v(\oc\%) describes a subdivisible process for | cc |2=a2+...+af3 for all Jc^l, then v(y) must be of the form (4.3.2). 4.5. A theorem of Hardy, Littlewood and Paley Definition 4.5.1. Given (4.4.11) we call a function/^, y)symmetric, iff{x9 y) =f(y, x), and we call {f(t; x, y)} symmetric if it is so for each t. Note that in Ek or Tk if f{x, y) =f(x~~y) and f(z) e Ll3 then f(x, y) is symmetric if and only if <j)f{(x) is real-valued. Theorem 4.5.1. Iff(x, y) is symmetric and f(x,y)Z0, jf(x,y)dvy=l=jf(x,y)dvx> (4.5.1) and if we introduce the distributive transformation m=[ f{x,y)g(y)dvy, (4.5.2) 7-2
100 LAPLACE AND MELLIN TRANSFORMS then for g(y) ^ 0 we have \g(yydvy (4.5.3) [h{xydvx<[g for every p ^ 1, and for #> = 1 equality holds. Proof. Obviously \h(x)dvx=\l lf{x,y)dvJg(y)dvy = Jg(y)dvy as claimed for p = 1, and for p ^ 1, if we apply Holder's inequality to f(x> y)1!a .f(x, y)1,p g{y) we obtain first h(x)* ^ (jf(x, y) dvy\ PlQ jf(x9 y) g(y)* dvy = jf(x, y) g{y)*dvyi and if we now integrate with respect to x we obtain (4.5.3), Note that only assumptions (4.5.1) have been used and not the symmetry in its entirety. Definition 4.5.2. We say that a Markoff density {f(t; x,y)} is of special kind if it is symmetric and if there is a constant A > 0 such that f(r-M; x,y)^A[f(r; x,y)+f{$\ x,y)] (4.5.4) identically in r > 0, s > 0, x, y e B. Theorem 4.5.2. Any density subordinate to one of special kind is likewise so, with the same A. Proof. f{u+v; x,y)=\ f{t; x,y)dty(t; u+v) f{r + s; x,y)dry(r; u)dsy(s; v) o f(r;x,y)dry{r;u)dsy{s;v) /*oo /*oo +A\ f{s\x,y)dTy{r\u)dsy{$;v) Jo Jo =A?(u;x,y)+Af{v;x,y). Note that the property (4.5.4.) is not possessed by the Gaussian density in El9 i f(t; x, y) = -e-WQ <*-*>2, (4.5.5) but it is possessed by the Poisson-Cauchy density 1 t Jo Jo /* oo (*o Jo Jo
LAPLACE AND MELLIN TRANSFORMS IOI for which it was introduced by R.E.A.C. Paley, in its periodic version (2.5.9) on Tv Also, their characteristic functions are e"""tc^ and e-27r*iaij so that the second is subordinate to the first. Theorem 4.5.3. For a density of special hind, if we take a Baire function t(y),yeR,0<t<oo and set up for g^Othe transformation AfaO = f(t(x); x, y) g(y) dvy = f(t(x); y, x) g{y) dvy, (4.5.7) then we have jh(x)* dvx S ±A*[g{yf dvyi (4.5.8) Remark. As known, under certain secondary assumptions, by suitably choosing t(y) the function (4.5.7) will be a.e. equal to h(x)= sup f{t;x,y)g{y)dvy, (4.5.9) 0<t<co J R in which case (4.5.8) is [h (xf dvx S ±A2 \g(y?dyy, (4.5.10) and this is the manner in which assertations of this sort are usually stated. Proof. Using Hilbert space theory, if for given t(x) we denote the operator (4.5.7) by h=Lg and introduce the adjoint operator g(y)=L*x s \f(t(z); z, y) x(z) dvz, then we have L(L*g) = jjf(t(z); y, x)f[t(z); z, y) g(z) dvydvz = (f(t(x)+t(z);x,z)g(z)dvz £ A f(t(x); x, z) g(z) dvz+A\ f(t{z); x, z) g{z) dvz =ALg+AL*g, and hence (L*g, L*g) = (fir, LL*g) ^A(g, Lg) +A(g, L*g)=2A(g, L*g), and this implies \\Lg\\l=\\L*g\\l<2A ||g||2. \\L*g\\z and this is (4.5.8)
102 LAPLACE AND MELLIN TRANSFORMS 4.6. Functions of the Laplace operator With any given subdivisible process {F(t; x^} in Ek we associate the semigroup of operators *»(%)=! 9(Xi-yi)dvF{t;ys)=Lt9> (4-6A) and in characteristic functions we write for this also $*(**) = &(**) *-*#**. (4.6.2) It follows with the aid of theorem 2.2.4 that these operators are bounded distributive transformations of the Banach spaces Lx and L2 into themselves, and we will view them as transformations of the intersection L12=LX fj L% into itself. Theorem 4.6.1. A semigroup of distributive transformations ht = Ltg, (4.6.3) 0 ^ t < oo of Llt 2 into itself is presentable in the form (4.6.1) if and only if (i) each Lt is commutative with translations and bounded in L2-norm3 (ii) for g^Owe have ht^0, and (Hi) it converges strongly in Lrnorm to identity at &e origin, ^ {] L^_g ^=^ (4 6 4) Proof. The 'only if' part is quite easy, and for the proof of the 'if' part the following proposition will be taken as known. Lemma 4.6.1. A bounded linear operator h=Lg of L%(Ek) into itself is commutative with translations (if and) only if there is a bounded measurable ' multiplier' xfaj) ^n $% ^ch that <f>h(^) = <f>Mi)X(<Xi) a.e. (4.6.5) Now since L1>2 is dense in L2, by assumption (i) of theorem 4.6.1 there exists a family of bounded measurable functions {x(t; oc)} such that ^(«) = ^(a)^;a), (4.6.6) X(r; oc)x(s; a) = v(r+s; oc), (4.6.7) a.e. Next, since L1>2 contains e-*ke~ne~llxl* it follows by assumption (ii) that, for fixed t, x(t\ oc)e-^iai2 js a positive-definite function a.e. for each e, and by theorem 3.2.1 the function v(t; oc) is therefore positive-definite after alteration on a set of measure 0. Finally, assumption (iii) implies Km e~i*i2| l-x(t; oc) |2aX~>0, t ^Oj Ek
LAPLACE AND MELLIN TRANSFORMS IO3 /. and all this implies that {x(t; a)} is a subdivisible process {e-^(a)} as claimed. Next, the following facts from operator theory will be taken as known: Lemma 4.6.2. If x(aj) ^s a (finite) measurable function in Ek (not bounded) then there is a distributive operator h — Ag which is expressed by 0*(a) = &(<*) *(<*)■ (4.6.8) It is defined for those elements geL%for which >.(a)|2(l + |z(a)|a)<to«^ "° (4-6.9) and these elements are dense in L2; also the operator is self-adjoint if X(&) is real-valued, and a (more general) 'normal' operator, if x(a) is complex-valued. These operators are commutative, and if we denote the operators pertaining to #(a) = otj,j=l,...,k,by B^~hh (4"6-10) then the previous operator can be denoted by according to a known definition of a 'function' of several normal operators which commute. The application of this is as follows: Theorem 4.6.2. The function (4.6.1.) satisfies the 'diffusion equation' ^£=-*&!,...sDk)hi9 (4.6.11) the derivative existing in norm in 0 fa t < 00. whenever the initial element g=h0 is one for which rJr(Dlt.... Dk) g is defined. In particular, if we introduce the (negative) Laplacean then for a process subordinate to it, i]r(a) — v(\ u\2).we have dht_ dt ™" whenever v(A) g exists, at any rate. ^=-v(A)ht, (4.6.13)
io4 LAPLACE AND MELLIN TRANSFORMS Proof. The Plancherel transform of l/£(A*+£—A*), 0^t<oo, 0<t + £<oois e-w«)_i <j)g(<x)erW*) and for fixed t > 0, this converges in L2-norm to $g(<x) e-W (_ f(cc)), whenever <j)g(oL) i/r(oc)eL2, at any rate. Turning now to spaces other than Ek we first note as follows: Theorem 4.6.3. The preceding definitions and statements can be adapted to {periodic) functions and subdivisible processes on Tk. Now, if a process on Th consists of densities, f(t; x, y) =f(t; x - y) ~ 2 e"¥(m) e2ni{m' x~v)> (to) then on putting fm(x)—e27riim>x\ we may write for this M x,y)~2e-W*fn(x)fn(3f), (4.6.14) (m) and this we now axiomatize as follows. On a measure space (4.4.11) with v(E) = 1 there is given a countable complete (complex-valued) orthonormal system r fm(x)fm>(%)dvx=8mmfm, (4.6.15) J R which is indexed by some labels {m}, and there is given a stationary Markoff density as previously denned which has an expansion (4.6.14) with exponents for which _ r Re^(ni)^0, (4.6.16) and this expansion is convergent in the following manner. If we take any element g(x)eL2(R) and assume that say g[x)£:Ot and if we introduce the expansion , . ,_ , . , .. n 7m * #)-Erm/^)5 (4.6,17) (m) then for every t > 0 the (non-negative) integral **(»)= (/(*; x,y)g(y)dvy=Ltg (4.6,18) is rigorously what a formal substitution of (4.6.14) indicates, namely, again an element in L2(R) and, in fact, the element whose expansion is **(*)-2 ^mW*(s). (4.6.19) (m) Now, with any set of complex numbers {v(m)} we can associate
LAPLACE AND MELLIN TRANSFORMS 105 a normal operator h —Kg on the Hilbert space L2(R) which is denned for those elements (4.6.17) for which S|rW|2(l + U(m)|2)<a)J (m) the value of it being Afir^ £ x(m) y(m)fm(x), (m) and we may now state as follows: Theorem 4.6.4. If we are given (4.6.14) and A pertains to the given multipliers {i/r(m)}, and if g is an element for which Ag is defined, then (4.6.18) satisfies j-u =J—AA» (4.6.20) strongly in 0 ^ t < 00. Also, iff(t; x, y) is subordinated tof(t; x, y) by (4.3.2), then, subject to secondary assumptions we have f(t; x,y)~Xe-*«K<>®fm(x)fm(3f)9 (4.6.21) (m) and for the corresponding normal operator we have A-v(A). (4.6.22) Returning to Tk for a moment, if fr(m) is real-valued, ^(m)£0, (4.6.23) or, equivalently, if the density is symmetric, and if we replace e-27ri(m,x) -fay j^s reaj an£ imaginary parts, then we can also write alternately f(f; X)y)^^e-^m)fm{x)fm{x)i (4.6.24) (m) where the symbols {m}, {fm(x)} are not quite the same as before and the latter again constitute a complete orthonormal system J B but now a real-valued one. In the corresponding analogue to theorem 4.6.4 the density/(t; x, y) is of course symmetric now, and the corresponding operator A which now carries again fm into i/r(m)fm is at present self-adjoint and not only normal, and this is an important difference indeed. The archtype of such an operator on Th (or Eh) was the ordinary Laplacean A, and with it we formed the functions v(A), and these constructions can be undertaken effectively on manifolds in general. In fact, if R is a
io6 LAPLACE AND MELLIN TRANSFORMS manifold, compact or not, and {g#} a positive-definite symmetric tensor on it, both of sufficiently high differentiability class Cr, and if we introduce the operator A=-7.A((^'i)' <4A26> and the volume element dvx~ ^Jgdx1 ...dxk, then a suitable f closure5 of A is self-adjoint, and for all compact R and many noncompact ones there exists a Markoff density for which (4.6.20) is the equation ^f=-M. (*-6-27) now. Also, if R is compact and v(R) = 1 then we again have an expansion (4.6.24) with {fm(x)} and {^-(m)} being eigenfunctions and eigenvalues of the operator, the latter occurring multiply, and for many noncompact E analogous expansions exist with the index m not being discrete any longer. Also, in all these cases, the function/(t; x, y) itself satisfies the equation st -=-AJ(t;z,y), (4.6.28) which is formally equivalent to (4.6.27) for 'all' initial functions ht—g, and our subordinate densities/(t; x, y) satisfy the corresponding equation %%. * dM^>=-HA)J(t;X,y), (4.6.29) in which the right side must be interpreted ' operationally * however, since v(A) is not a literal differential expression any more. If we view (4.6.28) as defining a Gaussian process then f=-AP/, 0<p<l (4.6.30) defines the corresponding stable processes say. But all these generalizations are symmetric processes only, and the problem of exhibiting nonsymmetric Markoff densities on manifolds in general remains yet to be broached. • 4.7. Multidimensional time variable The semigroup requirement F(rr)*F(s;')=F(r+8'r) (4.7.1) for a subdivisible process {F(t; A)} (4.7.2)
LAPLACE AND MELLIN TBANSTORMS I07 on Ek say, and also for a stationary Markoff density on any (4.4.11), if we ignore continuity, can be set up for elements r, s of any ' semigroup of addition' T0 in which a commutative associative operation r-\-s=t is defined. A situation of some interest arises if T0 is a cone in a Euclidean En: (t„) according to definition 3.2.1; and if in this case the characteristic function x(tl oc5) of (4.7.2) is continuous in (t, a) then it must be of the form exv[-(hfi(*j) + -+tnfn(*M (4.7.3) as can be shown. Also, the 'infinitesimal' description of the process is now given by the system of differential equations ^=-^(Dl3...,Dfc)h„ ^=l,...}n, (4.7.4) in which the operators ^(-^u • ••>A:) are normal operators which are such that for every (t1? ..., tn) in T0 the operator t^^-... + tnifrn has no spectrum in the left half-plane. Assuming that T0 is a proper cone we take in En: (£„) the cone X whose dual T is the closure of T0; and in another space Ef. (97A) we take a proper cone Y whose dual in Ez: (wx) will be denoted by W and, as in section 2.2, we introduce a completely monotone transformation ?a = *a(£i,..-,£»), A=1,...,Z, (4.7.5) from X to Y. This gives rise to Laplace expansions e-(^i(£+...+^Z(£))== e~(t>®dty(t; w), J t in which {y(B; w)} are Lebesgue measures on the Borel field {B} of T, and they again have the properties (4.4.2), (4.4.3), (4.4.4) of a sub- ordinator as before. Obviously, the function eXp l~ ? ^A(^l(a)) ***' ^(a)) I (4'7'6) describes a process subordinate to (4.7.3), and if (4.7.2) has a density f(t; x) and y(B; w) is a proper subordinator in the sense that it is zero outside T0, then the distribution function F(t; x) of (4.7.6) has again a density/(W; x) and the relation J{w\x)=\ f(t;x)dty(t;w) J TD holds.
io8 LAPLACE AND MELLIN TRANSFORMS This again applies to functions on the multitorus and to Markoff densities with expansions 2 exp [ - (t± rjr^m) + ...+*„ irn(m))]fm(x)fm(x) (w) on general spaces as before. Returning to (4.7.2), if this is Gaussian symmetric for every t in T, then ^(flfy) is a symmetric quadratic form Qv(a>j)i not necessarily positive semidefinite by itself, and we think that the following ought to be a pertinent definition: Definition 4.7.1. A subdivisible process {F(w; A)} on a proper closed cone W in some Ez: (wt) is subordinate to the Gaussian generally, if for some dimension^, there is given a completely monotone mapping (4.7.5) as described, and a set of quadratic forms Qi(ay)....,(2n(%) such that the characteristic function of ff(w; A) is of the form It would be of interest to study this class of processes for any given h and W, say. 4,8. Riemann's functional equation for zeta functions Theorem 4.8.1. In the plane of the complex variable s=<x+it let X(s) be defined and holomorphic in a domain containing the exterior of a circle , ,. ,,«,.. 1*1 ^Pq> (4.8.1) and its Laurent decomposition there be X(s)=#°(s)+X*(s), (4.8.2) X°«= S ^> %*(*) = S7n«n, (4.8.3) 5m|cwp»gp0. (4.8.4) 1/ #(s) possesses in a right half-plane an absolutely convergent expansion /•«, X(*H trM^dy, cr^<rr(>p0) (4.8.5) with a continuous $r(y)9 and a similar expansion XM=| My)*-ysdy, o-ScrI(<-p0)> (4.8.6) J —oo
LAPLACE AND MBLLIN TBANSFORMS 109 in a left half-plane, and if we have lim /v(o-+it) = 0 (4.8.7) uniformly in every finite interval a' g or ^ or", say, then the difference <l>M-Hy)-p{y) (±.8.8) is an entire function of exponential type, namely, =2^if ^)62/S^=2^i9 X°(s)eysds, (4,8.9) where C is the circle J s | =p0 say. Proof. By Fourier inversion of (4.8.5) we obtain, subject to a criterion, that is, «2/)=~r *°°x(s)e*sds, (4.8.10) and similarly ^z(2/)=5™. #(s)eysds, (4.8.11) and if these integrals are actually convergent, then, on account of (4.8.7), by the use of Cauchy's theorem we obtain for <j>r(y) — <j>i{y) the value ^ /• WiJ0x{s)eVSis- However, x*($) *s an entire function and therefore we may herein replace #(s) by #°(s), and if we substitute (4.8.3) we obtain the series (4.8.9). In general, the two integrals (4.8.10) and (4.8.11) can be evaluated as limits, for e->0, of 1 /V-Hco eeS*x{s)eysds, 2ni (theorem 2.1.3), and if we apply Cauchy's theorem first and put e = 0 afterwards, the result follows again. Lemma 4.8.1. For any entire function (4.8.9), subject to (4.8.4), we have J%(2/)e-^2/ = i^(=/(5))
no LAPLACE AND MELLIN TRANSFORMS for cr>p0and - p(y) e~ys dy=#°(s) for or< -p0. This follows readily by direct substitution of the series (4.8.9) and term-by-term integration. Theorem 4.8.2. If we are given two absolutely convergent integrals J — 00 J —00 the first for cr>crr and the second for c<orl} and if the difference (4.8.8) is a function of exponential type, then there exists a holomorphic function X(s) in a domain (4.8.1) with the property (4.8.7), such that xM^Xi8) for cr>0*r and Xi(s)=X(s) for <*«Tv Proof. It is easily seen that the two integrals in the sum x*W= f $r(y)e-ysdy+ [™Hy)z-ys&y J -oo J0 are convergent everywhere, thus defining an entire function. By the preceding lemma, and on using (j>r{y) = <l>i{y)Jrp{y)i we obtain, for <r>(crr,pQ), ro /*oo Xr(s)=\ My)e-ysdy+\ (My)+p(y))e"ysdy J -oo Jo =X*(s)+X°(s)> and for cr<(orli — p0) we obtain again Too /*0 Xi(8) = My) e~*sdy+ {(f>r{y) ~p{y)) e^dy JO J -oo Also, (4.8.7) is easily verified for #*(s), #°(s) separately, q.e.d. For application we introduce the variable x—ery, 0<#<oo and wite &(y)=*r(*), Hy) = ®i(x), Pto)=P(x), so that we have <Dr(a,)-®i(x) =P(<b), (4.8.12) >(s)ds 1 f 3 where P(x)=—■. (b - 2mJ x* and for either or > ar or or < crl we have the Mellin inversion formulas so-called ^ -, o«+i«x(8)d8 X« i = $(#) a^"1 dx, <b(x) = — Jo 27T*Ja._4.0C £s
LAPLACE AND MELLIN TRANSFORMS III If we have ®r(z) = 2X e-A^, 0<\1<X2<,.., (4.8.13) 1 and, for some 8 >0, ^W-lS^e-^, 0</«1</6a<...> (4.8.14) X 2 and P(<e)=—a0+-|, T which corresponds to v°(s) = —- -I—^, s s — o and if we put A0—/£0=0, then due to m As /•oo = e~Xxxs-1dx, <r>0, As Jo '(<?-*)_ f00 1 erPlaftf-ifai ^>o", the following conclusion ensues: Theorem 4.8.3. Given a 'modular relation' 2Xe-^ = -r Sftne-zW*, £>0, (4.8.15) o ^ o if the series gW = S^, ^) = 2.-~ (4.8.16) 1 ^» 1 /*w are absolutely convergent somewhere, then they are meromorphic functions and we have r^ ^ ^ T(S-s) Tj(d-s) (=X(s)), (4.8.17) X(8) = -?2+A.. + (entire function). (4.8.18) s s —6 Conversely, (4.8.16)-(4.8.18) implies (4.8.15). Riernann's own application was to derive from OO 1 00 v e~"7rtm2—— y\ e~(7r^^2 — oo * — oo the equation T(s) £(s) = T(| - s) £(| - s) ™ 1 for the function £(s) = 2 -7-^ • More generally, theorem 2.6.4 implies as follows. If in the notation of that theorem we put 1, RT(m+x) e27Ti^mj+x^ *j m=x ?w=s , B'T{m + y) e-*"&fimj+vj) *$ (S ^'(m+y))*
112 LAPLACE AND MELLIN TRANSFORMS then we have *v and there are two simple poles at most, and e(x)Br(x) i%y)R'M 1 *°w= — (detQ)* r+ik-89 where e(x) — l if x=zO (mod(m1,...,m&)) and e(» = 0 otherwise. In particular, for the function we have r(5)g(s)=irr(r + |^--s)g(r+p-5), where Pr(x) is any harmonic polynomial of order r=0,1,.... 4.9. Summation formulas and Bessel functions in one and several variables Theorem 4.9.1. Formally, if a pair of functions VW.tM} (4-9-1) are connected by the formula giji) =/.-»-«[" Js_x{2 V(/*A)) A*w-»/(A) dX (4.9.2) f l //A) in which case the pair 3(A#), -- gl -J > (4.9.3) is obviously likewise so connected for any x>0, then any modular relation oo j co 0* 0 ^0 implies the summation formula S^/(Aj = Ebn^n)J (4.9.5) 0 0 and therefore also £ aj^a) =~ £ bwf7 (—) - (4.9.6) o x° o \#/ Proof. We will take as known the formula ^e-W^=^-^-"i)rj^1(2V(M))A^-1)e"-Aa!dA, (4.9.7) x Jo which simply means that the pair of functions {e"\e^} (4.9.8)
LAPLACE AND MELLIN TRANSFORMS 113 is linked by relation (4.9.2). Now, the original modular relation (4.9.4) is a special case of the formula (4.9.6) for this pair of functions, and thus our theorem holds in this special case at any rate. However, starting from (4.9.7), if we multiply (4.9.4) by dp(x) and integrate over (0,00) then we obtain (4.9.5) for the functions /(A) - e-^dp(x), gUi)=\ -er^dp(x\ and if we substitute (4.9.7) we obtain (4.9.2) in this more general case, as claimed. In this way we have obtained actual criteria, which we will not reproduce here, for the validity of (4.9.5) for completely monotone functions /(A), in particular for a^O, bn^0, but the following alternate procedure was likewise featured. If we extend the Laplace transform (4.9.7) into the complex half-plane z=x+ioc, 0<x<co, — 00 < oc < 00, then by Fourier inversion it becomes equivalent with (•*+*» e-filz+te (p-hit-D J$_x(2 J(M) ^_1) for A > 0 - dz—{ 27ri]x-ico z* w~~[ 0 forA<0, (4.9.9) which is an important formula due to Sonine. And in keeping with this approach, if we start from CO 1 CO for some x=e (> 0 but small), multiply both sides by d<x(<x) and integrate over (™oo, 00), then this verifies our theorem for appropriate classes of functions representable by integrals of the form /(A) = j °° e-<e+ia*dcr(x), \ |<fcr(a)|<oo, J —CO J —CO and with some caution we can even let e->0 as well. However, we will not reproduce details here, but we will turn to functions in several variables instead, for some first statements at any rate. In the Euclidean Ek of the points #= (xx, ...,xk) we take a proper cone which we now denote by P. Its (closed) dual will be denoted by 0, its points by A=(A1,.... A&) and fi—Qi1, ...,/*&), and we will also introduce sequences of points Aw — (/§,..., A*) and /in — (/4, • • •, /4) • As a generalization of the half-plane z=x + i<xwe introduce in the space Ez7e of the Jc complex variables Zj—Xj + ioij
ii4 LAPLACE AND MELLIN TRANSFORMS the point set (xv...,xk)<=P, -oo<a*<ao, j=sl,...,fc, and denote it by TF ('Tube' with basis P). Next, as a generalization of the transformation y=x~x which dominates the structure of (4.9.4), we assume that there is given on P an analytic involution y=Ux, (4.9.10) that is, a transformation ys = U$(x) for which UU=identity, (4.9.11) and we assume that an analytic continuation of (4.9.10) transforms jTp into itself holomorphically. Finally, as a generalization of the denominator x8 in (4.9.4) we assume given in TP a holomorphic function R{z) for which R(U(z)) = (R(z))~\ P(z) + 0 in TF) and R{x) is real for x in P. As a generalization of (4.9.4) we introduce a relation (if existing) of the form ^ 1 «> 2 a^-tt^ S &we-WW>, (4.9.12) with An,/iw in 67, and again (A, z) = A1^ 4-... + A*fcfc, and our aim is to find a suitable transformation 9(M)=jG8(p;\)f(X)dvx, (4.9.13) in which S{/i; A) is meant to generalize the function /l~^-i)j,_l(2V(M))A^-1), such that (4.9.12) formally implies the summation formula ^anf(Xn)^bng(/in) (4.9.14) for any pair of functions so connected. For ft in G and A in Ek we set up the multiple complex integral (4.9.15) for a point (#l5 ...,#fc) in P. If for some fixed #££ 1 we introduce the quantity
LAPLACE AND MELLIN TRANSFORMS "5 then the integral (4.9.15) obviously exists if M1 (x) < 00, and if, more precisely, we have sup \mi(x)\<cx> (4.9.16) for every compact subset A of P then the integral is indeed independent of the point x chosen, as can be proven by shifting the coordinates x5 individually, and altogether several times. Furthermore, if also M2(x) < 00, then by Plancherel equation we have (4.9.17) and if we assume that for every x° in P we have sup M2(tx°) <oo, (4.9.18) then, because of (/i,Re U(z))^.Q, (4.9.17) implies that we have #(^^ = 0 for A not in 67. Theorem 4.9.2. Under the assumptions (4.9.16) and (4.9.18) we have for ju, in G, 1 /• xx + i 00 /» xtc+i co e-(/i, Z7(2))+(A, z) ( 0 for A not in 67, the integral being independent of x in P. For B{x)~xd the two assumptions are fulfilled only for 8>1 and 8 > J respectively, whereas the conclusion happens also to hold for 8>0. Now, it could be shown that our theorem also holds if the assumptions are fulfilled for M^x) for some p > 1 instead of only p = 1,2, and such a generalization would include 8 > 0 in particular. Next, by partial differentiation under the integral sign in (4.9.19) the following conclusion can be obtained: Theorem 4.9.3. Ifp(z1, ...,z&), q(zl3 ...,2fc) are polynomials for which q(U(z)).p(z) = l, (4.9.20) and if q*(z) is the polynomial adjoint to q(z), then for any R(z) we have the two-fold partial differential equation t iP\A^\ Sfo A) =£(/<; A), (4.9.21) subject to convergence assumptions which bear on B(z)i this equation being usually of an elliptic-hyperbolic type peculiarly ' mixed '. 8-2
ii6 LAPLACE AND MELLIN TEAKSFOBMS For the ordinary Bessd function we obtain in particular ^ {ir*Jy(2 V(M) *iy) = -riyJ7(2 V(M)) A*?, (4.9.22) which is equivalent to the classical equation, of course. Frequently we will have q{z) =p(z), B{z) =p(z)s, and then in addition to the 'principal' cone P there may be other cones to which our analysis applies, thus giving rise to 'Bessd* functions of 'second' kind, and possibly also of 'third' kind, etc By Fourier inversion of (4.9.19) we obtain as follows: Theorem 4.9.4. For z in TP we have !-_= S(ii-\)e-Mdvx (4.9.23) -#(z) J a {the integral converging absolutely), and after replacing z by U(z) we also have the inverse relation e-(/^)-= S(FiX)-—rdvx. (4.9.24) Now for a finite sum /(A) = £ C/, e~&>*"> (4.9.25) p the transform (4.9.13) has the value e-(fi,U(zP)) and therefore theorem 4.9.4 leads to the following conclusion: Theobem 4.9.5. Formally, (4.9.14) holds and the transformation (4.9.13) is self-inversive, meaning that it implies /(^)-Jff5(wA)jr(A)ctoA. (4.9.26) An obvious (but important) multidimensional case arises if P is the octant x1 > 0, ...sxk> 0 and R(x) is a product a?Ji... xpc for positive exponents 8l9..., Sk, equal or not. Another situation arises (and apart from variants and combinations it is the only nonobvious one known) if for a dimension k=m(m +1)/2 we consider the symmetric real matrices and as independent variables take the quantities x^p, l^p^m, ^2 xm, l^p<q^m. The domain P shall consist of matrices which are strictly positive-definite, and the dual space consists then of matrices
LAPLACE AND MELLIN TRANSFORMS 117 which are positive-semidefinite, the inner product (A,#) being then the matrix trace, m trace (Xx)= 2 Xmxm. 2>,ff—1 The involution U{z) is the matrix reciprocation z~x, and E(z) is (det {z))$, for 8 sufficiently high. The function S(ju,; A) is then the multiple integral and for this function we showed, in a somewhat more general setting, that the product Up; A) = (det \fiX-1 |)*-««+«S(/t; A), which is the actual generalization of/^_x(2 *J(/iX)), is symmetric in/i3 A, !(/*; A)=/(A;/0, and this is a rather nontrivial fact, apparently.
u8 CHAPTER 5 STOCHASTIC PROCESSES AND CHARACTERISTIC FUNCTIONALS 5.1. Directed sets of probability spaces We are recalling the definitions in section 3.7 to which we will refer soon. Definition 5.LI. A set of elements A: (A) is called directed if for some pairs of elements there is given an order relation ' <' such that (1) A<A, (2) A</£and/£<v implies \<v, (3) \<ja and /i< A implies A=/fc and, what is important, (4) given A, /i there is an element v such that A < v, fi < v simultaneously. Definition 5.1.2. If we are given a family of sets {QA: M) (5.1.1) which is indexed by a directed set, and if for every A, /i with \</i there is given a transformation from Cl^ to QA, also called projection, such that /AA = identity (5.1.3) and hv^hiAffiv) for \<[i<v\ (5.1.4) then the so-called projective limit of the family—which we will denote by Qqo : (wOT) or also by |Q: (w)—is defined as follows. A point w^ is any assemblage of points ^ ^^ (5 L5) from the sets (5.1.1) such that for each A in A there occurs one point wA from QA, also called the Ath component of co^ and that for any \</i these components are connected by (5.L2). We are also intro- ducing the mapping ^^ ^j . {51Q) to the Ath component for which we obviously have fto^hfJtw A</c, (5.1.7) and we call it the projection from 0,^ to £2A. Also, we call our family (5.1.1) simply maximal if all projections /a/^/aco are into all of 0A, so that in particular every point wA in QA, for every A, is the Ath component of at least one point in O^.
CHARACTERISTIC FUNCTIONALS II9 We will call it sequentially maximal if for every increasing sequence of indices A1<A2<A3<... (5.1.8) (which may be finite) and any choice of points u)\r in :QA with there is point o)° in QM for which o)lr=fXrQO(o)°)} r—1,2,.... Definition 5.1.3. A stochastic family is a family of probability spaces {QA;^A;PA} (5.1.10) which is indexed by a directed set together with simply maximal transformations (5.L2) from Q^ to Clx such that each fX[i is also a consistent mapping of the entire probability space {0,; .?V, P/J (5-L11) into the entire (5.1.10). We call the stochastic family topological if each probability space (5.1.10) is topological and the mappings (5.1.2) are continuous. Now, due to the (simple) maximality stipulated, eachfAoo generates a probability space ^; p*. p*j (5.1.12) such that S^=fxZ(Srx), I%SZ)=PMUSt)), (5-1.13) and we denote by ^* = UA ^* the union of all sets in all cr-algebras ^*, this being a finitely additive algebra of sets in Q^, obviously. If a set S* occurs in both S?\ and ^*, and if A </i then our consistency assumptions imply p*((Sf*)=p*(^*); (5.1.14) and for arbitrary indices A, /i we can find a third v, so that A < v and fjb<v and thus (5.1.14) holds in all cases. If therefore we denote the common value of the two sides in (5.1.14) by P*(S*), then we are led to introduce a ' finitely additive' probability space {Q;^*;P*}, (5.1.15) in which, however, ^* and P* are only finitely additive both. Definition 5.1.4. We say that a stochastic family has the Kolmo- goroff property i£ there is a cr-extension P of the measure P* from ^* to its er- closure £f\ and we then also call the probability space {Q; ^;P}, (5.1.16) the projective limit of the family (5.1.10) and the entire set-up a stochastic process.
120 STOCHASTIC PROCESSES AND Theobem 5.1.1. If a stochastic family is topological and sequentially maximal then it has the Kolmogoroff property, and thus gives rise to a stochastic process. Proof. By a general measure-theoretic criterion of Kolmogoroff it suffices to show that in the given circumstances there exists no infinite sequence of element S* in 5^* for which we have £p£p#p...->0 (5.1.17) and P*(£*)^a>0, r=l,2,..., (5.1.18) simultaneously. Suppose that such a sequence does exist. We can then find a sequence of indices (5.1.8) such that $*€^*r, and if we form Sr=fxr00{$*) then there is a compact set Cr in QAr such that CrCS„ PXr(Gr)>P,r(Sr)^r for all r. We put G* =fxX {@r)> and then so that obviously C*C#*, CpCp...->0, (5.1.19) P*(C*)^P*(^*)~ag + ...+^1^ia>03 and if we form back the sets Cr=fXrCO(C*) in DAr then they have the following properties. Each Cr is compact; we have PA (Cr) =P*(C*) > 0, so that Cr is nonempty; and we have f^x (C«) C @r f°r T<s- On ^ne basis of this we can pick a point in each Ori which we denote by ofr, and to this we add the further points ^r^A^A^C^s)? r < s> an(i we can assert that the sequence wj, ofr+1,..., has a limiting point in Cr. Next, if we replace {AJ by a subsequence of itself we may even assume that the limits v s 0 « lrm (i)*=z(jj\r€Cr exist, and from the continuity of fXfl it follows that (5.1.9) likewise holds. The sequential maximality now implies that there is a -point o) in Q with these limits as components, and this means that the sets C* have a point in common, which contradicts (5.1.19). Hence the theorem. We do not know whether the sequential maximality is really needed, and also whether the limit space (5.1.16) is likewise topological, and if the given spaces (5.1.10) are strictly topological to what extent the space (5.1.16) is likewise so.
CHABAOTEBISTIO FUNOTIONALS 121 A measurable function F^oo^) on (5.1.10) gives rise to a function ^H=n(/riK)) on (5.1.16), again measurable, and as previously stated we also have f F$MdPK=\ FA(w)dP (=E{FX}). (5.1.20) An interesting situation arises if we are given a monotonely increasing family of measurable functions {Fx{o))} for which we have 0£Fi(<o)£Fp((o), AS/*. (5.1.21) If the index set A is not countable then the classical theorem that the limit of the integral is the integral of the limit cannot be asserted literally, but, for instance, the following version of it remains: Theorem 5.L2. Given measurable functions with property (5.L21), if we admit limit values -f oo then there exists a measurable function F(oj) as follows. For each increasing sequence of indices {AJ we have ]imFAr{o))^F{o)), a.e., (5.1.22) and there is some such sequence for which equality holds; so that if any other F*(a)) satisfies all relation (5.1.22), then F(q))?£F*(g)), a.e. Proof Assume first that we also have 0£Fi(co)£l (5.1.23) for all A, o), choose an increasing sequence {AJ for which sup^{FAr} = supA^{FA}} (5.1.24) and put F(o))=supr FArM • If for any other increasing sequence {/ts} we put G(o))=mVsFXs(co)t (5.1.25) then we can find a third increasing sequence {vt} such that for every pair r, s there exists a t for which Ar ^ vt, fis ^ n. For the function H(a))=sixptFVt(a)) we thus have F(a))S 27(a)), G(o))^H(o))} and in particular E{F}<E{H}. A comparison with (5.1.24) now implies E{F}=E{H}, and hence H(o))=F(o)) a.e. Therefore, we also have 6r(o>)SF(6>). a.e. for every function (5.L25), and this proves our theorem for (5.1.23). If this additional assumption is not given, then we can introduce the new functions jf*(^»i _ e-uF*(co) (5 j ^6) for some fixed u>0 for which assumptions (5.1.21) and (5.1.23) hold both, and the theorem follows.
122 STOCHASTIC PROCESSES AND We have introduced the transformation (5.1.26) rather than another one because of the following theorem which will be suitable for applications: Theorem 5.1.3. If all functions are as in Theorem 5.1.2 and if we introduce the functions 0A(%)=| e~uF^dP(<D), <p{u)=\ e-uF^dP(o)), (5.1.26) then obviously <fi(u)—wf^$x(u), (5.1.27) and also <f>(0 + )=P{F(<o)<ao}, (5.1.28) that is, sJ(+0)=P(jS°), where S° is the set where F(u) <oo. Thus, F(oj) is finite almost everywhere if and only if <D(0 + ) = 1, and it is infinite almost everywhere if and only if <&(u) — Ofor some and then for all u > 0. In fact, <t>{u)=\ e-uF^dP(o)), Js° and for u 10 this converges increasingly to f dP(a))=P(SQ). J 5° Also, since <j>(u) is completely monotone in 0<w<oo it is either everywhere > 0 or =0, and hence the last statement in the theorem. 5.2. Markoff processes If T is any set and {A} a family of its subsets which is closed under set .addition A1U A2 then the point set inclusion A' C A" defines an order relation by which {A} becomes a directed set suitable for indexing. In our first application T will be the half-line 0 < t < oo, and the subsets will be the finite sets A=(t\...,**) (5.2.1) whose totality is obviously so closed. We take, as in section 3.7, a measure space {B; S\ v} (5.2.2) and denote by {Bl; g*>\ v1} (5.2.3) the familiar Z-fold product of (5.2.2) with itself, so that in particular Bl=BxBx...xB, I times, (5.2.4) and we also fix a point x° of B, put £°=0} and order the points in (5.2.1) thus: (0=P<)&<P<...<tK (5.2.5)
CHARACTERISTIC FUNCTIONALS 123 With any Markoff chain density f(r, x; s,y) on (5.2.2) we now set up the non-negative set function Fl{Bl\ x°) = J J UN*'1,a?3*"1; tp,s*)l dvl(x) = I II [/P*-1, a*"1; *p, a?*) &>(&*)], (5.2.6) and if for Bl=Rl we evaluate the integral by integrating successively with respect to x1^1"1, ...,x2}x1ia this order, then the property J /(t*-1, a?*-1; t*, x*) dv(x*>) = 1 (5.2.7) implies Fl(Rl; x°) = 1. Therefore ax=Bl; S?x=@l; Px(Bl)=Fl(B1;^) (5.2.8) is a probability space, and we associate it with the index (5.2.1) subject to (5.2.5). Given an index /i> A, we denote it by /*=(^...,^;t«,...,t™), and we must permit the new points tl+l,..., tm, if any, to be distributed on the line in any manner whatsoever. We now put £lp=Rm; ^=-Bw; P^m)=Fm(^m; x°), (5.2.9) denote the points of QA by (x1, ...,xl) and those of Q^ by and introduce the 'literal' projection x*=y*; j? = l,...,Z, (5.2.10) from 0,^ to QA, denoting if cOA=/A/t(w^), and we claim that our family (5.2.8) is made into a stochastic family hereby. In fact, properties (5.1.3), (5.1.4) and/^^) C ^ are obvious, and for the proof of the last and decisive property FWl %0)=Fm(fti(Bl); x% (5.2.11) it is enough to deal with the case m~l+I only, as is easily seen. We denote the additional point tl+1z=tm by s and we have either (i) t^1 <s<tp for some p = 1,..., I or (ii) tl<s. In either case the set fxKB1) iQ Bl x ft 9 and in case (i) say> a point of the set may be denoted by (x1, ...,xlh"lix,xp9 ...,xl) accordingly, and the expression (5.2.6) for JWfjB1*1; x°) is then f ...f(t^\x^; s,x)f(s,x; t*,x*)dv(x)...,
124 STOCHASTIC PROCESSES AND where the dots indicate factors not involving (s9 x). By (4.4.13) however this is /* .../(t^i,^-1;^^)..., J Bl which is the original expression for Fl(Bl; x°) itself, as claimed. In case (ii) relation (5.2.7) must be used for 3? — Z+l. Finally, the projections (5.2.10) are sequentially maximal, and finite products of strictly topological spaces are again so, and hence the following definition and theorem: De:etnition 5.2.1. The stochastic family (5.2.8) with projections (5.2.10) will be called a Marlcoff process (for the initial point x°) if the Kolmogoroff property is present. Theorem 5.2.1. If the underlying measure space (5.2.2) is strictly topological then the Kolmogoroff property is present, and we have indeed a MarJcojf process. If the process (that is, the underlying density) is space homogeneous then to a certain extent it is the same process for all initial points x°, as the following statements will imply in which more generally than before we put 0£P<fi<P<... <t\ where not necessarily t° = 0. Theorem 5.2.2. If the process is homogeneous for transformations {U} of (5.2.2), and if a bounded Baire function ^>(x°ix1}.... xl) on Rl+1 has the invariance property <j!>(Ux0, Ux19...9Uxl)^0(x°9x19...9xl)9 (5.2.12) then there is a function A$(t°, t1,..., tl) such that I <t>(x\x\...>xl)dF\x*,x«)=A${t\t1,...itl) (5.2.13) J Rl for all x°, and if we have in particular i <j>{x»9x\ ...9xl)=z n <l>»{xv-\x») (5.2.14) $9(Ux,Uy)=*4>9(z9y)9 p = l9...,l (5.2.15) then A^91\...9fi) =A^(tG9 fi) A^t\ **)... A^\ ft). (5.2.16) Proof. Ey (4.4.18) we have Fl(UB; Ux°)=Fl(B;x°)9 and if we denote the integral (5.2.13) by A^x0;^), then the substitution z*-+ U(x*>)9 p=091,2,..., I, leads to Afi(afi;t*)**Af(Uafl;t*),
CHARACTERISTIC JUNCTIONALS and since a transitive group carries any point x° into any other point y° this is indeed independent of x°. Therefore, we have 4v(x,y)f(r9x; s,y)dv(y) = A^p(r,s), J R and (5.2.16) follows if we evaluate (5.2.13) by integrating with respect to xl,x1'1,.... a;1 in this order. If (5.2.2) is the Euclidean {Ek; Jtfk; vk} (5.2.17) with translations, for any &^1, and if in {Fk)l=sRl: (x1,...^1) we replace, for fixed x°, the (vectorial) components x1, ...,xl by the new ones gi^^a* ga = a?a_aa> _f g*=a-i_a-i-i, (5.2.18) then space homogeneity means that we have f(r,x; s,y)=f(r,s; y-x)9 (5.2.19) <f>(x°>x\...}xi) = <j>(£\...,?), 4>,(x*-\x*) = fa(p), and because of dv^x1)... dvk(xl) == dvk(E})... dvk(^l)3 we have A^<\ti,...,**)= f $(p,...,p)dgF*(g), (5.2.20) where F*(£z) = f n S^"\ **\ £*) ^*(£3,)« (5.2.21) This gives rise to several remarks. Remark 1. If we associate the entity E? with the interval t*-1 <t^tp9 then we obtain an additive interval function and the probability measures (5.2.21) lead to a stochastic family and process for such, whereas the MarkofT process itself pertains to path and point functions as such. The 'randomization5 of interval functions will be presented later in a more comprehensive context (section 5.5). Remark 2. We note, however, immediately that the 'density' (5.2.19) can be replaced by a more general 'distribution' F(r,s; A%) with the properties F(r, s;A)^ 0, F(r, s; Ek) = 1 and F(r,8;-)*F(8,tr)=F(r,t;-) foTr<s<t, not only for a generalization of (5.2.21) which is f •••(" 'uFit^t^dA^Fit^^A^, J Ek J Ek P=*l
126 STOCHASTIC PROCESSES AND but also for a generalization of (5.2.6) which is f nF(tp-\tv; dB^-x^Fit1-1,?; B^-x1-1), and we have then in particular ^(^M^f 0,(£)W*-M*;£). J Elc Thus, if our Euclidean chain is also stationary, F(r, s; A)=F(s-r; A), then in forming the integral J Ejc we will be able to employ the most general subdivisible process {F(r; A)} as previously analyzed. All this applies of course also to locally compact Abelian groups other than Ek. Remark 3. Next, theorem 5.2.1 applies to (5.2.17), so that a Hmiting space exists and we claim that for a density (5.2.19) the quantities (5.2.18) are stochastically independent in the sense that E{fa(P) ...&(?)}= A E{U&)}- (5.2.22) In fact, as is easily verified, this is precisely the meaning of the relation (5.2.16) now. Theorem 5.2.3. Conversely, if we are given a Markoff density f(r, x; s, y) on (5.2.17) which is continuous in (x, y) and, what is restrictive, *>&$** f(0, xQ; r, y) > 0, (5.2.23) and not only ^ 0, for allr>0 and all y in E7c and x° fixed, and if for all 0<r<s,the two vectors £—x(r)—x°, rj — x(s) — x(r) are stochastically independent on (5.1.16) then f(r,x\ s,y) depends on r,s,y—x only. Proof. Put x°~0, say. We then have m(£)m}=jjmmmo; r,£)f(r,g; *,£+??) cm^m, Em)}^Em)^l}=j^)f(0,0;r^)dv^) E{f(r,)}^E{\ • f(V)}~jVfo)/(r,a; V)dv(7j) where f(r,s; n)sJ/(0,0; r, £)/(r, £; a, Uv)*>(£).
CHARACTERISTIC FUNCTIONALS 127 Now, if we are to have for all hounded Baire functions 0(g), rjr(y), then this implies /(0,0; r, g)/(r,& *,g+?)=/(0,0; r,g)/(r,«; 7/), except for a set of measure 0 in (£, 7;), the set depending on r, s. However, by our special assumption (5.2.23) we can divide through by flO,0;r,a,.othat /(f,|;,f£+„^r,,;,), and, due to the continuity of the function on the left, this is precisely the assertion made in the theorem. Remark 4. In all classical and modern population (and fission) problems known, it is customary (and probably also appropriate) to assume f(r,x; $,y) = 0 if either x<0 or y<0, which is an outright violation of our special assumption (5.2.23), and in the solutions of these problems the random increments {t;p} are usually not stochastically independent, which is in sharp contrast to the solutions of the classical diffusion problems (Brownian motion), in which the stochastic independence of the 'infinitesimal' increments has been a matter of axiomatic postulation, traditionally. Remark 5. Stationary Markoff chains as such, f(r9x;89y)=f(8-r;x,y), even nonhomogeneous ones, were easily generalized in chapter 4 from the half-line 0<t<oo to a multidimensional time variable in an octant, but the corresponding generalization of the processes offers a difficulty which we cannot readily overcome, and the diniculty is this that no order relation for the indices (5.2.1) can be then suitably introduced for which the decisive property (4) of definition 5.LI can be retained. Paul Levy in defining what he terms a 'multiple Markoff process' has also encountered a certain diniculty in actually estab- fishing the existence of his processes, and it is a difficulty of the same origin perhaps. 5,3. Length of random paths in homogeneous spaces If we restrict the points tp in (5.2.5) to lying in an open or closed interval Q^t<a or Q^t<a, then the resulting indices are again a directed set, and the previous constructions again apply. Also, in every case the limit space -Q: (oj) can be identified, in a familiar
128 STOCHASTIC PROCESSES AND manner, with the space of paths co: x=x(t) in B originating at the fixed point x°, 0 S t < & (or g a) and x(0) = x°. We now take in 0 ^ t ^ 1 a Markoff process which is space homogeneous and for which there is defined a non-negative Baire function p(rtx;e,y)£0 (5.3.1) for 0<r<s^I, xsyeB, with p{r,x;s,y)+p{s,y;t,z)^p(r,x;t,z), 0£r<s<t (5.3.2) (triangle property) and p(r, Ux; s, Uy) =p(r, x\ s, y) (5.3.3) (homogeneity). By theorem 5,2.2 the function A{r,s;u)=\ e~uP^x;s>y)f{r, x; s,y)dvy, u>0 (5.3.4) is independent of #, and in conjunction with (5.3.2) it also implies A(r}s; u)A{s3t\u)^A{rit\u). (5.3.5) Since also 0 <A(r, s; u) ^ 1, if we put A{r,$;u) = e~B(r>s>u) (5.3.6) then the new function satisfies B(r,s; u) ^0, B(r,s-, u) +B{s, t; u)ZB{r, t\ u), (5.3.7) and thus is, for each u, a distance function on the interval O^tg 1. If with the indices (5.2.1) we associate the functions then these fall under theorems 5.L2 and 5.1.3, and we are introducing the limit function F(o)) and the Laplace transforms <fix(u) &ncl ${u) as in these theorems. We note that the number F(o)) which is defined for almost all a), whether it be finite or infinite, is in the nature of a length of the path, even if the underlying distance function (5.3.1) should indeed involve the variables r, s, as we have permitted it to do. But we ought to point out that we know of no case in which it does so depend without F(a)) being infinite for almost all co. By theorem 5.1.2 we have (j>x{u) = n A{t*~\ t*; u)=ex$ [- S B{P~\ t*; u)\, p**i L #=a J and thus for the function <f>(u)—E{e~uf{^} we can write $(u) = e-°M9 (5.3.8)
CHARACTERISTIC FUNCTIONALS i2q I where C(u) = supA 2 S^"1, t2'; w) (5.3.9) is in the nature of a total variation of B(r, s; u) over the interval 0^r<sgl. (5.3.10) Also, if for each 0 < r S1 we introduce the corresponding variation C(r; u) over the interval (0, r) instead of (0,1) and put C(0; u) = 0, then C(r; u) is monotonely increasing in r and we have C(u) = C{l;u)=rdrC(r;u), of course. Frequently 0(r; u) will be absolutely continuous in r, so that the derivative _ ~, dC(r; u) DM=__ exists a.e. in 0 g r g 1, and we then have 0(^) = exp — D(r; u) dr . Also, if/(r, x; s, y) and p(r, x; s, y) both depend on s—r only (stationary), then we have A(r,s; u)=A(s — r; u)> B(r,s; u)=E.B(s — r\ u) and 0{s) — C(r) = <7(s — r); and D(r; u) = D(u) is independent of r altogether. Theorem 5.1.3 immediately implies the following criterion: Theorem 5.3.1. If C(u0) = oo for some u0>0, then C(u)==co, and almost all paths have infinite length. If C{u) is finite, then P(£0)=e-C(0+) is the measure of the set of paths of finite length, and almost all paths have finite length if and only if C(Q +) = 0. From this we will deduce as follows : Theorem 5.3.2. If for each e>0 there is a 8> 0 such that for u^8, s—r^d, we have 1-A(r,8;u)£e(l-A(r,8: l))+e(*-r), (5.3.11) then either almost all paths have infinite length or almost all have finite length. Condition (5.3.11) is fulfilled if the distance function (5.3.1) isbounded °nR' p(r3x;s,y)^p09 (5.3.12)
130 STOCHASTIC PROCESSES AND or more generally if for some p0>Qwe have e-uP(r,x;s,y)f(riX; S)y) dvy^e{s-r) (5.3.13) J pi f p(r,x; sty)^pa for 8-r£8zz8(e). Proof. We will denote by MXiM2,M3,..., certain finite positive numbers which are independent of r, s, e, d, the latter all in (0,1). By theorem 5.3.1 it suffices to prove that C(v)£Mv 0<^1, (5.3.14) implies rim C(w) = 0. (5.3.15) u JO Now (5.3.14) implies B(r, s; v)£Mv and from l~A(r,s;v) = l~~e~B(r>s'>*\ we thus obtain M2B(r>s; v)£l-A(r,s; v)SM3B(r,s; v). (5.3.16) Therefore M2B(r9s; u)£e(l-A(r98; l)) + e(s-~r)^eMzB(r,s; l) + e(*-r), and hence M% S Btp~\tpl u) £eMz S Bit*-1, P;l)+e^ (tv-tv-1) P = l p — 1 $=*1 £eJf80(l)+e==edf4. Therefore, M2C(u) ^eMi9 which proves (5.3.15). Next we have I~A{r,s;u)=\ {l-e~uP)fdv J R = + =JiM;«)+JaM; ^). J p^pa J p>po However, for p<>pQ and v^lwe have and hence M5vplI-e~«^Mevp, (6.8.17) 1-A(r,s;^)<rf7| (l-e^)/dv + J2^wM7(l-A(r,5;l))-hJ2, J P^po and by assumption (5.3.13) this implies (5.3.11) as claimed. Next, for stationary processes the reader will easily obtain the following statements from the definition of the symbols. Theorem 5.3.3. If our process is also stationary then either — 1-A(e;l) — B(e;l) Inn i =hm v = +oo, e\0 e e^Q e
CHARACTERISTIC JUNCTIONALS 131 in which case C(l) = oo and almost all paths have infinite length, or, however, we have - 1-A(e;u) B{e\u) lim —ijm _^ i = j}(%) < oo5 e|0 6 e 4.0 e the limits existing, and almost all paths have finite length. By applying (5.3.17) and theorem 5.3.2 this easily gives the following conclusion: Theorem 5.3.4. In the fully homogeneous case, if we have p{t;x,y)^p0 on E, or more generally if for some p0>Qwe have I f(t;x,y)dvy = o(t), t~*0, then almost all paths have infinite length or finite length depending on whether 1 /• lim- p(e;x,y)f(e;x,y)dvy = cc or <oo. e \0e J p^po If in Ek we are given a fully homogeneous process {f(t; Xj)} with transform {e"^(a^}, then for the ordinary distance p(x; y)=2n \ y — x \ we have r A(t;u)=\ e^^f(t;x,)dv99 and theorem 5.3.4 could be applied to this expression. However, by Fourier transforms this is also A(t; V-rrWm**!)}!^^?^. and an analysis of the behaviour of the difference quotient l/e{l-A(e;u)) from the latter expression is rather easier to undertake. Under the integral sign we then have to deal with the expression —TT-r-^(a), eyr{a) and since for e 10 this converges majorizedly towards i/r(oc), theorem 5.3.3 leads to the following conclusion: Theorem 5.3.5. If ijr{(x5) is real valued then 9-2
132 STOCHASTIC PROCESSES AND and almost all paths have either infinite length or finite length depending on whether r f{ch,...,gk)dva In particular if ^(ofy) ~i/r(\ot\) depends only on the distance |«| = (««+.. .+«*•)* $en. tfAe alternative is L = co or <oo; 1 /?* and thws /or the stable process ft(&j)~\a\Q we- have the result that irrespective of the dimension of the space we may expect infinite length if l^q<*2 (from Gauss's process to Cauchy's process, inclusive), and for 0<q<lwe may expect finite length. If ^/{olj) is not real valued we have at any rate the incomplete alternative that almost all paths have infinite or finite length depending on whether Re^(a,)rifog f | ffa) \dva f ^ffi-co or f m a l*+l <oo. The following supplement might be of some interest: Theobem 5.3.6. If iJr(a,j) = iJrG + iJrp is real valued then almost all paths have finite length if and only if the Gaussian part is missing, ijfG=0, and also the distribution function F(A) in the representation $*(<*,) = f (1 - cos 2iT(a, £)) d,F(£) (5.3.18) is such that we have | £ | d£F(£) < oo Jo<U!<i for the first power | £ |, and not only for the second power | £ |2 as must be the case automatically. Proof. In fact, for i/fl^kO the integral is always infinite, and for (5.3.18) it has the value c[ (l~e-2*i£i)d£F(£), whence the conclusion. Finally, we will very fleetingly comment on a certain type of process which arises in the following manner. Take a continuous function i/r(t; ccj) for Ogt<oo, aeEk, and assume that eruW>*$ is a characteristic function in Ek for each t and u > 0. If we approximate £ r/r(t; oc) dt
CHARACTERISTIC JUNCTIONALS 133 by Riemann sums it follows easily that exp - i/r(t; oc^dt is again a characteristic function for 0^r<s. Assume that it is the transform of a density (this could be generalized) and denote the latter by f(r, s\x). Obviously f(r, s; •) #/(s, t •) =f(r, t; •), and thus f(r,s; y3- — xd) is a Markoff density which is space homogeneous but not stationary. It satisfies the pair of differential equations and the statement we wanted to make is as follows: Theorem 5.3.7. Subject to secondary assumptions, the statements of Theorem 5.3.5 remain literally in force if we put now ijr(^)= i/r(r; oc^dr, D(u)= D(r; u)dr, Jo Jo where D(r- u) = 7r^k^rH(h + l)\ f U^T' ^ dV* where U[r,u) n l™lc + X)t]Bh(u* + c4+ ...+4P+1>* 5.4. Euclidean stochastic processes and their characteristic functionals We take a directed family of Euclidean spaces {QA} with connecting transformations fA/t as in section 5T, giving rise to a projective limit Q,: (a)), and we assume that eacht/A/^ is of the form m Zv=2,cPQyq, p = l,...,l, (5.4.1) ff*i where QAs= 2^: (xP) and £lfl=Em: (yq). We note that rank of matrix {cpg} = I, (5.4.2) since /A/t is a mapping into all of QA, as always assumed. With each given ^ _, tx a o\ & Qx: (0)^ = ^: {x9), (5.4.3) we now associate its Fourier analytic dual ^(^^K,), (5.4.4)
*34 STOCHASTIC PROCESSES AND and correspondingly with each 'inverse' mapping (5.4.1) the contra- gradient 'forward5 mapping i A=Sc^a^ q=l,...,m, (5A.5) 5P = 1 for which we also write ^>{i—9[ix &a> (5.4.6) and we note the following properties, the last of which is a consequence of (5.4.2): Lemma 5.4.L We have gVfig^=9vx for v>{i>\, (5.4.7) and E^a2> = 2..y<A> (5.4.8) so that (/(y),a) = (y^(a)), (5.4.9) and if for two points 2)^ o>\ and jlo>X, we have ^a(^a) s °m^a) then Definition 5.4.1. The ('forward') projective limit fi:(a)=IimA{fiA:(SA)}, A€A, (5.4.10) is defined as follows. A point & of it is any maximal collection of points {SA} (5.4.11) (called its 'components') and a subset A0 of A which depends on S such that (i) for each Ae A0 the collection contains one and only one point cox of &x> (fi) ^ ^v ^2 € A0 (where Ax=A2 is permissible) and A3 > A1? A2 then A3€A° and (iii) for the corresponding components we have flW^) = 9fA3A2(SA2) = SV (5.4.12) Theobem 5.4.1. (i) If two points of O have one component in common then they are identical, and every point 5A of every DA is the component of a point co which can be described as follows. For every A > A0 the point ^a=i7aa0(^a0) is a component of it and if for an index /i < A, ji < A0 and a point (0^ it so happens that gA/A(w/t)=:=i7AA0(^A0) then it is likewise a component. (ii) The mapping S=gcoAo(o)Ao) is a transformation of 0Ao into *5 such that „ „ ,* a io\ ^coA^oo/^A (5.4.13) for \<fi. (iii) For any n+1 points <or in &, r=0,1,2, ...}n; n^ 1, there is an index Xfor which the components <DA exist simultaneously, and & can be
CHARACTERISTIC FUNCTIONALS 135 made into a real vector space in such a manner that for any such com- bination of points and an index A the relation toQ=a1& + .,.+an(on (5.4.14) in tl implies the relation in&A. This can be proved readily with the aid of lemma 5.4.1. Also, the original projective limit Q can be likewise trivially made into a vector space in such a manner that the given spaces QA are vector projections of it. If now we introduce the notation i (^aA)=IX^ (5.4.15) and recall the consistency relations (5.4.8), then the following statement is easily proven: Theorem 5.4.2. On the product vector space £lx£l there exists a bidistributive functional (co, a)) such that we have (w.6) = KA) (5.4.16) for every Xfor which the component (i)\ exists. If we take a nonempty directed subset A! of A and introduce the corresponding limit spaces Q'=limAQ', £'=limAQA; AeA', (5.4.17) then £l' can be mapped vector-isomorphically into a part of Cl by assigning to a point u)r the point co having with it any one component in common for AeA'. The corresponding statement for the inverse limits is also true, but nontrivial, and if in particular we choose for A' a directed sequence {Aw} then this asserts the sequential maximality needed in theorem 5.1.1. Theorem 5.4.3. With each o)r € Q' there can be associated at least one point coed having the same components as a)' for AeA'. Proof. For fixed a)' and variable a)' we view (co', a>r) as a distributive functional on ft'. Now ft' is a vector subspace of Q, and since we are not concerned with ' completeness' or ' closure' of the spaces and functionals at all, there is a very simple general theorem available (based on the axiom of choice) stating that any distributive operation from a vector space U' to a vector space of values V (which for us are real numbers) can be extended to any vector space &D&'. We pick one such extension and denote it by (o)r, co). Now, if v is any index in
136 STOCHASTIC PROCESSES AND the large set A, thenfor the points co having components top = (yx,..., yn) in £lv=-En\ (y) the functional must have the familiar form and if now we put o)v=(z±i..., zn) then {o)v} == w is a point having the property asserted. Definition 5.4.2. We call a stochastic family Euclidean if each given probability space has a Euclidean structure Qy=Ez; &x^i\ Px(S)=Fl(A)} (5.4.18) and if all connecting transformations /A/t have the affine structure (5.4.1), with origin going into origin. By theorem 5.4.3 we may now assert as follows: Theorem 5.4.4. Any Euclidean stochastic family has the Kolmogorojf property so that a limiting probability space {Q; ST\ P} (5.4.19) exists, and the resulting stochastic process will be called 'Euclidean9. For each A, we now introduce the characteristic function 0Ate)=| e~27ri^^dFl(x), (5.4.20) J Ei and, in keeping with (5.4.15), we write for this also ^A(£}A)=f e-2^z>A^A)dpA. (5.4.21) But now the following statements become obvious: Theorem 5.4.5. (i) Our Euclidean process can also be characterized 'dually' by a directed family of data Clx=El:(^); 0a(&a) = ^i>.-.3^), (5.4.22) with connecting affine transformations (5.4.6) of the form (5.4.5) for which (5.4.2) and (5.4.7) hold; the functions 0A(a3-) bemg characteristic functions which are linked by the identities hi**)=tf,(A)=0/,Ma))> (5-4-23) or written differently 0A(6A) = ^^(gM(SA)), for \</i. (ii) On the limit space Q there is given a function 0(£)) such that for any component a>A of any a) we have ^(fi)=#A(fi)A), (5.4.24)
CHABACTEBISTIC FUNCTIONALS 137 and we can also write <fi((o) =\ e~27Ti^> & dP(o)) (5.4.25) J a = | e-^^'^dP^G)), the function (a), fi), and hence also e-2ni&A being a Baire function on (5.4.19) for each (b. Definition 5.4.3. This function 0(6) will be called the characteristic functional of the given Euclidean stochastic process. 5.5. Random functions Definition 5.5.L On a general point set T: (t), a family of its subsets D={A} will be called admissible if corresponding to any finite numbers of its elements A A A /■-,-■,% &v&2,--,&i (5.5.1) (which we will call a union) there is another finite number Ai,Aj,...,A^ (5.5.2) in which any two A,,Ag are disjoint for r=£s (which we then call a sum) such that each A^ is a point set sum of certain A'a which we will indicate by writing Qp K=ZK«> p = l,...,l (5.5.3) .2=1 We now take a function X(A) on D with real values say, and we call it additive, if every relation (5.5.3) implies Z(A,) =2 *(*«). (5.5.6) An admissible family D arises if we take for A any and every single point t of T (only), and in this case an 'additive' function X(A) is any point function X(t) whatsoever. On the other hand, if D is a cr-algebra of sets then our X(A) is an additive' set' function in the familiar sense; and finally we may take, for instance, on T={0<£<oo}, for D the family of intervals Aay?=(oc <t£J$), either for all real end-points oc, fi, or only rational ones, or even only diadic ones m/2n, with X(A) being then an interval function in the usual sense. Remark 1 (Important). A function X(A) on' A5, whether real-valued or later random-valued or Banach-valued, will always be intended to be additive as in definition 5.5.1, and sometimes, but not always, we will recall this for emphasis.
i3« STOCHASTIC PROCESSES AND Given D, we introduce a directed set A={A} in the following manner. An index is any sum • A= ^ ^ ^ ^ (g 6JJ) and if we are given another index /* = (Ai,Ai,...,Aj, (5.5.8) then we put A </£ if each A^ of A is a sum (5.5.3) of elements of ja, with some elements is!r left over perhaps. Thus, if we view A as an (incomplete) partitioning of T then jjl > A means that ft refines A and adds some elements, perhaps. This is indeed a directed set, and, in fact, the crucial property (4) of definition 5.1.1 is fulfilled because for any two indices Ax = (A*x), A2=(A|2) we can form the union {A*x, AJJ, and by definition 5.5.1 there exists then a sum /£= (A^) such that /\.1</i,X2</i simultaneously. Frequently the set A will contain a subset A' which will be a directed set likewise, and thus be ' admissible' in the forthcoming theorem itself. For instance, if the set T itself is a sum (5.5.1), then the subset A' of indices which ' refine' (5.5.1) is of this kind, and a very suitable directed set of indices it is. If for a function X(A) we associate with two indices (5.5.7), (5.5.8) the values _/A . V/A, ^ x1=X(A1)i...ixl=X(Al)s ) } (5.5.9) y1=Z(Ai)>...,yw«Z(A^,J then corresponding to (5.5.6) we have a?»55sy.pi+ — +y«.«J,» #-=i,...,l, (5.5.10) and if we put this in the form m A/.: 3*^2^.7, .p=l....,Z, (5.5.11) where cm are certain constants = 1 or 0, then these transformations satisfy requirements (5.1.3) and (5.1.4), among others. We can now apply the results of section 5.4, and we first state the following theorem, the first half of which is the easy one: Theorem 5.5.L If on an admissible family D we are given an additive function JT(A), 2(A1+...+A,) = l(A1)4-...+l(A2), (5.5.12) whose values X are random variables on some given probability space {&; y; P}, (5.5.13)
CHARACTERISTIC FUNCTIONALS 139 and if with every index in A (or only some A') we associate the Euclidean space {a^S%; ^s^; PA = FJ, (5.5.14) where Ft is the joint distribution function of the I random variables x1—%(A1),...,xlz=%(Al); then these Euclidean spaces constitute a stochastic family for the transformations (5.5.11). Conversely, if we are given D and for every A a probability measure Ft(A) in Elt and if they are linked by the given transformations (5.5.11) on D, then (5.5.14) constitutes a stochastic process whose limit space Q consists of all numerical additive functions co=X(A), X( A1 +... + A,) = X(AX) +... + X( Az), (5.5.15) and for each A the set function Ft(A) is the probability for the I values x1^X(A1), ...,xl = X(Al) to lie in the Borel set A. Remark 2. In fulfilling requirement (5.5.12) we may admit exceptional sets of measure 0, the sets varying with (A1} ....A?); for the constructed random function X(A), however, (5.5.15) is fulfilled without exception. Thus any random function may be replaced by another having the same joint distribution functions for which (5.5.15) is fulfilled without exceptional sets, but the new random variables are on a probability space altogether different from the original one. Also in what follows the symbol X(A) without the symbol'~' will frequently denote an arbitrary (additive) random function on D and not necessarily the one constructed in theorem 5.5.L Remark 3. If the value of the functionX( A) is a ^-dimensional vector, then not much need be changed, provided we replace the dimension I in (5.5.14) by kl and then take for Fn(A) the joint distribution function of the various components of the I vectors X(A$); and for k = 2 the case of complex-valued random functions is fully included in this way. Definition 5.5.2. For given D a step function on an index A is a real function on T which for t € Ap has a constant value h^ p — 1,..., I, and for teT~(A1 + ...+Al) has the value 0, and thus can also be written as 1 Ht)=HhpU*Jfi> (5-5.16) where o)A(t) is the characteristic function of the set A. For given A, all step functions on it are a vector space. Also for /f>Aa step function on A is also one on ft, and since {A} is directed we obtain the following statement: Lemma 5.5.1. All step functions together constitute a vector space which we will denote by L0.
140 STOCHASTIC PROCESSES AND If X(A) is a random function on D, then for a given step function (5.5.16) the sum 1 ZA^A,) (5.5.17) is independent of the special index A used, and depends therefore on the step function h(t) as a point function on T only. Deitnition 5.5.3. For any heLQi we denote this value by f h{t)dX(t)=~ J h{t)dX (5.5.18) J T J T and view it as a Stieltjes integral then. If now we put h^=ocP, F^) = x^ then (5.5.17) is the sum #!<%! + ...+^az, and theorem 5.4.5 implies the following one: Theorem 5.5.2. The space Q: (co) dual to the space Q of theorem 5.5.1 can be aptly represented by the vector space LQi so that (<tf,c!>)= J h(t)dX (5.5.19) and $(&) = $(h) = Ebxp[-2mt h(t)dxl\, (5.5.20) and this has the same value as r~27rir h(t)dii|, where J?(A) is any other random-valued additive functions on D having the same joint distribution functions FZ(A) as X(A) itself. Remark 4. If X(A) is vector-valued, then nothing else is changed if we also make h(t) vector-valued and in (5.5.17) interpret h*X as h'X'-^h"X',+ ...+hfMX<®, and for i-2we may also write for this hX+hX9 where h=4(h'+ih"), X=\(X'+ %X')9 so that then (*>,£>)« f h(t)dX+\ h{t)dX. J T J T We now assume that there is given on D a finitely additive measure v{A)~°> KA1 + ...-fA,) = KA1) + ... + (A,), and we are introducing the following definition: Definition 5.5.4. Given {D, v(A)}, we call the random function X(A) a homogeneous process, if there is a function ifr{ct), —oo<a<oo,
CHARACTERISTIC FUNCTIONALS I4I as in a subdivisible process, such that the characteristic function of the distribution FZ{A) in (5.5.14) is fafa,..., a,) = exp [ - vfAJ ^K) -... - i?(A,) ^(a2)]. (5.5.21) The random variables X(AX),..., X(AZ) are then obviously stochastically independent for any disj oint A1?..., A2 in D. Also, if T=(0 < t < 00) and A=~Aa/5=((x<t^/3), then this homogeneity is the same as the (full) homogeneity of a Markoff process if we interpret X(A) as the increments g=x(/3)—x(oi) of a path x(t), as was already stated in remark 1 of section 5.2. For a step function (5.5.16) the sum can be designated as the symbol f{h{t))dv, (5.5.22) L f T and we now have the following statement: Theorem 5.5.3. For a homogeneous process X(A) the characteristic functional has the value ft{h)=ex$[-( ft(h(t))dvA, h€L0, (5.5.23) so that for instance for ijr((x) = x2 [symmetric Gaussian homogeneous process] we have 0(h) =exp P.- f h(*)2d*T|, (5.5.24) 0<p<2 [symmetric Levy homogeneous process] ^(h)=exp £-JjA(*)|>cfoJ. (5.5.25) and for }Jr(oL)=\ot\P, 0<p<2 [symmetric Levy homogeneous process] we have Proof. By (5.5.20) and stochastic independence we have <j> (h)=E{e~*" Sfy-"(-*>} = n E{er 2™ V5^} 0=1 = fl e-<*A*> W=exp |- S ^W ^(A^ I, • 33=1 L 3>=1 J and this is the integral (5.5.23), as claimed. Remark 5. If X(A) is a &-vector, and correspondingly ^(a)s^(aV..,a*),
142 STOCHASTIC PROCESSES AND then (5.5.23) generalizes to <f>(h\ ...^EEexpT- f r/rih1®, ...,h*(t))dv\, (5.5.26) and if, for instance, f(a\...,(#)=* £ cMa%^0, (5.5.27) 3?,ff=l then 0(h\ ..., hk) = exp | - f Sc^h^t)h*(t)&?(*)|. (5.5.28) Remark 6. If all A's are individual points so that our function is a proper point function X(t), then A = (tv ..., tt), and (5.5.19) is a finite z discrete sum 2 ^-^fe) according to our general definition 5.4.3. If now T is the continuous interval 0 < t _; 1 say, then one would prefer having an integral ~ h(t)X(t)dt (5.5.29) instead, and we would like to point out that this can be so obtained by adjusting the stochastic process X(t) itself rather than modifying the definition 5.4.3 as such. In fact, if the integral (5.5.29) exists for such simple noncontinuous functions as are step functions on intervals Aa/S, then by putting h(t) = (oA(XJt) we arrive at random variables J. X(t)dt=Y(Aa/l). They constitute a random interval function, and for a step function h(t) the integral (5.5.29) can now be written as the Stieltjes integral Ch^dY according to our general definition. If, therefore, X(t) is such that 'averaged' values 7(A) are a fair substitute for it (more general random functions X(t) would hardly be sufficiently 'observable' at individual points t to be of interest stochastically) then the expression I jexp — 2iri E exp ~27ri h(t)X{t)dt may be claimed to be the characteristic functional for X(t) itself, indirectly.
CHARACTERISTIC FUNCTIONALS H3 5.6. Generating functionals If a joint distribution function FX(A) of random variables X1,...,Xl is 0 outside the closure of a proper cone Ch then for (zl3 ...,«j) in the (closed dual) cone Zl (or more generally for complex fa) whose real parts lie in Zt) we can introduce the completely monotone function J Ci called the generating function of Fl9 which by its real values already determines Fu and which for the complex values zj=2mocj is the characteristic function itself. In particular, if Xv ...,XZ are all ^0 (as they are deemed to be in 'population' problems), then Cz is the standard octant a;1>0,...,#j>0, but we do not demand that our octants be so normalized. If in the case of a Euclidean stochastic process each Fx is 0 outside a closed cone Ol9 and if for A </i we have then we also have */*(j7^a(«)) = &(«)• For the points a) in a certain c cone' of the space & we can then introduce a certain function 0(a)) having the form inerating functional. For a ranc <D(h)=^[exp(^27rj h{t)dx\\, 'Oa and we term it a generating functional. For a random function X(A) it has the form naturally. Theorem 4.2.3 implies the following one: Theorem 5.6.1. If x(£x,..., gk) is a completely monotone function in a proper cone Ck, and is a completely monotone mapping from another Cz into Ok) then X(c1,c^...,ck) is a generating function for any (cv ..., ck) in Ck, and if'x(£j) ^s bounded in Ok, then (fy) may also lie on the boundary of Ck. Another, perhaps more important criterion based on it, is as follows.
144 STOCHASTIC PROCESSES AND Theorem 5.6.2. Given the data of theorem 5.6.1, if £=A(7/1?..., rjj) is a further completely monotone mapping from Cl to 0<£<oo, and (yv ..., yk) is also in Gk, then the function jptlXfa + fa-'-'Ck + irti-Xfa^ (5.6.1) is completely monotone and bounded in Olt and thus a generating function except for a numerical factor. Proof. If we introduce the representation *&,...,&)= f e-<«ifi+-+"*Wi/)(tt), then the function (5.6.1) is i^"*L-Aw—J^* (5-6-2) and (ii, y) is ^ 0. If (w, y) = 0 then the bracket is — 1, and thus completely monotone, and if (u, y) > 0, then the bracket is (u,y). J e-teVMiDdt, and thus completely monotone. Also, e-(M><H-^) = e-(w,e)e-(w,^) jg com pletely monotone for u in 27ft, and thus the entire function (5.6.2) is so. Finally, for % =... =^=0, (5.6.2) has at most the value /. f-c e-(Ulcx+...-i-ukck) (Uiyi + % tm +ukyk) dp(u), ' Uk dy I which is — S.-y,---- and thus finite for c € Oft. By the use of theorem 5.6.2 it can be shown that for any completely monotone function #(t) in 0 ^ t < oo with #(0 +) = 1 and #(+oo)=0 the expression for a(t)^0 and 0 < 0(t) < 60 < I is the generating functional of an interval function X(Aa/?) on the intervals (0 ^ )oc < t <> fi( < oo), and for 0(t) = const, it was introduced (in its finite dimensional versions) by J. Neyman in a study on accident proneness.
145 CHAPTER 6 ANALYSIS OF STOCHASTIC PROCESSES 6.1. Basic operations with characteristic functionals To operate with characteristic functionals on a given dual space & means to generate a functional <fi(co) from given data, especially from other functionals, and by theorem 5.4.5 this amounts to constructing a family of characteristic functions {fixfav)} f°r which the consistency relations &(«*)-i*,Ma>> = */•(&) (6.1.1) have to be verified, but also nothing else. For instance, it follows N immediately that a product JJ ftn){&) of characteristic functionals, w=l or a sum n n .Er»0(n)(&), y»so, Ey»=i, w=l n=l or Km 0(n)(£>)j is again a functional, and if, for instance, on the same {D, A'} we are given two functions X1(A), X2(A) with random values in the same probability space, and if the two sets of random variables {Z1(A)}, {X2(A)} are stochastically independent (definition 3.7.6), then for the sum function XX(A)+X2(A) the characteristic functional is the product of the two given ones. Definition 6.1.1. We say that a Euclidean stochastic process is symmetric, or Gaussian, or Poisson, etc, if each (fi^tty) is so. Thus, for instance, a Gaussian process is one for which <f>(aj) = exp (iQx(&) - Q2(o>)), (6.1.2) where QX(S) is a real distributive functional on Q and Q-2{(o) is a real non-negative symmetric bi-distributive functional on QxiQ; and if Q-jju))=0, then the process is symmetric Gaussian. We say that the process is Gauss-Poisson if for each A we have Q\(ctP) = e~^ap\ where ^A(a)=^ + ^+^ (6.L3) as in (3.4.13). Note that if a function X(A) is a homogeneous process as in definition 5.5.4, then it is a particular case of a Gauss-Poisson process as just denned, but only a rather particular case of it.
146 ANALYSIS OF STOCHASTIC PROCESSES Relations (6.1.1) are equivalent with and if we introduce new functions then (6.1.4) implies ^A(a) = ^(/O, and hence the following conclusion: Theorem 6.1.1. The general rules of subordination (theorem 4.3.1 and others) also apply to characteristic Junctionals, so that in particular no symmetric Gaussian process on U is subordinate to any Oauss-Poisson process. Definition 6.1.2. For any admissible family D, a complex function K(A, A') on D x D is a positive-definite kernel if for any complex step function (5.5.16) we have S M«*(A*A)sf f Ht)W)d2K(t,T)^o] »,«-i JtJt y (6.1.5) Z(A',A)=Z(A^7). J We call it a Hilhert kernel if it is an inner product 2T(A,A')s(Z(A),X(A')) for an (additive) function X(A) with values in some (perhaps non- separable) Hubert space, and we call it a covariance kernel if the Hilbert space can be chosen as an L2-space over a probability measure space, that is, « Z(A,A0-^{X(A)X(A')}= X(A,a))X(k',G))dP(G)). Let X(A) be a Gaussian process with characteristic functional exp( — Q2((b)). If (0 is a step function (5.5.16), then, except for the factor 4:7T2, Q2(w) is == S A,*4Z(AtJ,Ae). (6-1.6) where Z(A, A') =E{X(A) X(A% Also in the case of a vector function X,(A), j=l, ...,k, (6.1.6) is to be replaced by 2 ^A/^A.A'), where Z,,(A, A') = JB{X,(A) X,(A')},
ANALYSIS OF STOCHASTIC PROCESSES HI and in particular for k=2 we obtain for a complex function Z(A)=X1(A)+iZ2(A) the expression 2hphqK(A99 Aff) 2>, a with a complex kernel K(A}A') = E{X(&)X(A% But, as on Ez so also on Q, an arbitrary bi-distributive function Q2(w) on L0 gives rise to a Gaussian process because the consistency relations (6.L4) are then obviously fulfilled, and hence the following conclusion: Theorem 6.L2. For any admissible D, any positive-definite kernel is a covariance kernel, and in particular any Hilbert kernel is a covariance kernel. We now take in Et: (<zp) a function ^(oc^) which in the neighborhood of the origin (a) = (0) can be written, for some t^ 1, as a sum ^K) = 1+P'K)H-P^), (6.1.7) where P^aJ is a polynomial Pvog=- 2 ani...ni(»27rix1)»i...(-2riaI)»i, (6.1.8) and Br{ocP) is of smaller order at the origin, #"(0^ = 0(1 a |'), |a|->0. (6.1.9) By taking the logarithm of (6.1.7) we can obviously write ^K)=exp(P-K)+^K)), (6.1.10) where again Sr(aS)) = o(\a\r), and (6.1.10) conversely implies (6.1.7). The representation (6.1.10) is unique. Otherwise we would have a relati0n -4K)=.BK), (6.1.11) in which B(ocP) = o(\ cc \r) and A{a9) is again a polynomial (6.1.8) and not =e0. For some s^rwe could then put p=8 where each Cp(ci^) is a homogeneous polynomial of degree s and Cs{<x^)^k0. However, if we introduce spherical coordinates ocP=jSP | a J, /?*+...+A2=1, and divide (6.1.11) by \oc\s and let a->0, then we obtain C8(jSJf) = 09 which would be contradictory.
148 ANALYSIS OF STOCHASTIC PBOCESSES If in a stochastic process each characteristic function is of the form ^AK)=exp(P^)+^AK)) for the same r, then (6.1.1) implies and because of the uniqueness just established this implies the relations P5<«HP;M«».j ^(*)=^(Ma))>J separately. In fact, if Pffl) is a polynomial of degree r m the /?'s, then P^(g/AA(a)) is such a polynomial in the a's, and Sffi) =-o(| /? |r) as /?->0 implies S^g^cc)) =-o(| a |r) as a~>0. As we have seen in section 3.5, if r=2m then we have a representation (6.1.7) if and only if the distribution function F(A) pertaining to ^(a,p) has moments of order ^ 2m, and we then have in fact .xfidF(xp). For r=2rn +1 the connection is not so direct, but in the next theorem we will denote the coefficients ani...Wz as moments nevertheless. The first moments a0...i...o which are usually called means and denoted by (mx, ...,mz) can always be 'removed', that is, made equal to zero, by envisaging the translated set function F(A — m) instead of the original F(A). Hence the following theorem: Theorem 6.1.3. (i) In any Euclidean stochastic process, especially in any random function X(A), the means, if existing, can be made equal to zero in all distribution functions {FV(A)} simultaneously, and the moments of any one order have values connected with each other by relations (6.1.12). (ii) // the process has (first and) second moments then there is another process (with the same consistency transformations g^{(x)) which is Gaussian and has the same first and second moments as the given one. Part (ii) of the theorem is the most comprehensive precise version known to us of the proposition that if in a 'stochastic process' (however this term be defined) only first and second moments of joint distribution functions are being envisaged, then any conclusion that can be drawn for the particular case of Gaussian processes will also be valid for processes other than Gaussian. Finally, if a stochastic process is Gauss-Poisson, then the following
ANALYSIS OF STOCHASTIC PROCESSES 149 precise decomposition statement can be made even though the second moments need not exist any longer. Theorem 6.1.4. If in a Gauss-Poisson process we put (6.1.3) in the Mm &(*,) = exp Wf(a9) + itfi(ap) +*#[(«,)], (6.1.13) as in section 4.3, then the Gaussian characteristic functions exp[^K)] (6.1.14) and the Bemoulli-Poisson functions exp[^f(a,)]. exVW*(^)+ifl(ocP)] (6.1.15) separately constitute a stochastic process each. Proof. The relation nw+i$M+iidw=W)+w+W) splits into the real and imaginary parts ^(a)+^f(a) = W) + ^), (6-1-16) and (6.1.16) is the specific relation 2 4aa3>a<z+ /sin(a,a))2dFz(;r) = S«fiAA+ i$m{P,y))HFm{y). r,s=l j£^ (6.1.18) Now, the second integral is j /sin/ SmI'^M- (6-1-18) m where ^^ Sc»ay<r> and the latter is a transformation from E'm to E[. By using its inverse we can define a set function F[(A) in E[ such that (6.1.19) is identical with r (sin(a.?/))W;(7/), J Ei but by the uniqueness assertion in theorem 3.4.2, relation (6.1.18) now splits into ^a) = ^y^ and also ^f(a) = ^(/?). If we combine the latter with (6.1.17) we obtain ») + ifi{x) -= ^(/?) + ifiifi), and this completes the proof of the theorem.
ISO ANALYSIS OF STOCHASTIC PROCESSES 6.2. Convergence in probability and integration Convergence in probability was introduced in definition 3.7.5, but for the present purposes another definition, known to be equivalent to it, will be more appropriate. Definition 6.2.L On a probability space {O; SP\ P) a sequence of random variables xn=fn(a))> each determined only a.e., is said to converge in probability (in measure) to a random variable x—f(a)), in symbols xn->x (prob), if for each e > 0 we have \imP(\fn(a))-f(a>)\>e)=0, (6.2.1) w->oo and a sequence {xn} is a Cauchy sequence (prob) if lim P(|/„(a>)-/n(«) | >e)=0 (6.2.2) for each e>0. m'n The following properties of limits in probability will be taken as known: Lemma 6.2.1. A sequence is convergent towards some limit element if (and only if) it is a Cauchy sequence, and the limit is unique a.e. A limit of limits of a set of random variables is again a limit of the set. If xn->x>yn~*y and ot,fi are real, then axn+fiyn-*ax+fiy. Theobem 6.2.1. For a Euclidean stochastic process, if we are given a sequence {o)n} in D, then the corresponding sequence of random variables Xn={u,K) (6.2.3) on Q is a Cauchy sequence (prob) if and only if for the characteristic functional we have ]im fta@m - con)) = 1 (6.2.4) for every real cc, and the limit (prob) of (6.2.3) is x0= (o), o)0) for co0 in & ifandonlyif ]jm (a(o)n-o)0))^l. (6.2.5) Proof. It will suffice to prove the first part of the theorem. If c(x) is bounded continuous in ( —oo,co), then, as is easily seen, for any Cauchy sequence we have Km f c(fwH-/wH)dPH = c(0), m, n-^-co J CI and on putting xn=fn(a)) and c(x) = e~27ria(c we obtain (6.2.4). Conversely if, for instance, in -< e-na^wiaxfa
ANALYSIS OF STOCHASTIC PROCESSES *5i we put x=(coi com) — (a), a)n) and integrate over Q we obtain E{e~* I am-ani2} = | e-™*<f>(ot{com - coj) da, J —oo so that^(6.2.4) implies f , m, n->oo But this implies (6.2.2), as claimed. Theorem 6.2.2. Let O be an everywhere dense part of a larger vector space Q,*: (a)*) on which there is given a topology in which aa)*+/?S* is continuous in the four quantities shown, and let Q* be complete in the sense that a sequence {fi*} has a limit point if and only if we have lim (6^-fi*) = 0 {in topology). (6.2.6) If a characteristic functional (j>{o))onilis continuous in this topology, then there exists a distributive operation (o)3 5*) from U* to the vector space of random variables on Q, which for &* = 6 reduces to the original expression (a), a>) and which is such that o>*->5* {in topology) (6.2.7) implies (M,&t)-> {<*>,&%) (prob). (6.2.8) Proof. Every S* is the limit of a sequence {5n} in Q, and by theorem 6.2.1 the corresponding sequence (6.2.3) has a limit (prob). Also, by a property stated in lemma 6.2.1, this limit depends only on 6* itself and not on the approximating sequence chosen, and thus may be denoted by (co, a>*). The properties claimed for the latter entity may now be obtained by full use of lemma 6.2.1. We now take an admissible family D={A} on a space T, and denote by D*={A*} the smallest <r-field of sets on T containing it, and we assume that on D* there is given a Lebesgue measure i?(A*) which for the elements of D is both, finite and 4=0. For p > 0, the measurable functions h{t) for which |Yf \h(t)\Pdv\1,P if l^p<oo if 0<pSl (6.2.9) is finite are determined almost everywhere, and they constitute a vector space LpDL0. Also, OXhi — h2) has the properties of a metric on Lfl for which the vector operations are continuous and the space is
152 ANALYSIS OF STOCHASTIC PBOCESSES I complete (even for 0 < p ^ 1, which is rarely so stated); and for tl=L0 the space Lp is a suitable extension H*. Hence the following definition and theorem: Definition 6.2.2. We say that a random function X{A) is Lp-finite if <j>{h) on LQ is continuous in the L^-metric. Note that if v{T) < oo (thus, for instance, if I7 is a bounded interval), then any function which is L^-finite is also L^-finite for p'>p. Theokem 6.2.3. // X(A) on D is Lp-finitefor some p>0 then, by limits in probability, the integral h(t)dX(t) (6.2.10) f t can be extended from h in L0 to h in Lp, and the i extended' integral is continuous (prob) relative to the Lp-metric. In particular, by putting h(t) = &>A*(t), X(A) can be extended to all sets A* for which v(A*) < oo, and the extension is o--additive {prob). If 0(h) is of the form <b([ \h{t)\PdvA, (6.2.11) where <£>(£) is defined and continuous in 0 ^ £ < oo, then it is obviously Z^-finite, and therefore the two processes 0(h)=exp Uj | h(t) |*d^, (6.2.12) 0(A)=exp(-(f \h(t)\*dv\*P\ (6.2.13) are both Infinite. But of ^h) = exp(-f \h(t)\PdvA, (6.2.M) we can only say that it is ^-finite even though it looks very similar to (6.2.13). But (6.2.14) is a homogeneous Levy process (see definition 5.5.4) just as (6.2.12) is homogeneous Gaussian; (6.2.13), however, is not homogeneous, although it is subordinate to (6.2.12). Remark. It is easily seen that every homogeneous symmetric Gaussian X{A) must be of the form (6.2.12), if X(A) is 'scalar'. But if it is a vector [Xl(A),..., X*(A)], then the general form of its characteristic functional is ${h\..., A»)«exp (- f S hv(t)h«(t)dvm(t)\, (6.2.15) where ^(A) is a (nonpositive) real additive function on D such that 2 h*hqvm(A) is positive semidefinite for each A.
ANALYSIS OF STOCHASTIC PROCESSES 153 In general, for k=2, we can introduce complex quantities X(A), h(t) as in remark of section 5.5. Now, if the sequence J. is convergent (prob) for every sequence {tin, h^} which is convergent in Z^-metric, then so is also the sequence r [KdX'{t)-h'JX"{t)l l hence the following statement: Theorem 6.2.4. If a complex random function Z(A) is Lp-finite, then the integrals (" h(t)dX(t) (6.216) and h(t) dX(t) have the completeness and closure properties stated in theorem 6.2.3. Deeinitton 6.2.3. We will call a complex random function X(A) a Wiener process if its characteristic functional is 0(h) = exp (- f I h(t) \*dv\. (6.2.17) If written in real components, a Wiener process is a Gaussian vector [XX(A),X2(A)] with characteristic functional ffl- l,h*) = exp(-[ ((&)* +(h*)*)dv\, (6.2.18) and this is a special case of (6.2.15) for &=2. Henceforth all functions X(A) will be complex. 6.3. Convergence in norm If for given ^^lwe consider on the given probability space only those random variables x—f(a)) for which the LP(Q)-norm \\f\\=(Ja\M\»iP(<»)J" (6-3.1) is finite, then we can envisage convergence in norm, ||/w— /|| ~>0, or /»-*/(£,)> (6-3-2) and we note the following statement which is easily proven: Lemma 6.3.L If fn->f(L9) then /„-*/ (prob); but not necessarily conversely.
i54 ANALYSIS OP STOCHASTIC PROCESSES Definition 6.3.L We say that X(A) is LP)i)-bounded if Z(A) is in LJO), meaning that r . * Y=\ h(t)dX{t) (6.3.3) is in L^Q) for h eL0, and if there is a constant M such that /' \h(t)\Pdv^l impfies E{\Y\*}£M. (6.3.4) Lemma 6.3.1 implies as follows: Theorem 6.3.1. For any p>0, if X(A) is Lp> ^bounded for some p*>l9it is also Lp-finite. If a process is of the form (6.2.12) or (6.2.13) or (6.2.14) then the (one-dimensional) characteristic function of the random variable (6.3.3) is of the form exp (-<*»(" \h\2dv), exp(-|a|^|| \h\*dv\ j, exp(-|a|/>| \h\Pdv\ respectively, and with the aid of theorem 3.5.3 we may state as follows: Theorem 6.3.2. A Wiener process is L2 ^bounded for every p ^ 1 and is in particular L22-bounded. The process (6.2.13), which is subordinate to a Wiener process, is L2> ^bounded for p<pm, and the homogeneous Levy process (6.2.14) is Lp> ^-bounded for p<p. We now start out not with a set function X(A) but with a (random) point function x(t) on T with values in Lp(0), and we assume that this vector-valued function x(t) is (strongly) measurable in v(A*) as in section 1.3, and that for some cr< 1 the integral l^^Wxm-dvJ^ (6.3.5) is finite, thus being a Banach norm itself. The number cr need have no relation to the number^) in (6.3.1) and in fact the norm (6.3.5) could be introduced for x in some Banach space B not necessarily an Lp(0), and in order to emphasize this we will denote the Banach space of functions {x(t)} with norm (6.3.5) by La(B) even for a given p. We will also consider the norm (6.3.5) for cr—oo defining it then by HI x HI == essential upper bound of || x(t) ||, a? i
ANALYSIS OF STOCHASTIC PROCESSES 155 so that we can introduce the quotient f=~l (6-3-6) for all p ^ 1. p = 1 included. By the theory of integration of vector-valued functions, if h(t)eLp(T) and x{^Lv{B)9 then the integral y= Mt) x(t) dv(t) (6.3.7) exists (and is an element of B—L^Q) again), and, in fact, by Holder's inequality we have , * ,ljp \\yU{jT\h(t)\Pdv) -\\\z\\\. (6.3.8) In particular, h(t) = o)A(t) eLp(T), so that the integral f a)A{t)x(t)dv=\ x(t)dv==X{/±) (6.3.9) exists and is an element of LP(Q), and as a function on D it is additive and in fact cr-additive (LJ. For heL0, the integral (6.3.7) can now be equated with the integral (6.3.3) and hence the following statement: Theorem 6.3.3. If X(A) is an indefinite integral (6.3.9), then it is Lp> ^bounded. This theorem does not have a converse; and we will see that even the Wiener process, which is the most 'bounded' process known, is not an indefinite integral if the variable t is continuous. Definition 6.3.2. We say that the measure v(A) is Lratomistic if for every e > 0 there is a 8 > 0 such that 2 v{Am) S 8 implies 2 v(Aw)* ^ e for any sum A-l+...-1-Aj. Theorem 6.3.4. J/X(A) is a Wiener process, then X(A) is the integral (6.3.9) of a function x(t) with values in L2{&) if and only if v(&) is L2-atomistic. Proof. For (6.3.3) we have E{\ Y\2} = c4 \h{t)\Hv for a constant c>0, and hence 2 hx(aj n= s m x(aj i2))4=c i. 1 »(a j 1*, m=l m=l m—1 so that X(A) is 'absolutely continuous' in L2(Q)-norm if and only if -y(A) is Lg-atomistic, as claimed.
i56 ANALYSIS OF STOCHASTIC PROCESSES I Ordinary measure on a < t < bis not ^-atomistic, and thus' ordinary' Wiener process is not an indefinite integral of a function x(t) with values x in L2(Q). If we apply definition 6.1.1 to a complex function X(A), then it is a Gaussian process if its characteristic function is of the form erpWiW-GaW). (6.3.10) where Qx{h) is real distributive and Q2(h) is non-negative Hermitian bi-distributive on the complex L0. Theorem 6.3.5. A Gaussian process X(A) is L22-bounded if and only if Qx(h) and Q2{h) are both bounded for \h{t)\*dv£l, (6.3.11) that is, if, and only if, on the complex L2(T), Q^h) is a real linear functional and Q2(h) = (Ah, h) where Ah is a non-negative bounded self- adjoint operator. Proof. The characteristic function of (6.3.3) is exp {iocQ^h) -a2#2(h)), and thus, except for a constant factor, ^{|F|}«=t2,(A)+i(«1(i))». Rut this is bounded in the unit sphere (6.3.11) if, and only if, Q^h), Q2(h) are bounded separately, as claimed. 6.4. Expansion in series and integrals We assume that there exists on {D*; v(A*)} a complete orthonormal SJStem fa®}* »=l,2,3y..., (6.4.1) with J gm(t)gjt)dv~dmn. (6.4.2) For any h(t) eL2(T) we can introduce the expansion ft(*)~Sy«?*W, r™= f Mt)JJt)dv, (6.4.3) 1 J T n and we have h,(t)=lim Sr»r7m(*)(ia), (6.4.4) by the Riesz-Fischer theorem. If for some /> > 1 the functions gn(t) belong to both Lp and Lp (p_1)}
157 ANALYSIS OF STOCHASTIC PROCESSES then the series (6.4.3) can also be set up for heLp(T), and it then frequently happens that there exists a matrix of real numbers {Km}, m,n,= l,2.35..., in which for each n only finitely many are =j= 0 such that 00 h(t)=]im 1lXnmyngm{t){Lp). (6.4.5) n—>co m=1 For instance, for ordinary periodic functions h(t) in — J^t<|, if we put oo /»£ HQ^lZyme27™, 7m= h(t)e*«imidtt (6.4.6) then, as we know, we have A(*)= lim £ (l-^)rmeto'"'(£p), (6.4.7) w-»oo — n \ n J for h(t) eL,p^l; and, in fact, by a renowned theorem of Marcel Riesz we have n h(t)=lim S yme2^(^)5 (6.4.8) n-^-oo —n not only for p = 2 but also for p > 1, although this is of no particular consequence to us. Now, if we substitute all these expansions into our integral (6.3.3) we obtain the following theorem: Theorem 6.4.1. (i) If for an Lp-finite X(A) we introduce the {random- valued) Fourier coefficients -J. 7m= 9m(t)dX(t), (6.4.9) then we have ^P ~ S Tmgm(fi (6.4.10) all m=1 in the sense that for heL.we have I h(t)dX(t)= lim S KmfmYm (prob), (6.4.11) T n->oo m=1 <tn<Z = lim 2 ymYm (prob) (6.4.12) w-^oo w.—1 for heL2. (ii) In particular, if X(A) is periodic on — J ^ t < |, tfhen we have dX(A) dA ~Ii7m&"imt> Ym=( e~*"imtdX(t) (6.4.13) -oo J — £
i5§ ANALYSIS OF STOCHASTIC PROCESSES in the sense that f* W)dX(t)= lim £ UJj!tl\ymYm (prob) (6.4.14) J -i n->oo77i--n\ ^ / n forp^l,and = lim S ywFm (prob) (6.4.15) w-»co m= — w /or/)>L (iii) If, more precisely, X(A) is Lp ^-bounded, then the limits (6.4.11), (6.4.12), (6.4.14) and (6.4.15) exist not only in probability but also in Lv(Q)-norm. We are now viewing the sequence of random variables {Ym}, as given by (6.4.9), as a random-valued function on the 'discrete' set T'={m}, and we assign to each point m the measure 1. The analogue to the previous vector space L0 will be denoted by L'0, and it consists of sequences y—{ym} in which only finitely many components are 4= 0, and the space Lp) which is meant to be the analogue to Lp, is determined by the norm for/)^l. Theorem 6.4.2. (i) If {X(A)} is L2-finite, then {Ym} is L2-finite; and conversely. Similarly for L2:p-boundedness. (ii) if {X(A)} is L22-bounded Gaussian, then {Ym} is L22-bounded Gaussian; and conversely. (iii) If {X(A)} is a Wiener process, then {Ym} is a Wiener process; and conversely. (However, if a non-Gaussian process {X(A)} is homogeneous, then {Ym} need not be homogeneous; nor conversely.) Proof. If {X(A)} is L2-finite, and if with {ym} in L'0 we form »0 = 2ymPm(*)» (6.4.16) m then we obtain | W)dX(t)^^ymYm, (6.4.17) J T m and by Parseval's equation we have f |M*)la*>-2.|y»l*. (6-4.18) J T m If now we take a sequence of elements {yJJ in L'0i r= 1,2,..., and if we apply these formulas to the differences ym=yrm—y8m, and if we let r, s -> oo, then we conclude that { Ym} is also Lg-finite. Conversely, if {Ym} is L^-finite, anc* iffor heLQ we introduce the expansion (6.4.3), then (6.4.12) implies (6.4.17), and if we now take / *> \VP 2 17ml' \t»«l /
ANALYSIS OF STOCHASTIC PROCESSES 159 a sequence of elements {hr} in L0 and apply this and (6.4.18) to the differences h = hr—hs, r, s->oo, then {X(A)} turns out to be L2-finite. Also, by the same reasoning, if either process is L2> ^-bounded then so is the other. Next, if {X(A)} is Gaussian, then its characteristic functional #jexp(i| J^dX(m\ (6.4.19) has the form (6.3.10). But by (6.4.15), (6.4.19) is identical with ^{exp(ipmFTO)J, (6.4.20) which is the characteristic functional on {Ym}, and on the other hand the substitution (6.4.15) also transforms (6.3.10) into exp(i<2'(y)-Q"(y)), (6.4.21) where again Q'(y) is real continuous on L'2 and Q"(y) is non-negative bounded self-adjoint there. Thus {Ym} is also Gaussian. Finally, the identity (6.4.18) shows that if either process is a Wiener process then so is the other. Remark. The above reasoning also gives the following conclusion. If {X(A)} is subordinate to an Z2> ^bounded Gaussian process, thus having a characteristic functional of the form ®(»GiW-«aW), (6.4.22) with 'bounded' Qx(h) and Q2(h)t then{Fw}is of thesamekind and has 'the same' characteristic functional; and conversely. We are now turning to Fourier integrals, and the following theorem, although insufficient for important applications, is a centerpiece of the L2-theory: Theorem 6.4.3. IfX(A) is defined and L%%-bounded on Ex: ( —oo<t<oo), then there exists a (unique) other process Y(A), likewise defined and L2t ^-bounded on EX) such that we have r h(i)dX(r)=r g^Ody(s) (6.4.23) J —00 J —00 for any h(r), g(s) eL2(Ex) which are Plancherel transforms h(r) ~ I e27rirs g(s) ds9 g(s) ~ | e~27risrh(r) dr (6.4.24) J —00 J —00 of each other.
i6o ANALYSIS OF STOCHASTIC PROCESSES In particular, by putting g(s) — o)A(s), we have poo f ffi \ f00 e—27rirJ2 Q—2irirct r(A°H - (J.^*) dX{rn - -2m dX{r) and inversely (or'dually'), Z(Aa/;)=J _m —. dY(s). (6.4.26) Also, if either process is Gaussian or Wiener, then so is the other, and if the characteristic functional of X(A) has the form (6.4.22), then the characteristic functional of Y(A) arises by insertion of the first integral (6.4.24) into it. Proof, If X(A) is Zr22-bounded onEx, then J —o h(r)dX(r) (6.4.27) is a bounded linear operation from L^E^ to L2(Q), and it is readily seen that any such operation can be so represented by a function X(A), and uniquely so. Now the Plancherel transformation (6.4.24) is a unitary mapping of L2(EX) into itself, and such a mapping carries a bounded linear operation from L2(EX) to L2(Q) into another such one, whence the theorem. 6.5, Stationarity and orthogonality Joint assumption. In the present section every process occurring is L2j2-bounded on its entire space of existence. Definition 6.5.1. We say that X(A) on ( — oo,oo) is JT-stationary (eE9 for Khintchine), if for any two finite intervals Am: am<r£fim, m—1,2, and all its translates A^: am +1 < r ^ fim+1, we have E{X(A{).X(^}^E{X(A1).'X(A^} (6.5.1) Lemma 6.5.1. X(A) is K-stationary if and only if the bilinear functional P(h1(-),h*(-))=Eir hHi)dX(r).r h\s)dX&\, (6.5.2) h1,h2eLz, is commutative with translations, P(W(r+t), h*(s+t))=-P(h\r), h*(s)), (6.5.3) — oo<t<oo. Also, if X(A) happens to be the indefinite integral (6.3.9) of an Lrcontinuous function x(t) then X(A) is K-stationary if and only if for the covariance function R(r, s) = E{x(r)x~^)}, (6.5.4)
ANALYSIS OF STOCHASTIC PROCESSES 161 we have R(r+t3s+t) = R(r,s), so that we have R(r, s) =R(r—s) (6.5.5) for a (continuous) function R(t). The proof of the first part follows easily if we approximate to h1, h2 by step-function, and of the second part if we substitute the integral (6.3.9) into condition (6.5.1). We also note without discussion (because it will be of no consequence to us) that if a Z-stationary X(A) is an integral (6.3.9) of a function x(t) which is integrable in L2(Q)-norm, then, after adjustment on a t-set of measure zero, x(t) is continuous in L2(0)-norm as envisaged in the second half of lemma 6.5.1. Debtnition 6.5.2. We call 7(A) on ( — 00,00) orthogonal if we have £7{7(A1).7(A2y} = 0 (6.5.6) for any two intervals which are disjoint, A1nA2=0. We call it orthonormal if, furthermore, E{\Y(A)\2} = c\A\, (6.5.7) where (A)=/? — a is the length of the interval, and c>0 is constant. Remarh 1. Requirements (6.5.6) and (6.5.7) can be contracted into E{Y(AX). 7(A2)} = c I A2 n A21. (6.5.8) Also, 7(A) is orthonormal if and only if it is both orthogonal and iT-stationary. Theorem 6.5.1. If X(A) and 7(A) are transforms as in theorem 6.4.3, then either of them is K-stationary if and only if the other is orthogonal. In particular, therefore, if either of them is orthonormal then so is also the other. Remark 2. Note that ' symmetry \ Gaussian character and ortho- normality are each 'invariant' under Fourier transformation, and a Wiener process is all three at the same time. Proof. Assume first that 7(A) is orthogonal. For Axn A2 = 0, we then have E{] y^^ ]2}==E{1 T{^ |2}+^{| 7(Af) ]% and thus if we introduce T(A)=E{\Y(A)\2}, (6.5.9) then this is a non-negative additive interval function which, however, is defined for finite intervals only. It is important to note for applications in other contexts that thus far we have only used the assump- 11 BHA
162 ANALYSIS OF STOCHASTIC PROCESSES tions that each F(A) is an element of L2(Q), and that the orthogonality property (6.5.6) is available. However, from full L2j2-boundedness we deduce r(A)=#j|P° uA(r)dY(r)f}^Mr \ a)A(r) \*dr = M. | A |, and by Lebesgue theory there exists now a bounded measurable function^), 0Sx(«)SJf, such that r(ArjS) = #(a) doc. We now introduce the bilinear functional (6.5.10) (6.5.11) QW-),g{-))=E[j"_n?(^ (6.5.12) g1, g2 € L2, and if g1 and g2 are both step functions on the same sum of intervals Ax +... + Au and if gf are the values of gm(r) on A3-; m = 1,2; j = l, ...,l; then (6.5.12) has the value £ ^fr(A,). Now, this expression is nothing but the integral /: and by L22-boundedness we obtain that this is the value of (6.5.12) for g\g2 in general. From this we conclude that (6.5.12) has the invariance property W(r), 9%$)) s Q(?(r) e^\ g2(r) e*"^) (6.5.14) for all t; and if we introduce the Plancherel transforms (6.4.22) for both functions g\ g2, then (6.5.14) goes over into the relation (6.5.3), meaning that X(A) is ^"-stationary as claimed. Conversely, assume that X(A) is JT-stationary. Since, for h1=A2 =h, P(h, h) is real and' bounded', it follows from Hilbert space theory that there exists a bounded self-adjoint operator Sh such that P(hi5h2) = (h2s^i)j (6.5.15) where the expression on the right is an 'inner product'. If we denote by 17% the operation which carries A(') into h(*+t), then the stationarity assumption (6.5.3) can be stated as (h2^h1) = (U*h2,^U*hi),
ANALYSIS OF STOCHASTIC PROCESSES 163 but the right side is also (h2, (Ut)*8U%1), so that we have £=(U*)*#U*. Thus, 8 is commutative with translations, and since 8 is bounded, our lemma 4.6.1 becomes applicable. Therefore, if we denote the Plancherel transforms of g\ g2 by h1, h2, then P(hx3 h2) has the value (6.5.13), where v(a) is a bounded measurable function associable with the operator 8. However, Q(g1ig2)=P(hl,h2), and thus Q(g\g2) is expressible as an integral (6.5.13), and if we put gm(oc) = o)Am(oc), m = l,2, then (6.5.13) has the value 0 for A-^ A2=0, which proves (6.5.Q), as claimed. Our lemma 4.6.1 has an analogue, more or less, in every situation in which a Plancherel duality theorem is available, and theorem 6.5.1 has then also an analogue of some kind, in 'noncommutative' cases as well. The following statement which will not be further discussed is typical of them all: Theorem 6.5.2. I/{Z(A)} is periodic on — J^£<|, and if 7m= f e-*"imtdX(t) (6.5.16) is the 'dual' process, then X(A) is stationary according to (the literally the same) definition 6.5.1, if and only if{Ym} is orthogonal in the sense that JE{YmYn}=0, m+n; and X(A) is orthogonal according to (the literally the same) definition 6.5.2 if and only if{Ym} is stationary in the sense that R(m,n) = E{YmYn} is a function ofn — m only, R(m, n) = B(m—n). Also, {X(A)} is orthonormal according to definition 6.5.2 if and only if{Ym} is orthonormal in the familiar sense that (usually with c—l). 6.6. Further statements Theorem 6.5.1 is not the familiar proposition due to EJiintchine and others. The latter does not introduce £2>2-boundedness or feature duality so prominently, but, in an indirect fashion it introduces L1>2-boundedness, and it is crucially based on our theorem 3.2.3 about 11-2
164 ANALYSIS OF STOCHASTIC PROCESSES positive-definite functions which it applies by way of the following proposition: Theorem 6.6.1. If a random point function x(t) is continuous in L2(Q)-ra?rm, and if we have E{x(r)~xJs)} = B(r-s), (6.6.1) then W3 can put R(r ~s)=\ e27Tilr-$«aT(a), (6.6.2) where T(A)Z0, r(E1)=i?(0)<oo. (6.6.3) For any hl( •), h2( •) e L-^E-j) we can form the bi-distributive functional P(hi,h2) = EJ p ¥(f)x{r)dr.r h2(s)^T)ds| (6.6.4) /•oo /*co = h1(r)h*(s)R(r-s)drds, (6.6.6) J — CO J — CO and if we introduce the transforms gm(a) = e~27ritahm(t)dt, m=l,2, (6.6.6) /*co /*co then we have P(h\h2) = gl(a)g2(cx)dT(a). (6.6.7) J —CO J —00 Proof. \\x(t) \\2=E{x(t)x(P)}=R(0) implies in particular that ||a;(t) || is bounded, and since x(t) is also continuous in norm, we can consider the integral /•» h(t)x(t)dt (6.6.8) for any h( •) eL^Ej), and it is even possible to define the integral /: x(t)dH(t) (6.6.9) as a Riemann integral for any H(*)e V(EX). We can also form (6.6 A) and since for he2^ we have 6^P(h,h)=r f° Mf)h(s)R(r~s)drds, we can indeed apply our theorem 3.2.3, and a representation (6.6.2) with (6.6.3) exists indeed. Also we can substitute (6.6.2) in (6.6.5) and (6.6.7) ensues, as claimed. Next, (6.6.8) is a distributive functional from Lx to L2(Q), and if we introduce the transforms g(a)=\ h(t)e-2nit«dt, (6.6.10)
ANALYSIS OF STOCHASTIC PBOCESSES 165 we may also view it as a distributive functional from the vector space {g( •)} to L2(Q). Now, formally such a functional can be represented as an integral g(a)dT(a) with 7(A) eL^Cl), so that we have I Mt)x{t)dt=\ l&)dY{<x), (6.6.11) J —00 J -co and if we denote the transforms of h\ h2 by g1, g2, then (6.6.4) is formally Q(g\g2)^B^ ^jHa)dY(a).^j^)dY(^, (6<6J2) just as in section 6-5. Rut formally this implies the orthogonality relation E{T(^) F(A2)} = r(Axn A2), (6.6.13) and the fact is that all this can be made rigorous and that the following continuation to the preceding theorem can be stated: Theorem 6.6.2. Also, there exists, on all bounded intervals, a finitely additive function Y(&) eL2(Q) for which (6.6.13) and (6.6.11) holds. Conversely, if we are given a function F(A) for which (6.6.13) with (6.6.3) holds then there exists a function x(t) as in theorem 6.6.1 for which (6.6.11) holds. We will not reproduce the proof. Bemarh 1. Relation (6.6.11) can be enlarged to the more self-dual one: P° x(t)dH(t)=r g&)dY(cc), (6.6.14) J-co J-co where H(')e V(E^), and g(a) is its transform. Remark 2. If a cr-additive F(A) is finite on bounded sets, and if there is a F(A) with property (6.6.13), and if we introduce the decomposition T(A)^T1(A) + T2(A) + T3(A) into discrete, singular-continuous and absolutely continuous addends, then F(A) can be written as F^A)* F2(A) + r3(A), where Fm(A) satisfies (6.6.13) for Ym, m = l,2,3. Hence the following further statement: Theoeem6.6.3. Finally, we can put uniquely x(t) =x1(t) + x2(t) +xB{t), where each xm(t) is K-stationary and has a transform Ym( A) as in remark 2 of section 6.6. Any cr-additive T(A)^0 with r(^)<oo (6.615)
i66 ANALYSIS OF STOCHASTIC PBOCESSES can be the covariance function of a function ]T(A), compare theorem 6.1.2, and thus, except for the restriction (6.6.15), our present theorem is much more comprehensive than was theorem 6.5.1 in which T(A) had to be absolutely continuous always. This being so, we will further note that even the restriction (6.6.15) can be removed, and that there is a proposition available, which can be stated in several versions, of which both of our previous theorems are special cases. Theorem 6.6.4. If {xe(t)} in 0<e<oo is a family of stationary random point functions as in theorems (6.6.1)-(6.6.3), and if for any e>0,7] > Owe have &+*(t)=\ yexp --(t-r)2 ^(r)dr, (6.6.16) then there exists a representation x*(t)~ e%7IiiTe~e7TT*dY(T) (6.6.17) J —00 by a joint orthogonal function Y(A), meaning that if we introduce the representations ^ a?(Q~ ea*«r dYe(r) (6.6.18) J —00 of theorem 6.6.2 then we have T*(&*fi)=\ e-^dY(r). (6.6.19) Proof. If we apply (6.6.11) with x(r)=&(r)9 Y{A)=Y*(A) and h(r) = -7- exp — (t—r)2 then we obtain V?7 L V J --exp -~-{t-r)2\xs(r)dr=\ e*"iiT-"iT*dY*(T), but since by explicit assumption the left side is e**i*rdYv+e{T)9 we therefore obtain 7^+e(Aa) = e^^dY^r) J oc for any tj > 0, e> 0. From this it is possible to conclude that we have e m e"(y+e)r2dYy+e{T)=\ ewer%dTe{T)9 l J OC J — o
ANALYSIS OF STOCHASTIC PROCESSES 167 for all 7] > 0, e > 0. and this means that the integral 7(A) =f fl^r^T) J cc is independent of e, so that (6.6.19) holds, which proves the theorem. Now, if we take a function x(t) to which theorems (6.6.1)-(6.6.3) apply, and if we form then this is a family to which theorem 6.6.4 applies, with 7(A) being the same as in theorem 6.6.2. Also x(t) is a limit in norm of xe(t) as e 10. On the other hand, if we take a set function X(A) as in theorem 6.5.1 and put then this is also a family to which theorem 6,6.4 applies with 7(A) being the same as in theorem 6.5.1. Furthermore, it is easy to show from L2j2-boundedness that Xe(Aa/?)==r &(t)dt J a converges in norm to X(Aay?) as e 10, and this theorem 6.6.4 does indeed include both previous ones, in a sense. Note that the function xe(t) is a random-valued solution of the equation d2xe(t) _ 8&(t) and that we could have also taken the 'harmonic' equation instead, in which case we would have had xe(t)~\" e2^T-2e7r,T'd7(T), but we will not pursue this topic any further.
i68 NOTES AND REFERENCES Chaptbe 1 1.1. General approximation theorems are also to be found in E. W. Hobson, Theory of Functions of a Real Variable, vol. 2 (Cambridge University Press, 1947). 1.2. Translation functions, after having been mentioned incidentally by H. Bohr, were introduced systematically in the second half of our paper, 'Beitrage zur Theorie der fast periodischen Funktionen. I', Math. Ann. vol. 96 (1926), pp. 119-47. For Stepanoff functions see, for instance, A. S. Besicovitch, Almost Periodic Functions (Cambridge University Press, 1932). 1.3. The Holder-Minkowski inequality here used is formula (6.13.9) on p. 148 in G. H. Hardy, J. E. Littlewood and J. Polya, Inequalities (Cambridge University Press, 1934). 1.4. The Lebesgue integral of Banach-valued functions was introduced in our paper, 'Integration von Funktionen, deren Werte die Elemente eines Vektorraumes sind', Fund. Math. vol. 20 (1933), pp. 262-76. See also T. H. Hildebrandt, 'Integration in abstract spaces5, Bull. Amer. Math. Soc. vol. 59 (1953), pp. 111-39, and pp. 40-52 in E. Hille, Functional Analysis and Semi- groups (Amer. Math. Soc, Providence, R.I., 1948). For additional Fourier analytic applications of the integral see S. Bochner and A. E. Taylor, 'Linear functionals on certain spaces of abstractly valued functions', Ann. Math. vol. 39 (1938), pp. 262-76, and R. P. Boas and S. Bochner, 'On a theorem of M. Riesz for Fourier series', J. Lond. Math. Soc. vol. 14 (1939), pp. 62-73. 1.5. For additive set functions in Euclidean space see E. J. McShane Integration (PrincetonUniversity Press, 1944); J. Radon, 'Theorie und Anwen- dungen der absolut additiven Mengen funktionen', S.JB. Ahad. Wiss. Wien, vol. 122 (1913), pp. 1293-439; Hans Hahn, Reelle Funktionen, I (Springer, Berlin, 1921); and S. Bochner, 'Monotone Funktionen, Stieltjessche Integrale und harrnonische Analyse', Math. Ann. vol. 10 (1933), pp. 378-410. Theorem 1.5.4 is due to A. Plessner, 'Eine Kennzeichnung der totalstetigen Funktionen', J. Reine Angew. Math. vol. 160 (1929), pp. 26-32. The extension to Haar measure on compact groups was given in our paper, 'Additive set functions on groups', Ann. Math. vol. 40 (1939), pp. 769-96 (theorem 15 on p. 789). Chaptbb 2 2.2. When a course of lectures on this topic was being given in the spring of 1953 at the Statistical Laboratory of the University of California, Berkeley, a remark made by Dr E. Parzen led to theorem 2.2.2 becoming as general as it is now, and a remark made by Dr H. Flanders led to theorem 4.1.3 becoming as general as it is now. For a discussion and proof of theorem 2.2.4 see S. Bochner and K. Chandra- sekharan, Fourier Transforms (Princeton University Press, 1949: Ann. Math. Studies, no. 19). 2.4. There are possibilities of extending the Poisson summation formula from sums over strict lattices to sums over more general point sets. For one variable some such statements were made in our paper, 'A generalization of Poisson's summation formula', Duke Math. J. vol. 6 (1940), pp. 229-34, and
NOTES AND REFERENCES for several variables the problem was mentioned in ' On spherical partial sums of multiple Fourier series', Rev. Ciena., Lima (1948), pp. 85-104. Formulas for Fourier integrals used in the text will be found in Vorlesungen tiber Fouriersche Integrate (Chelsea, New York, 1948). 2.5. For 'spherical summability' see our paper, 'Summation of multiple Fourier series by spherical means', Trans. Amer. Math. Soc. vol. 40 (1936), pp. 175-207 and also chapter iv in the book by K. Chandrasekharan and S. Minakshisundaram, Typical Means (Oxford University Press, 1952: Tate Institute of Fundamental Research, Monographs on Mathematics and Physics). The diffusion equation with general completely monotone transformations of the Laplacean which we will yet discuss in section 4.6,?especially with fractional powers of the Laplacean and the link to symmetric stable processes, and also the subordination of Markoff processes which we will discuss in section 4.4, were introduced in our papers, 'Quasi-analytic functions, Laplace operator, positive kernels', Ann. Math. vol. 51 (1950), pp. 68-91 [this paper contains details and applications that will not be mentioned in the text], and ' Diffusion equation and stochastic processes', Proc. Nat. Acad. Sci., Wash., vol. 35 (1949), pp. 368-70. The general transformations of the Laplacean by themselves were already introduced in ' Completely monotone functions of the Laplace operator for torus and sphere', Duke Math. J. vol. 3 (1937), pp. 488-502. The case of fractional powers of the Laplacean in — oo <# <oo was linked to Riemann-Liouville integrals in W. Feller, 'On a generalization of Marcel Riesz's potentials and the semi-groups generated by them', Comm. Sim. Math. Univ. Lund, Tome supplemental (1952), pp.* 73-81. 2.6-2.8. These sections reproduce, with additions, the major part but not all of the contents of the following two papers of ours: 'Theta relations with spherical harmonics', Proc. Nat. Acad. Sci., Wash., vol. 37 (1951), pp. 804-8; 'Zeta functions and Green's functions for linear partial differential operators of elliptic type with constant coefficients', Ann. Math. vol. 57 (1953), pp. 32-56. Chapter 3 3.1-3.4. These sections are an elaboration of our note, 'Closure classes originating in the theory of probability', Proc. Nat. Acad. Sci.f Wash., vol. 39 (1953), pp. 1082-8. In — oo <jc<oo, the famed structure theorem 3.4.2 was conceived, with an imperfection, in A. Kolmogoroff, * Sulla forma generale di un processo stocastico omogeneo', R.C. Accad. Lincei, vol. 15 (6), 1932, pp. 805-8, 866-9. The imperfection was soon removed by Paul L6vy, and the theorem was variously reproven, and a systematic analysis of this (one-dimensional) theorem will be found in M. Loeve, 'On sets of probability laws and their limit elements', Univ. California Publ. Statist, vol. 1, no. 5 (1952), pp. 53-88. This paper is also, in a sense, concerned with the peculiar relationship of this theorem to the central limit theorem, and on this some brtef comment had already been made in W. Feller, 'The fundamental limit theorems in probability', Bull. Amer. Math. Soc. vol. 51 (1945), pp. 800-32. The structure theorem 3.4.2 for several space variables was set up in P. Levy, Thiorie de Vaddition des variables aliatoires (Gautheir-Villars, Paris, 1937), pp. 212-20. 3.5. For an extension of the theorems to several variables see our paper, 'Stochastic processes with finite and nonfinite variance', Proc. Nat. Acad. Sci., Wash., vol. 39 (1953), pp. 190-7.
170 NOTES AND REFERENCES For functions of slow growth see J. Karamata, 'Sur un mode de croissance reguliere des fonctions', Mathematica, Cluj, vol. 4 (1930), pp. 38-53. 3.8. See S. Bochner, 'Completely monotone functions in partially ordered spaces', Duke Math. J. vol. 9 (1942), pp. 519-26; S. Bochner and Ky Fan, 'Distributive order preserving operations in partially ordered vector spaces', Ann. Math. vol. 48 (1947), pp. 168-9, and the last chapter of the Ann. Math. Studies by E. J. McShane, Order Preserving Maps and Integration Processes (Princeton University Press, 1952). For the result of H. Cramer see his paper ' On the theory of stationary random processes', Ann. Math. vol. 41 (1940), pp. 215-30. Chapter 4 4.1. One half of theorem 4.1.2 was given in our paper 'Stable laws of probability and completely monotone functions', Duke Math. J. vol. 3 (1937), pp. 726-8, and we used it in effect to show that the symmetric stable distributions are subordinate to the Gaussian in the sense of section 4.3. The second half of theorem 4.1.2 was then stated, for another purpose, in I. J. Schoenberg, 'Metric spaces and completely monotone functions', Ann. Math. vol. 39 (1938), pp. 811-41. 4.2. Theorem 4.2.2 was obtained in a Princeton doctoral thesis by W. Gilbert, 1952. 4.4. The notion of continuity as introduced in definition 4.4.2 is perhaps ill named because it is quite different from the one introduced in A. Kolmogoroff, '"tTber die analytischen Methoden in der Wahrschemlichkeitsrechnung', Math. Ann. vol. 104 (1931), pp. 415-58, and in W. Feller, 'Zur theorie der stochasti- schen Prozesse', Math. Ann. vol. 113 (1936), pp. 113-60. This latter continuity means more or less that almost all paths are continuous, and this in turn means more or less that in the associated diffusion equation the operator in the space variables is a partial differential operator of second order and in the case of a stationary process even a 'genuine' Laplacean. We might mention that an important problem on diffusion was studied for fractional powers of the Laplacean in M. Kac, ' On some connections between probability theory and differential and integral equations', Proceedings of the Second Berkeley Symposium on Mathematical Statistics and Probability (University of California Press, 1951, pp. 189-215.) To theorem 4.4.4 compare T. J. Schoenberg, 'Metric spaces and positive definite functions', Trans. Amer. Math. Soc. vol. 44 (1938), pp. 522-36. 4.5. See R. E. A. C. Paley, 'A proof of a theorem on averages', Proc. Lond. Math. Soc. (2), vol. 31 (1930), pp. 289-300. 4.8 and 4.9. These sections summarize part of the results from the following papers of ours: 'Some properties of modular relations', Ann. Math. vol. 53 (1951), pp. 332-63, 'Connection between functional equations and modular relations, and functions of exponential type', J. Indian Math. Soc. vol. 16 (1952), pp. 99-102; 'Bessel functions and modular relations of higher type and hyperbolic differential equations', Comm. Sem. Math. Univ. Lund, Tome supplemental (1952), pp. 12-20. Chapter 5 The content of this and the next chapter is meant either to supplant or elaborate most, though not all, of the results in the following interlocking papers of ours: 'Stochastic processes', Ann. Math. vol. 48 (1947), pp. 1014-61;
NOTES AND REFERENCES 171 'Partial ordering in the theory of stochastic processes', Proc. Nat. Acad. Sci., Wash., vol. 36 (1950), pp. 439-43; 'Length of random paths in general homogeneous spaces', Ann. Math. vol. 57 (1953), pp. 309-13; 'Fourier transforms of time series', Proc. Nat. Acad. Sci., Wash., vol. 39 (1953), pp. 302-7. 5.1. For directed sets and inverse mapping systems see, for instance, S. Lefschetz, Algebraic Topology (Amer. Math. Soc, Providence, R.I., 1942), pp. 31-3 and other pertinent passages. In connection with theorem 5.1.2, the author is indebted for enlightening advice to Dr Lucien LeCam and also to Dr L. Breiman. 5.4. The characteristic functional was systematically introduced in our 'Stochastic Processes' Ann. Math. vol. 48 (1947), pp. 1014-61, for random additive set functions, and it was also introduced for point functions in L. LeCam, 'Un instrument d'etude des fonctions aleatoires. la fonctionelle caracteristique', CM. Acad. Sci., Paris, vol. 224 (1947), pp. 710-11. For applications see D. G. Kendall, 'Stochastic processes and population growth', J.B. Statist. Soc. B, vol. 11 (1949), pp. 230-65; M. S. Bartlett and D. G. Kendall, ' On the use of the characteristic functional in the analysis of some stochastic process occurring in physics and biology', Proc. Comb. Phil. Soc. vol. 47 (1951), pp. 65-80; and also M. S. Bartlett, 'The dual recurrence relation for multiplicative processes', Proc. Camb. Phil. Soc. vol. 47 (1951), pp. 821-5. 5.5. Dealing virtually exclusively with random functions on the ordinary straight line, J. L. Doob, both in his papers and in his treatise, Stochastic Processes (Wiley, New York, 1945), prefers to speak of a 'process of increments' rather than of random interval functions. 5.6. See G. E. Bates and Jerzy Neyman, 'Contributions to the theory of accident proneness. II. True or false contagion', Univ. California Publ. Statist, vol. 1 (1952), pp. 255-76, in particular formula (56) and neighboring ones. This paper was enlightening to us for the emphasis it places on the fact that in population and fission problems the random entity is not a point function but a finitely additive interval function only, and that care must be taken in distinguishing which end point of the interval does belong to it and which does not. Chapter 6 6.1. For random-point functions on arbitrary point set, theorems 6.1.3 and 6.1.2 were presented in the memoir of M. Loeve, 'Fonctions aleatoires du second order', which was published on pp. 299-352 as a 'Note' to the book by P. Levy, Processus Stochastiques et Mouvement Brownien (Gauthier-Villars, Paris, 1949). Theorem 6.1.2, again for point functions only and with further restriction of 'separability', was also implicitly proven by Hilbert space theory on pp. 368-71 of N. Aronszajn, 'Theory of reproducing kernels', Trans. Amer. Math. Soc. vol. 68 (1950), pp. 337-404. 6.2. What we now call L-finiteness was in our 'Stochastic Processes', Ann. Math, vol 48 (1947), pp. 1014-61, called ^-stability. 6.2-6.5. The spectral theory of a stationary process was begunin A. Kintchine, 'Korrelationstheorie der stationaren stochastischen Prozesse', Math. Ann. vol. 190 (1934), pp. 604-15, and a rounded version of Kintchine's theorem was given in H. Cramer, 'On harmonic analysis in certain functional spaces', Ark. Mat. Astr. Fys. vol. 28B, no. 12 (1942), 17 pp., but in the meantime H. Wold, in A Study in the Analysis of Stationary Time Series (Almquist and Wiksells, Uppsala, 1938), had taken the first systematic steps towards utilizing the spectral resolution for purposes of 'prediction theory'.
IJ2 NOTES AND REFERENCES After 1945 the spectral theory was reexamined in several studies, virtually simultaneously and independently of one another. Loeve introduced, among others, the (nonstationary) 'harmonizable' process. K. Karhunen in 'tJber lineare Methoden in der Wahrscheinlichkeitsrechnung', Ann. Acad. Sci. Fenn. A, vol. 37 (1947), 79 pp., very much exploited the Hilbert space approach. J. L. Doob in his article, 'Time series and harmonic analysis', Proceedings of the Berkeley Symposium in Mathematical Statistics and Probability (University of California Press, 1949, pp. 303-43), stressed the viewpoint of Norbert Wiener and finally in our 'Stochastic processes' Ann. Math. vol. 48 (1947), pp. 1014-61, the first attempt was made (still abortively there) towards dualizing the theory by envisaging interval functions on both sides of the inversion formula.
173 GENERAL INDEX Absolutely continuous additive set function, 13 Absolute value of additive set function, 12 Additive set function, 12; absolute value of, 12; absolutely continuous, 13; periodic, 18 Adjusted Zeta kernel [def. 2.8.1] 44 Admissible family (of subsets) D, ] 37 Bernoulli convergence, 15 Bernoulli addend ^*(a), 67 Characteristic function (Fourier transform of a distribution function), 26; of a set <i)ai>{x)9 8; of a set u)&(t), 139 Characteristic functional [def. 5.4.3], 137 Closed octant T, 89 Completely monotone function [theorem 4.1.1], 83; in several variables, 89 Completely monotone mapping [def. 4.1.1], 83; in several variables [def. 4.2.2], 89 Cone [def. 4.2.1], 88; proper, 88 Consistent mapping, 77 Continuous Markoff: (chain) density, 97 Convergence: Bernoulli, 15; P-, 55; in probability (in measure), 78, 150; weak, 16 Convergence factor [def. 2.1.1], 25, 23; [def. 2.8.2], 44 Convolution, 15 Covariance kernel, 146 Density, see Markoff (chain) density Diffusion equation [formula (2.5.14)], 36; [formula (4.6.11)], 103 Directed set [def. 5.1.1], 118 Distribution function of ^-functions (or k random variables), 77 Euclidean stochastic family, 136 Euclidean stochastic process, 136 Gaussian, Poisson, symmetric, 145 Expected value, 78 Fejer kernel: non-periodic, 5; periodic, 20 Finite (Lebesgue) measure space, 76 Forward projective limit, 134 Fourier transforms ^(c^) and 0/(cfy), 22 Fully homogeneous Markoff (chain) density, 98 Gaussian addend i/rG(oc), 67 Gaussian kernel, 4 Gaussian law, 93 Gauss-Poisson process, 145 Generated field, 77 Generated measure, 76 Generating functional, 143 Harmonic, see Spherical harmonic Heat equation [formula (2.5.11)], 35 Hilbert kernel, 146 Homogeneity of distance function, 128 Homogeneous Markoff (chain) density see Fully homogeneous Markoff (chain) density Homogeneous Markoff process, 124 Homogeneous process X(A) [def. 5.5.4], 140-1 Independence, stochastic, 79 Infinitely subdivisible process (or law) [def. 3.4.1], 69; subordinate [theorem 4.3.1], 92 Joint distribution function, see Distribution function .EC-stationary random (additive set) function, 160
174 GENERAL INDEX Kernel, 1; covariance. 146; Fejer, non-periodic, 5; Fejer, periodic, 20; Gaussian, 4; Hilbert, 146; positive- definite, 146; product, 4; special [formula (1.1.5)], 2-3 Kolmogoroff property, 119 £2-atomistic measure [def. 6.3.2], 155 ^(Aj-norm, 3 L„(Mk) -space, 32-3 X^^J-norm, 3 X2?(n)-norm, 153 Laplace equation [formula (2.5.12)], 36 Lebesgue measure space, 76; finite, 76; cr-flnite, 76; topological, 76; strictly topological, 76 Mapping: consistent, 77; completely monotone [def. 4.1.1], 83; completely monotone in several variables [def. 4.2.2], 89 Markoff (chain) density, 97; continuous, 97; homogeneous, etc., 97-8; of special kind [def. 4.5.2], 100; subordinate [theorem 4.4.3], 98 Markoff process [def. 5.2.7], 124; homogeneous, etc., 124 Matrix space, 116-17 Maximal family: simply, 118; sequentially, 119 Measure space, see Lebesgue measure space Modular relation [formula (4.8.15)], 172 Norm: Lp(A)9Z;L9(JEk)9 3; L„(Tk), 3; LV(Q), 153 Octant, 14; closed, T, 89; open, X, 89 Open octant X, 89 Orthogonal random (additive set) function, 161 Orthonormal random (additive set) function, 161 P-closed, 56 P-closure, 56 P-convergent, 55 P-finite, 61 P-limit, 55 Partially ordered vector space, 80 Periodic additive set function, 18 Poisson addend i/r^ia), 67 Poisson character [def. 3.1.2], 53 Poisson summation formula [formulae (2.4.2) and (2.4.7)], 31-2 Poisson transform [def. 3.3.1], 59 Positive-definite functions [theorem 3.2.3], 58 Positive-definite kernel, 146 Probability space, 76 Product kernel, 4 Projection, 118 Projective limit, 118; forward, 134 Proper cone, 88 Proper subordinate, 95 Pseudo-character x(a; x) [def. 3.LI], 52 Pseudo-Gaussian function [def. 3.3.3], 61 Pseudo-transform [def. 3.2.2], 56 Radial function, 5, 37 Random function (random additive set function), 137; iC-stationary, 160; orthogonal, 161; orthonormal, 161 Random variable, 77 Sequentially maximal family, 119 Set function, see Additive set function Simply maximal family, 118 Space homogeneous Markoff (chain) density, 98 Special kernel [formula (1.1.5)], 2-3 Spherical average of a function [formula (1.1.27)], 6 Spherical harmonic [def. 2.6.1], 37 Stable laws [formula (4.3.11)], 93 Stationary Markoff (chain) density, see Time homogeneous Markoff (chain) density Step function [def. 5.5.2], 139
GENERAL INDEX 175 Stieltjes integral, 140, 142 Stochastic family, 119; topological, 119; Euclidean, 136 Stochastic independence, 79 Stochastic process [def. 5.1.4], 119; Euclidean [theorem 5.4.4], 136 Strictly topological (Lebesgue) measure space, 76 Subdivisible process (or law), see Infinitely subdivisible process (or law)' Subordinate Markoff (chain) density [theorem 4.4.3], 98 Subordinate (infinitely) subdivisible process [theorem 4.3.1], 143 Subordinate, 95; proper, 95 Summation formulae involving Bessel functions, 112 Time homogeneous Markoff (chain) density, 97 Topological (Lebesgue) measure space, 76 Topological stochastic family, 119 Translation function: Tf(u)9 7; r^w), 17 Weak convergence, 16 Wiener process [def. 6.2.3], 153 Zeta kernel [def. 2.9.1.], 47; adjusted [def. 2.8.1], 74
176 INDEX OF SYMBOLS AO(Ek)9 13 <7(r>, C°°, 6; CM, 83; CM{X), CM(Y)9 89 D, 137 Ektl;B{...},78 F(A), 12; P(«; xs)f F(t; x)9 F(t; A), F(t; •), 69; P(w; A), 93 i^,...,^), 1 L0> 139; £p, 151; L'p9 158; 2^), 3; L,(Tk), 3; £CT (5), 154; L1>2t 102 ^,32 P, P(£), P(fl), 76; P(a; *), 53 $(»), 53; Q(a; #), 53; §(a; a), 65 i2, 76; 5(a), 61-2 ^V 3 F, V+9 13; F(^), F+(.Tfc), F+(A<7), 13 X(A), X(A), 139 ^,76 #, 76 «^, 7; JJ, 8 &9 76 A:(A), 118; A', 138 ry, 7; 7* 17 $(&)9 136-7; 0(h), 140 X(oc; x)9 52; ^(a), ^ff(a), ^(«), 67; ^(a), ^(a), 93; ^B(a)-i ^(a), 55 ^oo, woo, 118; .3, <D, 137 (w, w), (wA, <yA), 135