On the remainder in some linear approximation formulas of the analysis

Abstract

Authors

T. Popoviciu
Institutul de Calcul

Keywords

?

Paper coordinates

T. Popoviciu, Asupra restului în unele formule liniare de aproximare ale analizei, Studii Cerc. Mat. (Cluj), 10 (1959) no. 2, pp. 337-389 (in Romanian).

PDF

About this paper

Journal

Studii si Cercetari Matematice

Publisher Name

Academy of the Republic of S.R.

Print ISSN

1220-269X

Online ISSN

google scholar link

??

Paper (preprint) in HTML form

ABOUT THE REST IN SOME LINEAR APPROXIMATION FORMULAS OF ANALYSIS

by
TIBERIU POPOVICIU
(Cluj)

T

Many of the approximation formulas of analysis are of the form

A[f]B[f], or A[f]=B[f]+R[f],A[f]\approx B[f],\text{ sau }A[f]=B[f]+R[f], (*)

whereA[f],B[f]A[f],B[f]are linear functionals defined on a vector set of real and continuous functions of a real variable and whose remainderR[f]=A[f]B[f]R[f]=A[f]--B[f]is canceled onn+1n+1given functionsfi,i=0,1,,nf_{i},i=0,1,\ldots,nThe restR[f]R[f]is also a linear functional and vanishes on any linear combination of the functionsfif_{i}.

We will only consider real functionals and by a linear functional we will understand an additive and homogeneous functional.

The usual formulas for interpolation (polynomial or trigonometric), differentiation, and numerical integration, etc., are of the preceding form.

In applications it is important to be able to conveniently delimit the remainder. For this, at least in certain well-determined particular cases, attempts have been made to put the remainder in various convenient forms. For example, it has been obtainedR[f]R[f]in the form of a given linear combination of one or more values ​​of one or more derivatives of certain orders of the functionff. The remainder was also expressed in the form of a definite integral. It is sufficient to quote Taylor's formula which gives an approximation of the value of the functionfffor a given value ofxxand whose remainder is given by the well-known Lagrange formula or by a well-known integral representation [4].

Much research has been done on the rest. We will content ourselves with citing the works of AA Markov [6]. GD Birkhoff [1], G. Kowalewski [5], R. v. Mises [7], J. Radon [21], E. Ya. Remez [22], A. Sard [23].

00footnotetext: *) This work is also published in French in the journal "Mathematica" vol. 1 (24), fascicula 1.

In this paper we will highlight another expression of the remainder, which is more general, in the sense that, in general, it does not require the existence of derivatives, other than those that actually intervene in the formula()(*).

The new shape we give to the remainder makes its structure stand out better. We obtained this result with the help of the theory of higher-order convex functions, which we studied elsewhere [12, 13].

Under a certain particular hypothesis made on the functionsfif_{i}, a case that nevertheless encompasses a vast field of applications, the expression we find for the remainder is closely related to certain average formulas.

We will make some considerations on these average formulas and thus we will find part of DV Widder's results [28].

In this case it is easy to deduce the remainder expressed as a linear combination of derivatives, if, of course, these derivatives exist.

We obtained some of these results [16] in the particular case when the functionsfif_{i}are reduced to successive powersxi,i=0,1,,nx^{i},i=0,1,\ldots,nhis/hersxx, so in the case where the remainder cancels out for any polynomial of degreennIn this case, we also gave applications to certain formulas for derivation [17] and numerical integration [19].

This paper is divided into 4 parts. In § 1 we study the new expression of the remainder in the case when it has the form that we agree to call simple. In § 2 we study the average formulas that we have indicated. In § 3 we give examples for certain criteria that allow us to decide whether the remainder is of simple form or not. Finally, in § 4 we say a few words about the case when the remainder is not of simple form and we conclude this paragraph with an application. The results§x3,4\S\xi 3.4shows us, on the one hand, their connection with other known results, in particular with the results of E. Ya Remez [22] and, on the other hand, the degree of generality of the expression obtained for the remainder.

§ 1.

  1. 1.

    All functions considered in this paper will be assumed to be real and of a real variable. We will denote byANDANDthe definition set of the function or the definition set of functions considered simultaneously. We will always specify, if necessary, the structure ofANDAND.

We denote with

In(g1,g2,,gmx1,x2,,xm)=|gj(xi)|i,j=1,2,,mV\left(\begin{array}[]{ll}g_{1},&g_{2},\ldots,g_{m}\\ x_{1},&x_{2},\ldots,x_{m}\end{array}\right)=\left|g_{j}\left(x_{i}\right)\right|_{i,j=1,2,\ldots,m}

determinant of function values

g1,g2,,gmg_{1},g_{2},\ldots,g_{m} (2)

on the pointsxiAND,i=1,2,,mx_{i}\in E,i=1,2,\ldots,mIn the determinant (1),gj(xi)g_{j}\left(x_{i}\right)is the element in line aii-a and column ajj-a.

The determinant is obviously zero if the pointsxix_{i}or if the functions (2) are not distinct.

We will keep the notation (1) only for the case when the pointsxix_{i}are distinct. Otherwise, we will suitably modify the definition of the determinant (1). This modification consists in replacing the lines corresponding to each group of pointsxix_{i}confused by lines formed by the values ​​of the functions (2) and their successive derivatives on these points. More precisely, eitherWith1,With2,,Withpz_{1},z_{2},\ldots,z_{p}the distinct points with which they coincide respectivelyk1,k2,,kp(k1,k2,,kp1,k1+k2++kp=m)k_{1},k_{2},\ldots,k_{p}\left(k_{1},k_{2},\ldots,k_{p}\geqq 1,k_{1}+k_{2}+\ldots+k_{p}=m\right)between the pointsxix_{i}. Then, for eachi=1,2,,pi=1,2,\ldots,p, there exists exactlykik_{i}lines formed by the values ​​of the functions (2) and their primeski1k_{i}-1derivatives on the pointWithiz_{i}This implies, of course, the existence of the derivatives considered. The numberkik_{i}is the multiplicity order of the pointWithiz_{i}.

Conveniently ordering the pointsxix_{i}, we can write the determinant (1) thus modified by

In(With1,With1,,With1k1,g1,g2,,gmk2With2,With2,,With2,,Withp,Withp,,Withpkp),V(\underbrace{z_{1},z_{1},\ldots,z_{1}}_{k_{1}},\underbrace{g_{1},g_{2},\ldots,g_{m}}_{k_{2}}z_{2},z_{2},\ldots,z_{2},\ldots,\underbrace{z_{p},z_{p},\ldots,z_{p}}_{k_{p}}), (3)

which is of the orderm=k1+k2++kpm=k_{1}+k_{2}+\ldots+k_{p}and in whichgs(r1)(Withi)g_{s}^{(r-1)}\left(z_{i}\right)is the element in a(k1+k2++ki1+r)\left(k_{1}+k_{2}+\ldots+k_{i-1}+r\right)-th line andss-th column,in=1,2,,ki,i=1,2,,p(k1+k2++ki1v=1,2,\ldots,k_{i},i=1,2,\ldots,p\left(k_{1}+k_{2}+\ldots+k_{i-1}\right.is replaced by 0 ifi=1i=1 ).

We highlight the following particular cases
11^{\circ}Ifgi=xi1,i=1,2,,mg_{i}=x^{i-1},i=1,2,\ldots,m, we will denote the determinant (1) byIn(x1,x2,,xm)V\left(x_{1},x_{2},\ldots,x_{m}\right)This is the Vandermonde determinant of the numbers 1.x1,x2,,xmx_{1},x_{2},\ldots,x_{m}and we have

In(x1,x2,,xm)=1,2,,mi<j(xjxi),(In(x1)=1)V\left(x_{1},x_{2},\ldots,x_{m}\right)\stackrel{{\scriptstyle 1,2,\ldots,m}}{{=}}\prod_{i<j}\left(x_{j}-x_{i}\right),\quad\left(V\left(x_{1}\right)=1\right) (4)

Also in this case the determinant (3) will be denoted by
In(With1,With1,,With1k1,With2,,With2k2,,Withp,Withp,,Withpkp)V(\underbrace{z_{1},z_{1},\ldots,z_{1}}_{k_{1}},\underbrace{z_{2},\ldots,z_{2}}_{k_{2}},\ldots,\underbrace{z_{p},z_{p},\ldots,z_{p}}_{k_{p}})where we assume that the pointsWithi,i=1,2,,pz_{i},i=1,2,\ldots,p, are distinct.
22^{\circ}In the case when all the pointsxix_{i}coincide withxx, we denote the determinant (1) modified withIN(g1,g2,,gm)W\left(g_{1},g_{2},\ldots,g_{m}\right). This is the Wronskian of the functions (2). We therefore have

In(x,x,,xm)=IN(1,x,x2,,xm1)=(m1)!!V(\underbrace{x,x,\ldots,x}_{m})=W\left(1,x,x^{2},\ldots,x^{m-1}\right)=(m-1)!!

where did I puta!!=1!2!a!(0!!=1)\alpha!!=1!2!\ldots\alpha!(0!!=1).
2. We can also obtain the determinant (3) by passing the first limit if all the derivatives involved are continuous onANDE, or at least in the vicinity of the pointsWithiz_{i}.

Whethermmdistinct pointsxj(i),j=1,2,,ki,i=1,2,,px_{j}^{(i)},j=1,2,\ldots,k_{i},i=1,2,\ldots,pand form the determinantDDof the ordermmwhose element of a(k1+k2+++ki1+c)a\left(k_{1}+k_{2}+\right.\left.+\ldots+k_{i-1}+\gamma\right)-aline andsas-acolumn is the difference divided (ordinary) by the orderr1,[x1(i),x2(i),,xr(i);gs],r=1,2,,ki,s==1,2,,mr-1,\left[x_{1}^{(i)},x_{2}^{(i)},\ldots,x_{r}^{(i)};g_{s}\right],r=1,2,\ldots,k_{i},s==1,2,\ldots,mIf we observe that this divided difference tends towards1(r1)!gs(r1)(Withi)\frac{1}{(r-1)!}g_{s}^{(r-1)}\left(z_{i}\right)when the pointsxj(i),j=1,2,,rx_{j}^{(i)},j=1,2,\ldots,rtend towardsWithiz_{i}, we see that the determinantDDtends to the determinant (3) divided byi=1n(ki1)\prod_{i=1}^{n}\left(k_{i}-1\right)  ! !, whenxj(i)Withi,j=1,2,,ki,i=1,2,,px_{j}^{(i)}\rightarrow z_{i},j=1,2,\ldots,k_{i},i=1,2,\ldots,p. Finally, if we multiply the determinantDDthrough the product

i=1pIn(x1(i),x2(i),,xki(i))\prod_{i=1}^{p}V\left(x_{1}^{(i)},x_{2}^{(i)},\ldots,x_{k_{i}}^{(i)}\right) (5)

and once we do a few elementary operations on the lines, we obtain the determinant

In(g1,g2,,gnx1(1),x2(1),,xk1(1),x1(2),x2(2),,xk2(2),,x1(p),x2(p),,xkp(p)).V\binom{g_{1},g_{2},\ldots\ldots,g_{n}}{x_{1}^{(1)},x_{2}^{(1)},\ldots,x_{k_{1}}^{(1)},x_{1}^{(2)},x_{2}^{(2)},\ldots,x_{k_{2}}^{(2)},\ldots,x_{1}^{(p)},x_{2}^{(p)},\ldots,x_{k_{p}}^{(p)}}. (6)

It follows that the determinant (3) is obtained by multiplying (6) by(ki1)\left(k_{i}-1\right)  ! !, dividing 1 by (5) and then making the pointsxj(i)x_{j}^{(i)}to tend towardsWithiz_{i}forj=1,2,,ki,i=1,2,,pj=1,2,\ldots,k_{i},\quad i=1,2,\ldots,p.

The procedure of passing to the limit by which determinant (3) was obtained starting from determinant (1) can be generalized. Namely, a determinant of the form (3) can be obtained in the same way starting from determinants of the same form. We do not insist on this generalization because it will not be used in what follows.

As a first application we find the formula

In(With1,With1,,With1k1,With2,With2,,With2k2,,Withp,Withp,,Withpkp)=\displaystyle V(\underbrace{z_{1},z_{1},\ldots,z_{1}}_{k_{1}},\underbrace{z_{2},z_{2},\ldots,z_{2}}_{k_{2}},\ldots,\underbrace{z_{p},z_{p},\ldots,z_{p}}_{k_{p}})=
=[i=1p(ki1)!!]1,2,,p(With¯jWithi)kikj\displaystyle=\left[\prod_{i=1}^{p}\left(k_{i}-1\right)!!\right]^{1,2,\ldots,p}\left(\bar{z}_{j}-z_{i}\right)^{k_{i}k_{j}} (7)

We have, based on the well-known formula (see, e.g., IV Gonciarov [3]),

In(1,cosx,sinx,cos2x,sin2x,,cosmx,sinmxx1,x2,,x2m+1)=\displaystyle V\binom{1,\cos x,\sin x,\cos 2x,\sin 2x,\ldots,\cos mx,\sin mx}{x_{1},x_{2},\ldots\ldots,x_{2m+1}}=
=2mi<j1,2,2m+1(2sinxjxi2)\displaystyle=2^{-m\prod_{i<j}^{1,2,\ldots 2m+1}}\left(2\sin\frac{x_{j}-x_{i}}{2}\right) (8)

from which it results and

In(1,cosx,sinx,cos2x,sin2x,,cosmx,sinmxWith1,With1,,With1,With2,With2,,With2k1,,Withp,Withp,,Withpk2)=\displaystyle V\binom{1,\cos x,\sin x,\cos 2x,\sin 2x,\ldots,\cos mx,\sin mx}{z_{1},z_{1},\ldots,z_{1},\underbrace{z_{2},z_{2},\ldots,z_{2}}_{k_{1}},\ldots,\underbrace{z_{p},z_{p},\ldots,z_{p}}_{k_{2}}}= (9)
=2m[i=1n(ki1)!!]i<j1,2,,p(2sinWithjWithi2)kikj,(k1+k2++kp=2m+1)\displaystyle=2^{-m}\left[\prod_{i=1}^{n}\left(k_{i}-1\right)!!\right]_{i<j}^{1,2,\ldots,p}\left(2\sin\frac{z_{j}-z_{i}}{2}\right)^{k_{i}k_{j}},\left(k_{1}+k_{2}+\ldots+k_{p}=2m+1\right)

If in the following we consider a determinant (1) with the pointsx˙i\dot{x}_{i}not all distinct, we will consider it modified in the manner explained above.
3. If among the pointsxix_{i}, on which the determinant (1) or the modified determinant (3) is defined, there exists one that has the order of multiplicitykkrespectively an order of multiplicityk\leqq k, we will say that this point is repeated bykk, or respectively, it is repeated at mostkktimes.

Definition 1. - We will say that the functions (2) form an interpolation system or a system (I) on the setANDE(having at least m points) if. we have

In(g1,g2,,x1,x2,,xm)0V\left(\begin{array}[]{lll}g_{1},&g_{2},&\ldots,\\ x_{1},&x_{2},&\ldots,\end{array}x_{m}\right)\neq 0

for any group ofmmdistinct pointsxiAND,i=1,2,,mx_{i}\in E,i=1,2,\ldots,m
The property of forming a system (I) on ANDE, for the functions (2), is more restrictive than their linear-independence property (onANDE). In other words, any system (I) is formed by linearly independent functions but not every system of linearly independent functions forms an (I) system.

It is also of interest to complete Definition 1 by
Definition 2. - We will say that the functions (2) form a regular system (I) of orderk(1km)k(1\leqq k\leqq m)onANDEif we have (10) for any group ofmmpuncturexiAND,i=1,2,,mx_{i}\in E,i=1,2,\ldots,m, each repeating at mostkktimes.

Ifk=mk=m, we will say that the functions (2) form a complete regular system (I) (onANDE ).

Regularity of the orderkkso it means that the determinant (3) is0\neq 0ifWithiAND,1kik,i=1,2,,p,k1+k2++kp=mz_{i}\in E,1\leqq k_{i}\leqq k,i=1,2,\ldots,p,k_{1}+k_{2}+\ldots+k_{p}=m, the pointsWithiz_{i}being distinct.

In definition 2 we always assume that ifk>1k>1, derivatives of the orderk1k-1of the functions (2) are continuous onANDEIn this way, the regularity of the orderk>1k>1implies continuity onANDEof derivatives of the orderk1k-1of the functions (2). The assumption made previously allows us to avoid any difficulty. This is obviously a restriction but, according to TJ Stieltjes [26], it ensures the validity of the passage to the limit from no. 2.

One can obviously define the modified determinant ( 3 ), assuming more general differentiability conditions, from which also a more general notion of regularity results, but then the limit properties are more complicated. We will systematically leave such generalizations aside.

The crowdANDEit can be any. In what follows the crowdANDEwill generally be an interval. Then the notion of derivative is the one known from elementary analysis.

It is clear that regularity of the orderkkimplies regularity of any lower order and that complete regularity implies regularity of any orderm\leqq mIn particular, the notions of system (I) and regular system (I) of order 1 are equivalent.

Finally, regularity of the orderkkis equivalent to one of the following properties:
11^{\circ}For any group ofmmpoints (counted by their multiplicity orders)WithiANDz_{i}\in E, respectively, by the orderskik_{i}of multiplicity,i=1,2,,pi=1,2,\ldots,p, k1+k2++kp=mk_{1}+k_{2}+\ldots+k_{p}=mand for any group ofmmNUMBERSandi(),j==0,1,,ki1,i=1,2,,py_{i}^{()},j==0,1,\ldots,k_{i}-1,i=1,2,\ldots,p, there is a linear combination of the functions (2) and a singlef(x)\varphi(x)for whichf(f)(Withi)=andi(f),j=0,1,,ki1\varphi^{(f)}\left(z_{i}\right)=y_{i}^{(f)},j=0,1,\ldots,k_{i}-1,i=1,2,,pi=1,2,\ldots,p.

Ifffis a function such thatandi(j)=f(j)(Withi),j=0,1,,ki1y_{i}^{(j)}=f^{(j)}\left(z_{i}\right),j=0,1,\ldots,k_{i}-1,i=1,2,,pi=1,2,\ldots,p, we will denote this linear combination and with

L(fx)=L(With1,With1,,With1k1,With2,k2,With2,,With2,,Withkp,Withp,,Withp;fx).L(f\mid x)=L(\underbrace{z_{1},z_{1},\ldots,z_{1}}_{k_{1}},z_{2,}\underbrace{}_{k_{2}},z_{2},\ldots,z_{2},\ldots,z_{k_{p}},z_{p},\ldots,z_{p};f\mid x). (11)

If the functions (2) form a regular system (I) of orderkkand ifkik,i=1,2,,pk_{i}\leqq k,i=1,2,\ldots,p, the linear combination (11) is well-determined (and unique).

It is clear that ifffreduces to a linear combination of the functions (2) and if the previous condition is verified, we haveL(fx)=fL(f\mid x)=f.
22^{\circ}A linear combination of the functions (2) cannot be cancelled onmmpoints, each of which is repeated at mostkktimes, without being identically null.

It is said that a function is canceled bykktimes on a point, if this function and its primesk1k-1derivatives are zero at this point.

Formula (7) shows us that the functionsxi,i=0,1,,m1x^{i},i=0,1,\ldots,m-1, forms a completely regular (I) system for any natural numbermmand on a certain crowdANDE.

Also formula (9) shows us that the functions1,cosix,sinix,i==1,2,,m1,\cos ix,\sin ix,i==1,2,\ldots,m, forms a completely regular system (I), for any natural numbermmand on any interval that does not contain a closed subinterval of length2p2\pi, so, in particular on the interval[0,2p[0,2\pi), closed on the left and open on the right.
4. If the functions (2) are continuous, we can find more complete results. Thus, we have
Theorem 1. - If the functions (2) :11^{\circ}. are continuous on the intervalANDE,22^{\circ}. forms a regular system (I) of orderkkonANDE,
the determinant (1) does not change sign, as long as the pointsxix_{i}, each of which is repeated at mostkktimes, do not change their relative order of magnitude (e.g., as long as the functionsgig_{i}remain in the order indicated by the string (2) and the pointsxix_{i}remain in the ascending order of their indices,x1x2xmx_{1}\leqq x_{2}\leqq\ldots\leqq x_{m} ).

According to the previous definitions, fork>1k>1(but not fork=k=1) condition22^{\circ}of the statement implies the continuity of functions (2).

Let us first assume thatk=1k=1. For the sake of demonstration, let us assume the opposite. We can then find the points

x1<x2<<xm,x1′′<x2′′<<xm′′x_{1}^{\prime}<x_{2}^{\prime}<\ldots<x_{m}^{\prime},x_{1}^{\prime\prime}<x_{2}^{\prime\prime}<\ldots<x_{m}^{\prime\prime} (12)

so that we have

In(g1,g2,,gmx1,x2,,xm)>0,In(g1,g2,,gmx1′′,x2′′,,xm′′)<0.V\binom{g_{1},g_{2},\ldots,g_{m}}{x_{1}^{\prime},x_{2}^{\prime},\ldots,x_{m}^{\prime}}>0,\quad V\binom{g_{1},g_{2},\ldots,g_{m}}{x_{1}^{\prime\prime},x_{2}^{\prime\prime},\ldots,x_{m}^{\prime\prime}}<0. (13)

points

xi=lxi′′+(1l)xi,i=1,2,,mx_{i}=\lambda x_{i}^{\prime\prime}+(1-\lambda)x_{i}^{\prime},\quad i=1,2,\ldots,m (14)

remain distinct for0l10\leqq\lambda\leqq 1and the determinant (1) is a function ofl\lambdaperfectly determined and continues on[0,1][0,1].

A well-known property of continuous functions shows us that there is al,0<l<1\lambda,0<\lambda<1so that the determinant (1), where the pointsxix_{i}are given by (14), to be equal to zero. This is in contradiction with the hypothesis that the functions (2) form a system (I).

Let us also note that from (12) it follows that, for0l10\leqq\lambda\leqq 1, points (14) also verify the inequalitiesx1<x2<<xmx_{1}<x_{2}<\ldots<x_{m}staying within a length rangemax(xmx1,xm′′x1′′)\leqq\max\left(x_{m}^{\prime}-x_{1}^{\prime},x_{m}^{\prime\prime}-x_{1}^{\prime\prime}\right)If in addition0<l<10<\lambda<1andx1x1′′,xmxm′′x_{1}^{\prime}\neq x_{1}^{\prime\prime},x_{m}^{\prime}\neq x_{m}^{\prime\prime}, the pointsxix_{i}are inside the smallest interval containing the pointsxi,xi′′x_{i}^{\prime},x_{i}^{\prime\prime}.

Let's assume nowk>1k>1. For the sake of demonstration, let us assume the opposite again. We can then find the pointsin1in2inmu_{1}\leqq u_{2}\leqq\ldots\leqq u_{m}, each of which is repeated at mostkktimes and pointsin1in2inmv_{1}\leqq v_{2}\leqq\ldots\leqq v_{m}, each of which is also repeated at mostkktimes, so that

In(g˙1,g2,,gmin1,in2,,inm)>0,In(g1,g2,,gmin1,in2,,inm)<0.V\binom{\dot{g}_{1},g_{2},\ldots,g_{m}}{u_{1},u_{2},\ldots,u_{m}}>0,\quad V\binom{g_{1},g_{2},\ldots,g_{m}}{v_{1},v_{2},\ldots,v_{m}}<0. (15)

We can then find the variable points (12), tending respectively to the pointsiniu_{i}andiniv_{i}, so that the product of the determinants (13), through some functions that remain positive, tends to the respective determinants (15). It follows that this time too it is possible to find the points (12) such that we have (13). The proof therefore returns to that of the previous case.

Theorem 1 is therefore completely proven.
5. The linear combination (11) can be written

L(fx)=f(x)In(g1,g2,,gm,fx1,x2,,xm,x)In(g1,g2,,gmx1,x2,,xm)L(f\mid x)=f(x)-\frac{V\binom{g_{1},g_{2},\ldots,g_{m},f}{x_{1},x_{2},\ldots,x_{m},x}}{V\binom{g_{1},g_{2},\ldots,g_{m}}{x_{1},x_{2},\ldots,x_{m}}} (16)

wherexix_{i}are the pointsWithiz_{i}, with their multiplicity orders, in some order. Formula (16) has a precise meaning ifxxdoes not coincide with one of the pointsxix_{i}. Otherwise we agree to replace the second term of the second member by zero. This convention is necessary to avoid confusion with the definition of the determinant (1) in the case when the pointsxix_{i}they are not all distinct.

From formula (16) it follows

L(fx)L(gx)=f(x)g(x)In(g1,g1,,gm,fgx1,x2,,xm,x)In(g1,g2,,gmx1,x2,,xm)L(f\mid x)-L(g\mid x)=f(x)-g(x)-\frac{V\binom{g_{1},g_{1},\ldots,g_{m},f-g}{x_{1},x_{2},\ldots,x_{m},x}}{V\binom{g_{1},g_{2},\ldots,g_{m}}{x_{1},x_{2},\ldots,x_{m}}}

In particular, if the pointsxix_{i}are distinct and the functions (2) are continuous, we deduce the inequality

|L(fx)L(gx)|Mmaxi=1,2,,m(|f(xi)g(xi)|)|L(f\mid x)-L(g\mid x)|\leqq M\max_{i=1,2,\ldots,m}\left(\left|f\left(x_{i}\right)-g\left(x_{i}\right)\right|\right) (17)

whereM(>0)M(>0)is the maximum, in the smallest closed interval containing the pointsxix_{i}, of the continuous function

i=1m|In(g1,g2,,gmx1,x2,,xi1,xi+1,xi+1,,xm,x)||In(g1,g2,,gmx1,x2,,xm)|\frac{\sum_{i=1}^{m}\left|V\binom{g_{1},g_{2},\ldots,g_{m}}{x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+1},\ldots,x_{m},x}\right|}{\left|V\binom{g_{1},g_{2},\ldots,g_{m}}{x_{1},x_{2},\ldots,x_{m}}\right|}

where the determinants that intervene (in the numerator) are defined by the second member of formula (1).

We can easily generalize this result to the case where the pointsxix_{i}they are not distinct.

We then deduce
theorem 2. - If:11^{\circ}. the functions (2) are continuous and form a system (I) on the intervalAND,2E,2^{\circ}. the linear combinationf\varphiof these functions is canceled onm1m-1distinct pointsxi,i=1,2,,m1x_{i},i=1,2,\ldots,m-1, without being identically null onANDE,
functionf\varphi(is continuous and) changes sign passing through a pointxix_{i}(which does not coincide with an extremity ofANDE ).

It is assumed, of course,m>1m>1.
This property is well known. For completeness, we will give its proof.

Let us suppose that, contrary to the statement,f\varphidoes not change sign passing through the pointx1x_{1}, which does not coincide with one of its extremitiesANDEWe can then find the pointsx1,x1′′x_{1}^{\prime},x_{1}^{\prime\prime}thusthat:1.x1<x1<x1′′,2\mathrm{ca}:1^{\circ}.x_{1}^{\prime}<x_{1}<x_{1}^{\prime\prime},2^{\circ}, none of the pointsxi,i=2,3,,m1ninx_{i},i=2,3,\ldots,m-1\mathrm{nu}belongs to the closed interval[x1,x1′′]\left[x_{1}^{\prime},x_{1}^{\prime\prime}\right], 3.f(x1)f(x1′′)>03^{\circ}.\varphi\left(x_{1}^{\prime}\right)\varphi\left(x_{1}^{\prime\prime}\right)>0Let us consider inequality (17) relative to the pointsx1,x1,x2,,xm1x_{1}^{\prime},x_{1},x_{2},\ldots,x_{m-1}, to the functionf\varphiand to the linear combinationf1\varphi_{1}of functions
(2) which takes the same values ​​asf\varphion the pointsx1,x2,x3,,xn1x_{1}^{\prime},x_{2},x_{3},\ldots,x_{n-1}and for whichf1(x1)=esgf(x1′′)\varphi_{1}\left(x_{1}\right)=-\varepsilon\operatorname{sg}\varphi\left(x_{1}^{\prime\prime}\right), uncle e\varepsilonis a positive number<12M|f(x1′′)|<\frac{1}{2M}\left|\varphi\left(x_{1}^{\prime\prime}\right)\right|We then have

|f1(x1′′)f(x1′′)|=|L(f1x1′′)L(fx1′′)|Me<|f(x1′′)|2\left|\varphi_{1}\left(x_{1}^{\prime\prime}\right)-\varphi\left(x_{1}^{\prime\prime}\right)\right|=\left|L\left(\varphi_{1}\mid x_{1}^{\prime\prime}\right)-L\left(\varphi\mid x_{1}^{\prime\prime}\right)\right|\leqq M\varepsilon<\frac{\left|\varphi\left(x_{1}^{\prime\prime}\right)\right|}{2}

It follows that sgf1(x1′′)=sgf(x1′′)\varphi_{1}\left(x_{1}^{\prime\prime}\right)=\operatorname{sg}\varphi\left(x_{1}^{\prime\prime}\right)
It is now seen that f1\varphi_{1}, without being identical to null, cancels out at the pointsx2,x3,,xm1x_{2},x_{3},\ldots,x_{m-1}and at least once more on each of the open intervals(x1,x1),(x1,x1′′)\left(x_{1}^{\prime},x_{1}\right),\left(x_{1},x_{1}^{\prime\prime}\right)This isin^\hat{in}contradiction with the fact that functions (2) form a system (I).

Theorem 2 is proved.
6. Suppose that then+2n+2functions

f0,f1,,fn,fn+1f_{0},f_{1},\ldots,f_{n},f_{n+1} (18)

are defined and form a system (I) onANDEIt is easy to see that then the firstn+1n+1of these functions

f0,f1,,fnf_{0},f_{1},\ldots,f_{n} (19)

are linearly independent onANDEWe say [ 13
] that the functionffis convex, respectively concave with respect to the sequence (19) of functions, if

In(f0,f1,,fn,fx1,x2,,xn+2)>0 resp. <0V\binom{f_{0},f_{1},\ldots,f_{n},f}{x_{1},x_{2},\ldots,x_{n+2}}>0\text{ resp. }<0 (20)

for any systemx1<x2<<xn+2x_{1}<x_{2}<\ldots<x_{n+2}ofn+2n+2his pointsANDEIf
the functionffis convex or concave with respect to the sequence (19), the terms of the sequence together with this function form a system (I) (onANDE). Conversely, if the functions (18) are continuous and form a system (I), the functionfn+1f_{n+1}and, in general, any one of these functions is convex or concave with respect to any series formed with the othersn+1n+1functions.

In what follows we will assume that the integernnis0\geqq 0.
The previous definition can also be given meaning in the case ofn=1n=-1Then the series (19) disappears and the convexity, respectively the concavity of the functionffreturns to its positivity and negativity respectivelyANDE.

The notion of convexity thus introduced generalizes that of higher-order convexity (of the ordernn) [12], which is obtained in the particular case

fi=xi,i=0,1,,nf_{i}=x^{i},\quad i=0,1,\ldots,n (21)

In this case the function

fn+1=xn+1f_{n+1}=x^{n+1} (\prime)

is convex with respect to the sequence of functions (21), the intervalANDEbeing any. 10 - Studies$1\mathdollar 1mathematical research
7. The definition inequalities (20) are not symmetric with respect to the pointsxix_{i}and the distinction between convexity and concavity depends on the order in which the functions occur (19). This is the reason why in the definition I emphasized that convexity and concavity are relative to the sequence and not to the set of functions (1).

We observe that ifffis convex or concave, the functionf-fis concave or convex respectively. The set of convex functions (or concave) with respect to the sequence (19) remains invariant or changes to the set of concave (or convex) functions by a permutation of the functions (19).

To remove these asymmetries we introduce the notation

[x1,x2,,xn+2;f]=In(f0,f1,,x1,x2,,xn+2):In(f0,f1,,fn,fn+1x1,x2,,xn+2),\left[x_{1},x_{2},\ldots,x_{n+2};f\right]=V\left(\begin{array}[]{lll}f_{0},&f_{1},&\ldots,\\ x_{1},&x_{2},&\ldots,\end{array}x_{n+2}\right):V\binom{f_{0},f_{1},\ldots,f_{n},f_{n+1}}{x_{1},x_{2},\ldots,x_{n+2}},

where we assume that the functions (18) form a system (I) and that the pointsxix_{i}are distinct. Then the expression (22) has a perfectly determined meaning and is symmetric with respect to the pointsxix_{i}In the particular case (21), (21') this expression reduces to the divided difference of the functionffon the nodesx1,x2,,xn+2x_{1},x_{2},\ldots,x_{n+2}We will continue to use the term divided difference for expression (22) and for the pointsxix_{i}the name of the nodes (of this divided difference or on which this divided difference is defined). In the notation (22) we omitted to highlight the functions (18) because we will never encounter two different systems (18) simultaneously in our considerations.

The divided differences thus defined enjoy some properties which are expressed by the formulas

[x1,x2,,xn+2;f]={0,i=0,1,,n1,i=n+1,\displaystyle{\left[x_{1},x_{2},\ldots,x_{n+2};f\right]=\left\{\begin{array}[]{l}0,\quad i=0,1,\ldots,n\\ 1,\quad i=n+1,\end{array}\right.} (23)
[x1,x2,,xn+2;af+bg]=a[x1,x2,,xn+2;f]+\displaystyle{\left[x_{1},x_{2},\ldots,x_{n+2};\alpha f+\beta g\right]=\alpha\left[x_{1},x_{2},\ldots,x_{n+2};f\right]+}
b[x1,x2,,xn+2;g]\displaystyle-\mid-\beta\left[x_{1},x_{2},\ldots,x_{n+2};g\right] (24)

whatever the functionsf,gf,g, the constantsa,b\alpha,\betaand the distinct pointsxiANDx_{i}\in E, i=1,2,,n+2i=1,2,\ldots,n+2. Formula (24) expresses the linearity property of the divided difference.
8. With the help of divided differences, the definition of convexity can be stated (in a somewhat more precise form) as follows:

Definition 3. - The function f is convex, non-concave, non-convex or concave with respect to the functions (19) if

[x1,x2,,xn+2;f]>0,0,0, resp. <0\left[x_{1},x_{2},\ldots,x_{n+2};f\right]>0,\geqq 0,\leqq 0,\text{ resp. }<0 (25)

pointsxiAND,i=1,2,,n+2x_{i}\in E,i=1,2,\ldots,n+2being distinct and arbitrary.
It is seen that the definition is independent of the order of the functions (19) and that the distinction between convex and concave functions is specified by the choice of the functionfn+1f_{n+1}which is, ipso facto, convex. We will see below
, when studying the restR[f]R[f], that the introduction of divided differences satisfies requirements that go far beyond our simple desire to restore the symmetry of certain formulas considered above.

Convexity (concavity) is a particular case of non-concavity (non-convexity). However, for what follows it is useful to make a clear distinction between non-concave (non-convex) functions in general and only convex (concave) functions.

Ifff- is convex or non-concave,ffis concave respectively non-convex and reciprocal.

The linear combination with all positive and all negative coefficients, respectively, of a finite number (at least 1) of non-concave functions is non-concave, respectively non-convex. If at least one of the functions considered is convex, the linear combination considered is convex or concave, respectively.

The limit of a convergent sequence (onANDE) of non-concave (non-convex) functions is a non-concave (non-convex) function.

A functionffcan be both non-concave and non-convex. The functions that verify this property are those and only those whose divided difference is zero on any group ofn+2n+2his pointsANDEFor this property to be verified it is necessary and sufficient that//to reduce to a linear combination of the functions (19). The condition is obviously sufficient. But it is also necessary. Indeed, since the functions (19) are linearly independent, there existsn+1n+1distinct pointsxi,i=1,2,,n+1x_{i},i=1,2,\ldots,n+1, so thatIn(f0,f1,,fnx1,x2,,xn+1)0[20]V\left(\begin{array}[]{ll}f_{0},&f_{1},\ldots,f_{n}\\ x_{1},&x_{2},\ldots,x_{n+1}\end{array}\right)\neq 0[20]We haveIn(f0,f1,,fn,fx1,x2,,xn+1,x)=0V\binom{f_{0},f_{1},\ldots,f_{n},f}{x_{1},x_{2},\ldots,x_{n+1},x}=0, forxANDx\in E, from which the property results.

Among the other properties of convex functions we point out
theorem 3. - If:11^{\circ}. the functions (18) are continuous and form a system (I) on the intervalAND,2E,2^{\circ}. function//is continuous but is neither convex nor concave onANDE,
can be foundn+2n+2distinct pointsxiAND,i=1,2,,n+2x_{i}\in E,i=1,2,\ldots,n+2, so that we have[x1,x2,,xn+2;f]=0\left[x_{1},x_{2},\ldots,x_{n+2};f\right]=0.

Indeed, if the function nut is neither convex nor concave, then it is either non-concave or non-convex and then the property is obvious or two groups of each can be foundn+2n+2distinct pointsxiANDx_{i}^{\prime}\in Eandxi′′ANDx_{i}^{\prime\prime}\in E, i=1,2,,n+2i=1,2,\ldots,n+2, so that the divided differences

[x1,x2,,xn+2;f],[x1′′,x2′′,,xn+2′′;f]\left[x_{1}^{\prime},x_{2}^{\prime},\ldots,x_{n+2}^{\prime};f\right],\left[x_{1}^{\prime\prime},x_{2}^{\prime\prime},\ldots,x_{n+2}^{\prime\prime};f\right] (26)

be different from zero and of opposite signs. It is then sufficient to apply Theorem 1, taking into account the definition formula (22) of divided differences.

We also deduce the more general property expressed by
Theorem 4. - If:11^{\circ}. the functions (18) are continuous and form a system (I) on the intervalAND,2E,2^{\circ}. functionffis continuous onAND,3E,3^{\circ}. CCis a number between the valuesA,BA,Bof the divided differences (26),
one can findn+2n+2distinct pointsxiAND;i=1,2,,n+2x_{i}\in E;i=1,2,\ldots,n+2, so that we have[x1,x2,,xn+2;f]=C\left[x_{1},x_{2},\ldots,x_{n+2};f\right]=C.

IfCCcoincides withAAor withBB, which necessarily occurs ifA=BA=B, the property is obvious. Otherwise we have(AB)(BC)<0(A-B)(B-C)<0. Taking into account (23), (24) it is easy to verify that the functionfCfn+1f-Cf_{n+1}is neither convex nor concave. It is then sufficient to apply Theorem 3 to this latter function.

From the observations made in the proof of Theorem 1, it follows that ifx1<x2<<xn+2,x1′′<x2′′<<xn+2′′x_{1}^{\prime}<x_{2}^{\prime}<\ldots<x_{n+2}^{\prime},x_{1}^{\prime\prime}<x_{2}^{\prime\prime}<\ldots<x_{n+2}^{\prime\prime}, the points can be chosenxix_{i}, so that we havex1<x2<<xn+2x_{1}<x_{2}<\ldots<x_{n+2}andxn+2x1max(xn+2x1,xn+2′′x1′′)x_{n+2}-x_{1}\leqq\max\left(x_{n+2}^{\prime}-x_{1}^{\prime},x_{n+2}^{\prime\prime}-x_{1}^{\prime\prime}\right), and if(AC)(BC)<0(A-C)(B-C)<0to have plus and minus(x1,x1′′)<xi<<max(xn+1,xn+2′′),i=1,2,,n+2\left(x_{1}^{\prime},x_{1}^{\prime\prime}\right)<x_{i}<<\max\left(x_{n+1}^{\prime},x_{n+2}^{\prime\prime}\right),\quad i=1,2,\ldots,n+2.
9. If the functions (18) form a regular system (I) of orderkk, we can take formula (22) to define any divided difference whose distinct nodes are repeated at mostkktimes. To highlight the multiplicity of nodes we will note this divided difference and with

[With1,With1,,With1h1,With2,With2,,With2h2,,Withp,Withp,,Withpkp;f][\underbrace{z_{1},z_{1},\ldots,z_{1}}_{h_{1}},\underbrace{z_{2},z_{2},\ldots,z_{2}}_{h_{2}},\ldots,\underbrace{z_{p},z_{p},\ldots,z_{p}}_{k_{p}};f] (27)

where the nodesWithiz_{i}of the respective multiplicity orderski,i=1,2,,pk_{i},i=1,2,\ldots,p, are distinct.

The results of 1a no. 2 show us that the divided difference (27) is the limit of the divided difference

[x1(1),x2(1),,xk1(1),x1(2),x2(2),,xk2(2),,x1(p),x2(p),,xkp(p);f]\left[x_{1}^{(1)},x_{2}^{(1)},\ldots,x_{k_{1}}^{(1)},x_{1}^{(2)},x_{2}^{(2)},\ldots,x_{k_{2}}^{(2)},\ldots,x_{1}^{(p)},x_{2}^{(p)},\ldots,x_{k_{p}}^{(p)};f\right]

on distinct nodes, ifxj(i)Withi,j=1,2,,ki,i=1,2,,px_{j}^{(i)}\rightarrow z_{i},j=1,2,\ldots,k_{i},i=1,2,\ldots,p
In particular, assuming that the functions (18) form a completely regular system (I), we have

[x,x,,x;f]=[IN(f0,f1,,fn,f)IN(f0,f1,,fn,fn+1)]x=x[\xi,\xi,\ldots,\xi;f]=\left[\frac{W\left(f_{0},f_{1},\ldots,f_{n},f\right)}{W\left(f_{0},f_{1},\ldots,f_{n},f_{n+1}\right)}\right]_{x=\xi} (28)

The various properties of divided differences defined on distinct nodes can be extended to divided differences on not all distinct nodes, defined in the way shown above. For example, formulas (23), (24) obviously remain valid.

We observe that if the functions (18) form a completely regular system (I) and if the functions (19) are solutions (necessarily linearly independent) of the linear and homogeneous differential equation of ordern1n\neq 1,

D[and]=and(n+1)+f1(x)and(n)++fn+1(x)and=0D[y]=y^{(n+1)}+\varphi_{1}(x)y^{(n)}+\ldots+\varphi_{n+1}(x)y=0

HAVEIN(f0,f1,,fn,f)=IN(f0,f1,,fn)D[f]W\left(f_{0},f_{1},\ldots,f_{n},f\right)=W\left(f_{0},f_{1},\ldots,f_{n}\right)D[f]and the formula (28) is clevin

[x,x,,x;f]=[D[f]D[fn+1]]x=xi[\xi,\xi,\ldots,\xi;f]=\left[\frac{D[f]}{D\left[f_{n+1}\right]}\right]_{\begin{subarray}{c}x=-\xi\\ i\end{subarray}} (29)

The divided difference (27) exists, based on the given definition, only if the determinantInVfrom the numerator of the second member of formula (22) exists in the sense of no. 2. In the following we will assume that the functionffhas all the intervening derivatives continuous. With this assumption the divided difference (27) exists under the above conditions.

More general divided differences can be defined on nodes not all distinct, by convenient limit crossings. These limit crossings can be done by means of the limits of ordinary divided differences (those corresponding to the particular case (21), (21')). In fact, this is how we proceed in this paper. One can also proceed directly, without going through the particular case (21), (21'). All these questions are closely related to the definition and existence of higher-order direct derivatives of a function.

To give an example, let us observe that in the particular case (21), (21'), the quotient (29) reduces to1(n+1)!f(n+1)(x)\frac{1}{(n+1)!}f^{(n+1)}(\xi)and this result is valid, by virtue of our convention, ifffhas a derivative of ordern+1n+1continues, at least on the pointx\xi. However, if for the first member of (29) (remaining in the particular case (21), (21')) we adopt as definition the limit of the divided difference[x1,x2,,xn+1,x;f]\left[x_{1},x_{2},\ldots,x_{n+1},\xi;f\right]when the pointsxi,i=1,2,,n+1x_{i},i=1,2,\ldots,n+1tend towardsx\xi, formula (29) remains valid, as TJ Stieltjes [26] showed, only under the hypothesis of the existence of the derivative of ordern+1n+1of the function at the pointx\xi(functionffis assumed to be defined and bounded onANDE ).

In what follows we will systematically leave aside such generalizations.
10. LetR[f]R[f]a linear functional, defined on a vector spacef\sqrt{f}made up of functionsffcontinue on the intervalANDE.

We assume that the functions (18) form a system (I) and belong to (f). In particular, they are continuous onANDE.

If the linear functionalR[f]R[f]vanishes on the functions (19), it is zero on any linear combination of these functions. Such a functional is, for example,

K.[x1,x2,,xn+2;f]K.\left[\xi_{1},\xi_{2},\ldots,\xi_{n+2};f\right] (30)

whereKKis a number independent of the functionffandxi\xi_{i}saintn+2n+2distinct points of the intervalANDE.

We will now introduce
Definition 4. - We will say that the linear functionalR[f]R[f], defined on𝕗\mathbb{f}, is of simple form if, for anyf𝕗f\in\mathbb{f}, it is of the form (30), whereKKis a non-zero number, independent of the functionff, andxi\xi_{i}saintn+2n+2distinctive points ofANDE(which may generally depend on the functionff).

We then have
THEOREM 5. - The necessary and sufficient condition for the linear functionalR[f]R[f]to be of simple form, is to haveR[f]0R[f]\neq 0for any functionf(ϵ)f(\epsilon)convex with respect to the functions (19).

The condition is necessary. Indeed, ifR[f]R[f]is of the simple form. From (23) it follows first thatR[fn+1]=K0R\left[f_{n+1}\right]=K\neq 0From the formula

R[f]=R[fn+1][x1,x2,,xn+2;f]R[f]=R\left[f_{n+1}\right]\left[\xi_{1},\xi_{2},\ldots,\xi_{n+2};f\right] (31)

it then follows thatR[f]0R[f]\neq 0ifffis convex.
The condition is sufficient. If we haveR[f]0R[f]\neq 0for any convex function, the same property is true for any concave function. Indeed, ifffis concave, the functionf-fis convex and we have

R[f]=R[f]0R[f]=-R[-f]\neq 0

Whetherf𝔽f\in\mathbb{F}and consider the auxiliary function

f=R[f]fn+1R[fn+1]f\varphi=R[f]\cdot f_{n+1}-R\left[f_{n+1}\right]\cdot f (32)

wef𝔽\varphi\in\mathbb{F}andR[f]=0R[\varphi]=0It follows thatf\varphiis neither convex nor concave. By virtue of Theorem 3 we can findn+2n+2distinct pointsxi\xi_{i}, i=1,2,,n+2i=1,2,\ldots,n+2, so that we have[x1,x2,,xn+2;f]=0\left[\xi_{1},\xi_{2},\ldots,\xi_{n+2};\varphi\right]=0Taking into account (23), (24), from (32) we deduce formula (31).

Theorem 5 is therefore proven.
IfR[f]R[f]is of simple form, it cancels out on the functions (19). This property can be deduced directly from the fact thatR[f]0R[f]\neq 0for any convex or concave function. To prove the property, let us assume the opposite, so thatR[fi]0R\left[f_{i}\right]\neq 0for ai,0ini,0\leqq i\leqq nIf we putf=fif=f_{i}in (32), we obtain a functionf\varphiwhich is convex or concave. The equalityR[f]=0R[\varphi]=0is then in contradiction with the hypothesis.

An analogous demonstration shows us that ifR[f]0R[f]\neq 0for any convex function, we have more preciselyR[fn+1]R[f]>0R\left[f_{n+1}\right]R[f]>0for these functions. In other wordsR[f]R[f]keeps its sign, which is the sign ofR[fn+1]R\left[f_{n+1}\right], for any convex function, so it keeps the opposite sign for any concave function.

It is also seen that ifR[f]R[f]is of simple form, we haveR[fn+1]R[f]0R\left[f_{n+1}\right]R[f]\geqq 0for any functionffnon-concave and we have the opposite inequality for any non-convex function.

§ 2.

  1. 11.

    RESTR[f]R[f], in the case when it is of simple form, is expressed, by the formula (31), as a divided difference. The structure of the remainder therefore depends on the structure of the divided difference (22). The structure of this divided difference is specified by an important mean theorem due to DV Wid der [28]. This theorem takes place under an additional hypothesis made on the functions (19), an hypothesis which we will indicate below.

We will find DV Widder's results in a different way. Our results, which are sufficient for the study of the rest, are a little more general, but they only allow us to find part of DV Widder's results in the particular case examined by this author.

The additional hypothesis we discussed above is that the functions (19) form a system (I). This is not a consequence of the fact that the functions (18) form a system (I) (see, e.g., the example given in no. 16). To avoid any difficulty, we will assume in the following that the functions (18) are continuous.
12. We will use the following formula

In(f0,f1,,fnx2,x3,,xn+2)In(f0,f1,,fn,ff0,f1,,fn,,xi1,xi+1,xi+2,,xn+3)=\displaystyle V\binom{f_{0},f_{1},\ldots,f_{n}}{x_{2},x_{3},\ldots,x_{n+2}}V\binom{f_{0},f_{1},\ldots,f_{n},f}{f_{0},f_{1},\ldots,f_{n},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3}}=
=\displaystyle= In(f1,,fn+3x2,x3,,xi1,xi+1,xi+2,,xn+3)In(f0,f1,,fn,fx1,x2,,xn+2)+\displaystyle V\binom{f_{1},\ldots,f_{n+3}}{x_{2},x_{3},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3}}V\binom{f_{0},f_{1},\ldots,f_{n},f}{x_{1},x_{2},\ldots,x_{n+2}}+
In\displaystyle--V (f0,f1,,fnx1,x2,,xi1,xi+1,xi+2,,xn+2)In(f0,f1,,fn,fx2,x3,,xn+3).\displaystyle\binom{f_{0},f_{1},\ldots,f_{n}}{x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+2}}V\binom{f_{0},f_{1},\ldots,f_{n},f}{x_{2},x_{3},\ldots,x_{n+3}}. (33)

To prove this formula, let us consider the determinant of order2n+32n+3

|ar,s|r,s=1,2,,2n+3\left|a_{r,s}\right|\quad r,s=1,2,\ldots,2n+3 (34)

uncle ar,sa_{r,s}is the element in line arr-a and column ass-a and where

as,s={fs1(xr),r=1,2,,n+30,r=n+4,n+5,,2n+3}s=1,2,,n+2\displaystyle a_{s,s}=\left\{\begin{array}[]{ll}f_{s-1}\left(x_{r}\right),&r=1,2,\ldots,n+3\\ 0,&r=n+4,n+5,\ldots,2n+3\end{array}\right\}s=2,\ldots,n+2
fsn3(xr),\displaystyle f_{s-n-3}\left(x_{r}\right),
ar,r=\displaystyle a_{r,r}=
fsn3(xrn2),r=1,i,n+3fsn3(xrn1),r=n+4,n+i+2,,n+i+1s=n+3,n+4,,2n+3\displaystyle\begin{array}[]{ll}f_{s-n-3}\left(x_{r-n-2}\right),&r=1,i,n+3\\ f_{s-n-3}\left(x_{r-n-1}\right),&r=n+4,n+i+2,\ldots,n+i+1\\ s=n+3,n+4,\ldots,2n+3\end{array}

This determinant is equal to zero. To see this, it is enough to transform it, first adding the line a(n+2+j)(n+2+j)- that dayjj-a forj=2,3,,i1j=2,3,\ldots,i-1and the line of(n+1+j)(n+1+j)- that dayjj-a forj=i+1j=i+1, i+2,,n+2i+2,\ldots,n+2and then subtracting its column from(n+2+s)(n+2+s)-a fors=1,2,,n+1s=1,2,\ldots,n+1In this way all the elements located in the lastn+1n+1columns and firstsn+3n+3lines become null.

If we expand the determinant (34) according to Laplace's formula after the firstn+2n+2columns, we obtain formula (33).

Formula (33) is valid for2in+22\leqq i\leqq n+2It is easy to see how we can write it fori=2i=2and fori=n+2i=n+2.

If the pointsxi,i=1,2,,n+3x_{i},i=1,2,\ldots,n+3are distinct, considering (2), from formula (33) we deduce

[x1,x2,,xi1,xi+1,xi+2,,xn+3;f]=\displaystyle{\left[x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3};f\right]=}
=A[x1,x2,,xn+2;f]+B[x2,x3,,xn+3;f]\displaystyle=A\left[x_{1},x_{2},\ldots,x_{n+2};f\right]+B\left[x_{2},x_{3},\ldots,x_{n+3};f\right] (35)

where, taking into account the fact that the functions (19) form a system (I), we have

A=In(f0,f1,,fnx2,x3,,xi1,xi+1,xi+2,,xn+3)In(f0,f1,,fn,fn+1x1,x2,,xn+2)In(f0,f1,,fnx2,x3,,xn+2)In(f1,,fn,fn+1x1,x2,,xi1,xi+1,xi+2,,xn+3)\displaystyle A=\frac{V\binom{f_{0},f_{1},\ldots,f_{n}}{x_{2},x_{3},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3}}V\binom{f_{0},f_{1},\ldots,f_{n},f_{n+1}}{x_{1},x_{2},\ldots,x_{n+2}}}{V\binom{f_{0},f_{1},\ldots,f_{n}}{x_{2},x_{3},\ldots,x_{n+2}}V\binom{f_{1},\ldots,f_{n},f_{n+1}}{x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3}}} (36)
B=In(f0,f1,,fnx1,x2,,xi1,xi+1,xi+2,,xn+2)In(f0,f1,,fn,fn+1x2,x3,,xn+3)f0,f1,,fn,fn+1\displaystyle B=\frac{V\binom{f_{0},f_{1},\ldots,f_{n}}{x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+2}}V\binom{f_{0},f_{1},\ldots,f_{n},f_{n+1}}{x_{2},x_{3},\ldots,x_{n+3}}}{f_{0},f_{1},\ldots,f_{n},f_{n+1}} (37)
In(f0,f1,,fnx2,x3,,xn+2)In(x1,x2,,xi1,xi+1,xi+2,,xn+3)\displaystyle V\binom{f_{0},f_{1},\ldots,f_{n}}{x_{2},x_{3},\ldots,x_{n+2}}V\left(\begin{array}[]{c}x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3}\end{array}\right)

If in (35) we putf=fn+1f=f_{n+1}, we findA+B=1A+B=1. But ifx1<x2<<<xn+3(1<i<n+3)x_{1}<x_{2}<\ldots<<x_{n+3}(1<i<n+3), theorem 1 shows us that the coefficientsA,BA,B, which are independent of the functionff, are positive. It follows that ifx1<x2<<xn+3(1<i<n+3)x_{1}<x_{2}<\ldots<x_{n+3}(1<i<n+3), the divided difference[x1,x2,,xi1,xi+1,xi+2,,xn+3;f]\left[x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3};f\right], is a generalized arithmetic mean (with positive weights) of the divided differences[x1,x2,,xn+2;f],[x2,x3,,xn+3;f]\left[x_{1},x_{2},\ldots,x_{n+2};f\right],\left[x_{2},x_{3},\ldots,x_{n+3};f\right].

In particular, in the case of (21), (21') we find the formula for the average of the ordinary divided differences

[x1,x2,,xi1,xi+1,xi+2,,xn+3;f]==(xix1)[x1,x2,,xn+2;f]+(xn+3xi)[x2,x3,,xn+3;f]xn+3x1.\begin{gathered}{\left[x_{1},x_{2},\ldots,x_{i-1},x_{i+1},x_{i+2},\ldots,x_{n+3};f\right]=}\\ =\frac{\left(x_{i}-x_{1}\right)\left[x_{1},x_{2},\ldots,x_{n+2};f\right]+\left(x_{n+3}-x_{i}\right)\left[x_{2},x_{3},\ldots,x_{n+3};f\right]}{x_{n+3}-x_{1}}.\end{gathered}
  1. 13.

    From the formula (35) of the average we deduce the more general property expressed by

THEOREM 6. - Ifx1<x2<<xmx_{1}<x_{2}<\ldots<x_{m}saintmn+2m\geq n+2his pointsANDE, the divided difference[xi1,xi2,,xin+2;f](1=i1<i2<<<in+2=m)\left[x_{i_{1}},x_{i_{2}},\ldots,x_{i_{n+2}};f\right]\quad\left(1=i_{1}<i_{2}<\ldots<\right.\left.<i_{n+2}=m\right)onn+2n+2between these points is a generalized arithmetic mean (with convenient positive weights) of the divided differences

[xi,xi+1,,xi+n+1;f],i=1,2,,mn1\left[x_{i},x_{i+1},\ldots,x_{i+n+1};f\right],\quad i=1,2,\ldots,m-n-1 (38)

eachn+2n+2consecutive points in the sequence of pointsxix_{i}So
we have

[xi1,xi2,,xin+2;t]=i=1mn1Ai[xi,xi+1,,xi+n+1;t]\left[x_{i_{1}},x_{i_{2}},\ldots,x_{i_{n+2}};t\right]=\sum_{i=1}^{m-n-1}A_{i}\left[x_{i},x_{i+1},\ldots,x_{i+n+1};t\right] (39)

coefficientsAiA_{i}being positive, independent of functionffand of a sum equal to 1.

The proof presents no difficulty. It can be done exactly as in the particular case (21), (21') [14], by complete induction on the numbermmof pointsxix_{i}The positivity of the coefficients is a consequence
of this proof ifx1<x2<<xmx_{1}<x_{2}<\ldots<x_{m}(and the fact thati1=1i_{1}=1, in+2=mi_{n+2}=m ).

In addition to the hypotheses of theorem 6, the following inequalities are also deduced:

mini=1,2,,mn1\displaystyle\min_{i=1,2,\ldots,m-n-1} ([xi,xi+1,,xi+n+1;f])[xi1,xi2,,xin+2;f]\displaystyle\left(\left[x_{i},x_{i+1},\ldots,x_{i+n+1};f\right]\right)\leqq\left[x_{i_{1}},x_{i_{2}},\ldots,x_{i_{n+2}};f\right]\leqq
maxi=1,2,,mn1([xi,xi+1,,xi+n+1;f])\displaystyle\leqq\max_{i=1,2,\ldots,m-n-1}\left(\left[x_{i},x_{i+1},\ldots,x_{i+n+1};f\right]\right) (40)

These equalities can only occur at the same time, namely if and only if the divided differences (38) have the same value.CC, so if and only if for the functionfCfn+1f-Cf_{n+1}, these divided differences are all zero. We know that for this it is necessary and sufficient that the functionffto depend linearly on the functionsfi,i=0,1,,n+1f_{i},i=0,1,\ldots,n+1on the pointsxi,i=1,2,,mx_{i},i=1,2,\ldots,m.
14. The previous results allow us to demonstrate, under the same assumptions,

THEOREM 7. - If the function f is continuous on the intervalANDEand ifxi,i=1,2,,n+2x_{i},i=1,2,\ldots,n+2saintn+2n+2distinctive points ofANDE, we can find, inside the smallest interval containing the pointsxix_{i}, a pointx\xiso that in any neighborhood of this point there existsn+2n+2distinct pointsxiANDx_{i}^{\prime}\in E,i=1,2,,n+2i=1,2,\ldots,n+2for which we have the equality

[x1,x2,,xn+2;f]=[x1,x2,,xn+2;f]\left[x_{1},x_{2},\ldots,x_{n+2};f\right]=\left[x_{1}^{\prime},x_{2}^{\prime},\ldots,x_{n+2}^{\prime};f\right] (41)

We will first demonstrate that in (41) we can choose the pointsxix_{i}^{\prime}inside the smallest interval containing the pointsxix_{i}and in an interval of length less than a positive numbere\varepsilonsome date.

We can assume from the outset thatx1<x2<<xn+2(n0)x_{1}<x_{2}<\ldots<x_{n+2}(n\geqq 0). We divide each of the intervals[xj,xj+1],j=1,2,,n+1\left[x_{j},x_{j+1}\right],j=1,2,\ldots,n+1 in mmequal parts,mmbeing a natural number>2>2and which checks the inequality

m>n+1emaxi=1,2,,n+1(xj+1xj).m>\frac{n+1}{\varepsilon}\max_{i=1,2,\ldots,n+1}\left(x_{j+1}-x_{j}\right). (42)

Whetherand1<and2<<and(n+1)m+1y_{1}<y_{2}<\ldots<y_{(n+1)m+1}all the division points thus obtained. We therefore havexi=and(i1)m+1,i=1,2,,n+2x_{i}=y_{(i-1)m+1},\quad i=1,2,\ldots,n+2and, finally, considering (42),

andi+n+1andin+1mmaxj=1,2,,n+1(xi+1xj)<e\displaystyle y_{i+n+1}-y_{i}\leq\frac{n+1}{m}\max_{j=1,2,\ldots,n+1}\left(x_{i+1}-x_{j}\right)<\varepsilon (43)
i=1,2,,(n+1)mn\displaystyle i=1,2,\ldots,(n+1)m-n

Whether[andr,andr+1,,andr+n+1;f]\left[y_{r},y_{r+1},\ldots,y_{r+n+1};f\right]one of the smallest and[ands,ands+1,,ands+n+1;/]\left[y_{s},y_{s+1},\ldots,y_{s+n+1};/\right]one of the biggest differences divided[andi,andi+1,,andi+n+1;f]\left[y_{i},y_{i+1},\ldots,y_{i+n+1};f\right],i=1,2,,(n+1)mni=1,2,\ldots,(n+1)m-n. Formula (40) gives us

[andr,andr+1,,andr+n+1;f][x1,x2,,xn+2;f]\displaystyle{\left[y_{r},y_{r+1},\ldots,y_{r+n+1};f\right]\leqq\left[x_{1},x_{2},\ldots,x_{n+2};f\right]\leqq}
[ands,ands+1,,ands+n+1;f]\displaystyle\leqq\left[y_{s},y_{s+1},\ldots,y_{s+n+1};f\right] (44)

We will distinguish two cases:
Case 1. The equalities do not hold in (44). Then based on Theorem 4, for

A=[andr,andr+1,,andr+n+1;f]B=ands,ands+1,,ands+n+1;f]C=[x1,x2,,xn+2;f]\begin{gathered}\left.A=\left[y_{r},y_{r+1},\ldots,y_{r+n+1};f\right]\quad B=\mid y_{s},y_{s+1},\ldots,y_{s+n+1};f\right]\\ C=\left[x_{1},x_{2},\ldots,x_{n+2};f\right]\end{gathered}

the property results if we take into account the observation made during the proof of Theorem 1 and taking into account (43). The hypotheses of Theorem 1 are satisfied here.

Case 2. The inequalities (44) both become equalities. We then have[x1,x2,,xn+2;f]=[and2,and3,,andn+3;f]\left[x_{1},x_{2},\ldots,x_{n+2};f\right]=\left[y_{2},y_{3},\ldots,y_{n+3};f\right], wherex1<and2<andn+3<xn+2x_{1}<y_{2}<y_{n+3}<x_{n+2}and the property still follows from (43).

It is now easy to prove the existence of the pointx\xiThe previous reasoning shows us that we can find the strings ofn+2n+2puncturex1(j)<x2(j)<<xn+2(j)x_{1}^{(j)}<x_{2}^{(j)}<\ldots<x_{n+2}^{(j)},j=1,2,j=1,2,\ldotsso that, assumingx1<x2<<xn+2x_{1}<x_{2}<\ldots<x_{n+2}, let's have

[x1,x2,,xn+2;f]=[x1(j),x2(j),,xn+2(j);f]j=0,1,;xi(0)=xi;i=1,2,,n+2\begin{gathered}{\left[x_{1},x_{2},\ldots,x_{n+2};f\right]=\left[x_{1}^{(j)},x_{2}^{(j)},\ldots,x_{n+2}^{(j)};f\right]}\\ j=0,1,\ldots;x_{i}^{(0)}=x_{i};i=1,2,\ldots,n+2\end{gathered}

and

x1(j)<x1(j+1),xn+2(j+1)<xn+2(j),xn+2(j+1)x1(j+1)12(xn+2(j)x1(j)),j=0,1,x_{1}^{(j)}<x_{1}^{(j+1)},x_{n+2}^{(j+1)}<x_{n+2}^{(j)},x_{n+2}^{(j+1)}-x_{1}^{(j+1)}\leqq\frac{1}{2}\left(x_{n+2}^{(j)}-x_{1}^{(j)}\right),j=0,1,\ldots

The common pointx\xiof closed intervals[x1(j),xn+2(j)],j=1,2,\left[x_{1}^{(j)},x_{n+2}^{(j)}\right],j=1,2,\ldotscheck the property you are looking for.

It is seen that the pointx\xialso enjoys the property that the points can always be foundxix_{i}^{\prime}so thatx\xito be inside the smallest interval containing these points (we say thatx\xiseparate the pointsxi)\left.x_{i}^{\prime}\right).

Proprietatea exprimată de teorema 7 , cel puțin în cazul particular (21), (21’), se datoreşte lui A. Cauchy [2]
15. Putem completa teorema 7, observînd că putem totdeauna alege punctele xix_{i}^{\prime} astfel ca ele să fie echidistante. Aplicînd proprietatea funcției f[x1,x2,,xn+2;f]fn+1f-\left[x_{1},x_{2},\ldots,x_{n+2};f\right]f_{n+1}, se vede că este suficient să demonstrăm că dacă avem

[x1,x2,,xn+2;t]=0,x1<x2<<xn+2\left[x_{1},x_{2},\ldots,x_{n+2};t\right]=0,\quad x_{1}<x_{2}<\ldots<x_{n+2} (45)

putem găsi n+2n+2 puncte echidistante xi,i=1,2,,n+2x_{i}^{\prime},i=1,2,\ldots,n+2, cuprinse în intervalu1 închis [x1,xn+2]\left[x_{1},x_{n+2}\right] astfel ca să avem (41).

Vom distinge două cazuri :
Cazul 1. Printre diferentele divizate pe noduri echidistante şi cuprinse în [ x1,xn+2x_{1},x_{n+2} ], există cel puțin una care este pozitivă și cel puțin una care este negativă. In acest caz proprietatea rezultă deoarece prin procedeul întrebuintat la demonstrarea teoremei l, se poate construi o diferentă divizată pe noduri echidistante și care să fie nulă.

Cazul 2. Toate diferenţele divizate pe noduri echidistante și cuprinse în [x1,xn+2]\left[x_{1},x_{n+2}\right] sînt de același semn. Vom arăta că atunci funcţia ff, presupusă
continuă, este neconcavă sau neconvexă pe [x1,xn+2]\left[x_{1},x_{n+2}\right]. Pentru fixarea ideilor, să presupunem că diferentele divizate pe noduri echidistante sînt toate 0\geqq 0 (sau toate 0\leqq 0 ). Din teorema 6 rezultă că toate diferentele divizate pe noduri care se divid rational (rapoartele mutuale ale distanțelor dintre noduri sînt rationale) sînt 0\geqq 0 (sau 0\leqq 0 ). Din continuitatea funcției ff rezultă atunci că toate diferențele divizate sînt 0\geqq 0 (sau 0\leqq 0 ). Funcția ff este deci neconcavă (sau neconvexă) pe [x1,xn+2]\left[x_{1},x_{n+2}\right].

Proprietatea căutată rezultă atunci din
L ema 1. - Dacă funcția continuă ff este neconcavă pe intervalul [x1,xn+2]\left[x_{1},x_{n+2}\right] si dacă avem (45), toate diferentele divizate ale functiei pe noduri apartinînd lui [x1,xn+2]\left[x_{1},x_{n+2}\right], sînt nule.

Pentru demonstrare să presupunem că proprietatea nu este adevărată. Există atunci puncte distincte xix_{i}^{\prime} astfel ca [x1,x2,,xn+2;f]>0\left[x_{1}^{\prime},x_{2}^{\prime},\ldots,x_{n+2}^{\prime};f\right]>0. Reunirea mulțimilor de puncte xi,xi,i=1,2,,m+2x_{i},x_{i}^{\prime},i=1,2,\ldots,m+2 formează un şir de cel putin n+3n+3 si cel mult 2n+42n+4 puncte distincte ale intervalului [x1,xn+2]\left[x_{1},x_{n+2}\right]. Aplicînd teorema 6, împreuna cu consecințele ei relative 1 a cazurile cînd egalitatea are loc în (40), succesiv şirurilor parti x1,x2,,xn+2;x1,x2,,xn+2x_{1}^{\prime},x_{2}^{\prime},\ldots,x_{n+2}^{\prime};x_{1},x_{2},\ldots,x_{n+2}, se ajunge la o contradictic cu (45).

In fine, dacă tinem seamă de rezultatele lui D. V. W id d e r [28], putem afirma că egalitatea (41) poate fi realizată cu noduri xix_{i}^{\prime} echidistante, distanța δ\delta a două noduri consecutive fiind suficient de mică. In cazul cînd intervalu1 E[x1,xn+2]E\supset\left[x_{1},x_{n+2}\right], teorema de medie a lui D. V. Widder afirmă că se poate realiza rezultatul precedent cu noduri xix_{i}^{\prime} echidistante pentru care distanta δ\delta este mai mică decît un număr fix independent de funcția ff.
16. Înainte de a merge mai departe să observăm că teorema 7 poate să nu aibă loc clacă funcţiile (19) nu formează un sistem (I).

Să considerăm funcţiile fi=xi+1i=0,1,,nf_{i}=x^{i+1}i=0,1,\ldots,n pe un interval EE care contine punctul 0. Aceste functii nu formează un sistem (I). Funcția fn+1=1f_{n+1}=1 este convexă sau concavă (convexă dacă nn este impar şi concavă dacă nn este par), în sensul definiției nesimetrice a convexității. Avem [x1,x2,,xn+2;xn+2]=(1)n+2x1x2,,xn+2\left[x_{1},x_{2},\ldots,x_{n+2};x^{n+2}\right]=(-1)^{n+2}x_{1}x_{2},\ldots,x_{n+2}. Dacă deci pentru funcţia continuă xn+2x^{n+2} avem egalitatea (41), unde unul dintre punctele xix_{i} coincide cu 0 , unul din punctele xix_{i}^{\prime} va coincide în mod necesar cu 0 . Rezultă ușor că teorema 7 nu se aplică.
17. Rezultatele acestui § se pot extinde și la cazul cînd nodurile nu sînt distincte.

Să presupunem că nu numai funcțille (18) dar și funcțiile (19) formează un sistem (I) regulat de ordinul kk.

Teorema 6 se poate extinde la cazul cînd punctele x1x2xm(mn+2)x_{1}\leqq x_{2}\leqq\ldots\leqq x_{m}(m\geq n+2) nu sînt toate distincte si acelaşi punct se repeta cel mult de kk ori. Pentru cele ce urmează va fi destul să ne ocupăm de extensiunea formulei (35) şi vom arăta că această formulă rămîne valabilă dacă
x1x2xn+3x_{1}\leqq x_{2}\leqq\ldots\leqq x_{n+3}, acelaşi punct repetindu-se cel mult de kk ori. Mai mult încă, coeficientii respectivi A,BA,B, de sumă egală cu 1 , rămîn independenți de funcţia ff și sint pozitivi dacă x1<xi<xn+3x_{1}<x_{i}<x_{n+3} (ceea ce implică 1<i<n+31<i<n+3 )

Formula căutată se scrie unde putem presupune p3p\geqq 3 şi avem kr=kr′′=kr′′′=kr,r=2,3,j1,j+1,,p1,2jp1(k_{r}^{\prime}=k_{r}^{\prime\prime}=k_{r}^{\prime\prime\prime}=k_{r},r=2,3,\ldots j-1,j+1,\ldots,p-1,2\leqq j\leqq p-1\quad( dacă p>3),k1=k1′′=k1kj′′=kj′′′=kj,kp=kp′′′=kp,k1′′′=k11,kj=kj1,kp′′=kp1p>3),\quad k_{1}^{\prime}=k_{1}^{\prime\prime}=k_{1}k_{j}^{\prime\prime}=k_{j}^{\prime\prime\prime}=k_{j},\quad k_{p}^{\prime}=k_{p}^{\prime\prime\prime}=k_{p},\quad k_{1}^{\prime\prime\prime}=k_{1}-1,\quad k_{j}^{\prime}=k_{j}-1,k_{p}^{\prime\prime}=k_{p}-1 ; 1krk,r=1,2,,p,k1+k2++kp=n+31\leqq k_{r}\leqq k,r=1,2,\ldots,p,k_{1}+k_{2}+\ldots+k_{p}=n+3.

Această formulă se obține din formulele (35) - (37), presupunînd x1<x2<<xn+3x_{1}<x_{2}<\ldots<x_{n+3} şi făcînd

xk1+k2++kr1+s=xs(r)zr,s=1,2,,kr(k0=0)\displaystyle x_{k_{1}+k_{2}+\ldots+k_{r-1}+s}=x_{s}^{(r)}\rightarrow z_{r},\quad s=1,2,\ldots,k_{r}\left(k_{0}=0\right) (47)
γ=1,2,,p\displaystyle\gamma=1,2,\ldots,p
z1<z2<<zp,i=k1+k2++kj\displaystyle z_{1}<z_{2}<\ldots<z_{p},\quad i=k_{1}+k_{2}+\ldots+k_{j}

Se vede ușor cum trebuie modificată formula dacă k1=1,kj=1k_{1}=1,k_{j}=1 sau kp=1k_{p}=1.

Rezultă imediat că A,BA^{*},B^{*} sînt independenți de funcţia ff și că A0,B0,A+B=1A^{*}\geqq 0,B^{*}\geqq 0,A^{*}+B^{*}=1. Rămîne să se demonstreze că A0A^{*}\neq 0, B0B^{*}\neq 0. Pentru coeficientul AA^{*} acest lucru rezultă observînd că, cu ajutorul notatiilor (47), el se obtine din membrul al doilea al formulei (36) împărțind cei 4 determinanți (1) care figurează la numărător și la numitor prin expresia (5) ( n=n+3n=n+3 ) multiplicată respectiv cu

V(x1(1),x2(1),,xk11(1))V(x1(1),x2(1),,xk1(1))V(x1(j),x2(j),,xk11(j))V(x1(j),x2(j),,xkj(j)),V(x1(p),x2(p),,xkp1(p))V(x1(p),x2(p),,xkp(p))\displaystyle\frac{V\left(x_{1}^{(1)},x_{2}^{(1)},\ldots,x_{k_{1}-1}^{(1)}\right)}{V\left(x_{1}^{(1)},x_{2}^{(1)},\ldots,x_{k_{1}}^{(1)}\right)}\cdot\frac{V\left(x_{1}^{(j)},x_{2}^{(j)},\ldots,x_{k_{1}-1}^{(j)}\right)}{V\left(x_{1}^{(j)},x_{2}^{(j)},\ldots,x_{k_{j}}^{(j)}\right)},\frac{V\left(x_{1}^{(p)},x_{2}^{(p)},\ldots,x_{k_{p}-1}^{(p)}\right)}{V\left(x_{1}^{(p)},x_{2}^{(p)},\ldots,x_{k_{p}}^{(p)}\right)}
V(x1(1),x2(1),,xk11(1))V(x1(1),x2(1),,xk1(1))V(x1(p),x2(p),,xkp1(p))V(x1(p),x2(p),,xkp(p)),V(x1(j),x2(j),,xk11(j))V(x1(j),x2(j),,xkj(j))\displaystyle\frac{V\left(x_{1}^{(1)},x_{2}^{(1)},\ldots,x_{k_{1}-1}^{(1)}\right)}{V\left(x_{1}^{(1)},x_{2}^{(1)},\ldots,x_{k_{1}}^{(1)}\right)}\cdot\frac{V\left(x_{1}^{(p)},x_{2}^{(p)},\ldots,x_{k_{p-1}}^{(p)}\right)}{V\left(x_{1}^{(p)},x_{2}^{(p)},\ldots,x_{k_{p}}^{(p)}\right)},\frac{V\left(x_{1}^{(j)},x_{2}^{(j)},\ldots,x_{k_{1}-1}^{(j)}\right)}{V\left(x_{1}^{(j)},x_{2}^{(j)},\ldots,x_{k_{j}}^{(j)}\right)}

și trecînd la limită. Mai sus determinantii lui Vandermonde care nu au sens (pentru k1=1,kf=1k_{1}=1,k_{f}=1 sau kp=1k_{p}=1 ) sînt înlocuiţi cu 1 . Dacă se efectuează aceste împărtiri, pe de o parte 1111 se schimbă valoarea coeficientului AA și, pe de altă parte, fiecare dintre determinanții (1) astfel împărțiți tinde către o limită bine determinată și diferită de zero. Rezultă că A0A^{*}\neq 0. Se demonstrează în acelaşi fel că B0B^{*}\neq 0. Demonstrația ne mai arată că
coeficientii A,BA^{*},B^{*} ai formulei (46) sînt bine determinați prin condiția ca să fie independenți de funcţia ff. Este uşor a se scrie valorile acestor coeficienti cu ajutorul determinanților (3).
18. Putem extinde teorema 7 la cazul cînd punctele xix_{i} nu sînt toate distincte. Intr-adevăr, presupunînd pe mai departe că funcţiile (18) și (19) sînt continue și formează cîte un sistem (I) regulat de ordinul kk, teorema 7 rămîne adevărată dacă printre punctele xix_{i} același punct se repetă cel mult. de kk ori.

Pentru a demonstra această proprietate, în virtutea chiar a teoremei 7, este sufficient să demonstrăm

L, e m a 2. - Dacă, pe lîngă ipotezele precedente, printre punctele xix_{i} există exact pp puncte distincte, cu 2pn+12\leqq p\leqq n+1,
se pot găsi n+2n+2 puncte xi,i=1,2,,n+2x_{i}^{\prime},i=1,2,\ldots,n+2, astfel ca : 11^{\circ}. fiecare se repetă cel mult de kk ori, 22^{\circ}. există printre ele cel putin p+1p+1 distincte, 33^{\circ}. sînt cuprinse toate în cel mai mic interval închis care contine punctele xi,4x_{i},4^{\circ}. egalitatea (41) este verificată.

Pentru simplificarea limbajului vom zice că o diferență divizată ale cărei noduri, aranjate în ordinea lor crescătoare, au succesiv ordinele de multiplicitate k1,k2,,kp(k1+k2+kp=n+2)k_{1},k_{2},\ldots,k_{p}\left(k_{1}+k_{2}+\ldots k_{p}=n+2\right), este de tipul (k1,k2,,kp)\left(k_{1},k_{2},\ldots,k_{p}\right). Condițiile 1,21^{\circ},2^{\circ} ale lemei însemnează că diferența divizată pe nodurile xix_{i} fiind de tipul ( k1,k2,,kpk_{1},k_{2},\ldots,k_{p} ), cu 1kik,i=1,2,,p1\leqq k_{i}\leqq k,i=1,2,\ldots,p, 2pn+12\leqq p\leqq n+1, se pot găsi punctele xix_{i}^{\prime} astfel ca diferența divizată pe aceste puncte să fie de tipul ( k1,k2,,kqk_{1}^{\prime},k_{2}^{\prime},\ldots,k_{q}^{\prime} ), cu 1kik,i=1,2,,q1\leqq k_{i}^{\prime}\leqq k,i=1,2,\ldots,q, qp+1q\geqq p+1.

Să considerăm deci diferența divizată pe nodurile xix_{i} și fie ( k1,k2,,kpk_{1},k_{2},\ldots,k_{p} ) tipul, iar CC valoarea acestei diferențe divizate. Să intercalăm între primele două noduri distincte un al ( n+3n+3 )-lea nod, diferit de toate celelalte. Să aplicăm formula mediei (46) șirului de n+3n+3 puncte astfel obținute, noul nod fiind acela care este eliminat în diferența divizată din membrul întîi. În membrul al doilea figurează diferentele divizate

[u1,u2,,un+2;f],[v1,v2,,vn+2;f]\left[u_{1},u_{2},\ldots,u_{n+2};f\right],\left[v_{1},v_{2},\ldots,v_{n+2};f\right] (48)
u1u2un+2,v1v2vn+2,u1<un+2,v1<vn+2,u_{1}\leqq u_{2}\leqq\ldots\leqq u_{n+2},v_{1}\leqq v_{2}\leqq\ldots\leqq v_{n+2},u_{1}<u_{n+2},v_{1}<v_{n+2},

care sînt respectiv de tipul (k1,1,k2,k3,,kp1,kp1)\left(k_{1},1,k_{2},k_{3},\ldots,k_{p-1},k_{p}-1\right) și (k11,1,k2,k3,,kp)\left(k_{1}-1,1,k_{2},k_{3},\ldots,k_{p}\right), unde trebuie suprimat k11k_{1}-1 dacă k1=1k_{1}=1, şi kp1k_{p}-1 dacă kp=1k_{p}=1.

Trebuie acum să distingem trei cazuri :
Cazul 1. Diferentele divizate (48) au valori diferite. Atunci una are o valoare A<CA<C si cealaltă o valoare B>CB>C. Tinind seamă de felul cum - diferentă divizată (27) se obține ca limită de diferențe divizate pe noduri distincte, rezultă că putem găsi diferențele divizate

[u1,u2,,un+2;f],[v1,v2,,vn+2;f]\left[u_{1}^{\prime},u_{2}^{\prime},\ldots,u_{n+2}^{\prime};f\right],\quad\left[v_{1}^{\prime},v_{2}^{\prime},\ldots,v_{n+2}^{\prime};f\right] (49)

pe noduri distincte și ale căror valori sînt numerele A,BA^{\prime},B^{\prime} respectiv oricît de aproape de numerele A,BA,B, deci în particular, astfel ca A<C<BA^{\prime}<C<B^{\prime}.

Se poate uşor constata că putem chiar lua nodurile primei diferențe divizate (49) în intervalul ( u1,un+2u_{1},u_{n+2} ) şi nodurile celei de a doua diferente divizate în intervalul ( v1,vn+2v_{1},v_{n+2} ). Aplicînd teorema 4 diferențelor divizate ( 49 ), putem găsi o diferență divizată avînd valoarea CC. Se vede că conditiile 2,42^{\circ},4^{\circ} ale lemei sînt verificate.

Cazul 2. Avem k1+kp>2k_{1}+k_{p}>2 şi cele două diferențe divizate (48) sînt egale. Atunci ambele sînt egale cu CC şi sau prima (dacă kp>1k_{p}>1 ) sau a doua (dacă k1>1k_{1}>1 ) verifică condițiile 22^{\circ} și 44^{\circ} ale lemei.

Cazul 3. Avem k1=kp=1k_{1}=k_{p}=1 și cele două diferenţe divizate (48) sînt egale cu CC. Avem atunci o diferență divizată egală cu CC şi de tipul (1,1,k2,k3,,kp1)\left(1,1,k_{2},k_{3},\ldots,k_{p-1}\right). Cu această diferență divizată se procedează în mod analog. Se vede atunci că dacă kp1>1k_{p-1}>1, cădem peste cazu1 1 sau 2 iar dacă kp1=1k_{p-1}=1 se construieşte o diferenţă divizată egală cu CC și de tipul (1,1,1,k2,k3,,kp2)\left(1,1,1,k_{2},k_{3},\ldots,k_{p-2}\right). Deoarece cel puţin un kik_{i} este >1>1, după un număr finit de operafii de acest fel se cade asupra cazului 1 sau 2 .

Astfel condiţiile 22^{\circ} și 44^{\circ} ale lemei sînt realizate. Să observăm că în timpul demonstrației, pe de o parte nu se întrece niciodată ordinul kk de multiplicitate şi, pe de altă parte, nu se iese niciodată din cel mai mic interval care conţine punctele xix_{i}. Deci și condițiile 11^{\circ} și 33^{\circ} ale lemei sint verificate.

Lema 2 este deci demonstrată.
Din cele ce preced rezultă și
teorema 8. - Dacă functilie (18) şi functiile (19) sînt continue si /ormează sisteme (I) regulate de ordinul kk pe intervalul EE si dacă functia ff este continuă și convexă, neconcavă, neconvexă resp. concavă in raport cu Junctiile (19),
prima, a doua, a treia respectiv a patra inegalitate (25) rămîne adevărată dacă nodurile xix_{i} nu sînt toate confundate si fiecare se repetă cel mult de kk ori.

Moreover, for non-concave functions and non-convex functions, the property results simply by passing to the limit and remains true if functions (18) and (19) form completely regular systems (I), even if the pointsxix_{i}they are all confused.

Theorem 8 results from the extension of Theorem 7 given in this issue
19. Theorem 7, extended in the above way, allows us to link the structure of a linear functional of simple form to the differential properties of the functions on which it is defined. Thus we have

THEOREMA9.\mathrm{OREMA}9.-If:11^{\circ}. the functionals (18) and (19) form completely regular systems (I) on the intervalAND,2E,2^{\circ}. linear functionalR[f]R[f]is of simple form,33^{\circ}, functionf𝔽f\in\mathbb{F}has a continuous derivative of ordern+1n+1inside itANDE,
can be found, inside itANDE, a pointx\xiso that we have

F[f]=R[fn+1][x,x,,x;f].F[f]=R\left[f_{n+1}\right][\xi,\xi,\ldots,\xi;f]. (50)

The proof follows immediately from Theorem 7 and the limit properties of divided differences with multiple nodes. The pointx\xiis one of those that verifies theorem 7.

The difference divided by the second term of (50) can be calculated using formula (28) or formula (29).

We do not intend to delve into these issues further in this paper. We only recall that, in the particular case (21), (21'), we gave a generalization of Theorem 7 [18] which allows us to further specify the connection between the properties of the remainderR[f]R[f]and the differential properties of different orders of the function\not.

§ 3.

  1. 20.

    In this § we will examine some criteria that allow us to decide whether a linear functionalR[f]R[f]is or is not of simple form. We will then make applications to the remaining few approximation formulas (*).

The linear combination (11) can be used to find an approximation formula of the form (*).

WhetherA[f]A[f]a linear functional defined on the vector space(f(fformed by continuous functions defined on the intervalANDEand which have continuous derivatives onANDEof all the orders that intervene. We will assume that the functions (18) and (19) belong to (f) and, to simplify matters, that they form completely regular systems (I). Moreover, for the validity of some of the results that follow, a regularity of an order lower thann+2n+2 resp. n+1n+1is generally sufficient.

We will take as an approximation forA[f]A[f]the defined and linear functional ont^\widehat{t},

B[f]=A[L(fx)],B[f]=A[L(f\mid x)], (51)

whereL(fx)L(f\mid x)is given by (11), relative to the functions (19).
This approximation procedure is well known and has been widely studied, especially in various particular cases.

we

L(fx)=i=1pj=0k1fi,j(x)f(j)(Withi),(f(0)(x)=f(x),(k1+k2++kp=n+1),L(f\mid x)=\sum_{i=1}^{p}\sum_{j=0}^{k-1}\varphi_{i,j}(x)f^{(j)}\left(z_{i}\right),\quad\left(f^{(0)}(x)=f(x),\quad\left(k_{1}+k_{2}+\cdots+k_{p}=n+1\right),\right.

pointsWithi,i=1,2,,pz_{i},i=1,2,\ldots,pbeing distinct andfi,j,j=0,1,,ki1\varphi_{i,j},j=0,1,\ldots,k_{i}-1, i=1,2,,pi=1,2,\ldots,pbeing well-determined linear combinations of the functions (19). We then have

B[f]=i=1pj=0ki1ai,jf(j)(Withi),B[f]=\sum_{i=1}^{p}\sum_{j=0}^{k_{i}-1}a_{i,j}f^{(j)}\left(z_{i}\right), (52)

whereai,j=Afi,j],j=0,1,,ki1,i1,2,,p\left.a_{i,j}=A\mid\varphi_{i,j}\right],j=0,1,\ldots,k_{i}-1,i-1,2,\ldots,p
There is an important particular case when the rest R[f]R[f]of the approximation formula thus obtained is of simple form. We have, namely

THEOREM 10. - If:11^{\circ}. linear functionalA[f]A[f]is positive,22^{\circ}. orders of multiplicitykik_{i}of all pointsWithiz_{i}which are foundI^n\hat{\imath}ninside the intervalANDE, are even,
the restR[f]R[f]of the approximation formula (*), constructed in the way shown above, is of simple form.

functionA[f]A[f]is positive if we haveA[f]0A[f]\geqq 0, for any functionff(continuous) non-negative, the equality being true (if and) only iff=0f=0onANDE.

Formula (16) gives us

f(x)L(fx)=ψ(x)[x1,x2,,xn+1,x;f]f(x)-L(f\mid x)=\psi(x)\left[x_{1},x_{2},\ldots,x_{n+1},x;f\right] (53)

pointsxix_{i}having the same meaning as in (16). In this formula we have

ψ(x)=In(f0,f1,,fn,fn+1x1,x2,,xn+1,x):In(f0,f1,,fnx1,x2,,xn+1)\psi(x)=V\binom{f_{0},f_{1},\ldots,f_{n},f_{n+1}}{x_{1},x_{2},\ldots,x_{n+1},x}:V\binom{f_{0},f_{1},\ldots,f_{n}}{x_{1},x_{2},\ldots,x_{n+1}}

ifxxis different from one of the nodesxix_{i}.
Formula (53) is true for anyxANDx\in E, provided that the second term is replaced by 0 ifxxcoincides with one of the nodesxix_{i}We have

R[f]=A[fL(fx)]=A[ψ[x1,x2,,xn+1,x;f]]R[f]=A[f-L(f\mid x)]=A\left[\psi\left[x_{1},x_{2},\ldots,x_{n+1},x;f\right]\right]

and the rest is of simple form because:11^{\circ}. the divided difference appearing in the second member of formula (53) is, by virtue of theorem 8, positive ifffis a convex function, except at mostn+1n+1points (pointsxix_{i}) of hisAND.2E.2^{\circ}. the function is not identically zero and does not change sign onANDE ; this property results from theorem 2 by taking the limit,33^{\circ}. functionfL(fx)f-L(f\mid x)is continuous onANDEIt follows that this latter function is not identically zero and that it does not change sign onANDEifffis a convex function. Theorem 10 follows immediately.

The remainder is of the form (30) and if(n+1)(n+1)-derivative offfexists and is continuous within itANDE, of the same form as indicated in theorem 9. The constantK=R[fn+1]K=R\left[f_{n+1}\right]can also be calculated using the formulaK=R[ψ]K=R[\psi], or using the formulaK=R[ψ+f]K=R[\psi+\varphi], wheref\varphiis a linear combination of the functions (19).

It is easy to generalize the previous result in the case when it is assumed that the functions (18) and (19) form regular (I) systems of orderkmax(k1,k2,,kp)k\geqq\max\left(k_{1},k_{2},\ldots,k_{p}\right)Finally, it is clear that an analogous property exists for a functionalA[f]A[f]negative, for which we therefore haveA[f]0A[f]\leqq 0for any functionffnonnegative, the equality being true only forj=0j=0.

We note that many classical approximation formulas, for example so-called numerical (or mechanical) quadrature formulas, are of the previous form. We will recall some of these formulas below.
21. The well-known numerical quadrature formula

A[f]=02pf(x)𝑑x=2pm+1i=0mf(2ipm+1)+R[f]A[f]=\int_{0}^{2\pi}f(x)dx=\frac{2\pi}{m+1}\sum_{i=0}^{m}f\left(\frac{2i\pi}{m+1}\right)+R[f] (54)

wheremmis a natural number andffa continuous function on the closed interval[0,2p][0,2\pi], is of the previous form.

In this caseR[f]R[f]is null on the functions

f0=1,f2i1=cosix,f2i=sinix,i=1,2,,mf_{0}=1,f_{2i-1}=\cos ix,f_{2i}=\sin ix,\quad i=1,2,\ldots,m (55)

to which the functions (19) now reduce. We have already shown that the functions (55) form a completely regular system (I) on the interval[0,2p)[0,2\pi)This property is equivalent to the fact that a trigonometric polynomial of degreemmcannot have2m+12m+1distinct roots or not in the interval[0,2p)[0,2\pi), without being identically null.

Let us also consider the function

f2m+1=xf_{2m+1}=x (\prime)

Then the functions (55), (55') together also form a completely regular system (I) on[0,2p)[0,2\pi)Indeed, a non-identical non-zero linear combinationf\varphiof functions ( 55 ), (5555^{\prime}) cannot have more than2m+12m+1distinct roots or not in[0,2p)[0,2\pi)Otherwise, the derivativef\varphi^{\prime}, which is a trigonometric polynomial of degreemm, would have at least2m+12m+1distinct roots or not in[0,2p)[0,2\pi)It would follow thatf=0\varphi^{\prime}=0, so thatf\varphiis a constant0\neq 0, which is impossible.

Formula (54) is of the previous form. To obtain it, it is sufficient to take the functionL(fx)L(f\mid x)(Lagrange-Hermite type trigonometric interpolation polynomial) relative to the simple node 0 and to the nodesdinbland2ipm+1duble\frac{2i\pi}{m+1}, i=1,2,,mi=1,2,\ldots,mIt is easy to verify that (54) is the only formula of the form

02pf(x)𝑑x=Af(0)+i=1m[aif(2ipm+1)+bif(2ipm+1)]+R[f]\int_{0}^{2\pi}f(x)dx=Af(0)+\sum_{i=1}^{m}\left[\alpha_{i}f\left(\frac{2i\pi}{m+1}\right)+\beta_{i}f^{\prime}\left(\frac{2i\pi}{m+1}\right)\right]+R[f]

in whichA,ai,biA,\alpha_{i},\beta_{i}are independent of the functionffand the remainder of which cancels out on the functions (55).

The rest of formula (54) is of simple form and we have

R[t]=2p2m+1[x1,x2,,x2m+2;t]R[t]=\frac{2\pi^{2}}{m+1}\left[\xi_{1},\xi_{2},\ldots,\xi_{2m+2};t\right]

functionffbeing continuous on[0,2p][0,2\pi]and having a continuous derivative on(0,2p)(0,2\pi). The pointsxi(0,2p),i=1,2,,2m+2\xi_{i}\in(0,2\pi),i=1,2,\ldots,2m+2are distinct.

Ifffhas a continuous derivative of order2m+12m+1on (0,2p0,2\pi), we find the remainder given by J. Radon [21]. In our case

[x,x,,x;f]=1(m!)2[ddx(d2dx2+12)(d2dx2+22)(d2dx2+m2)f]xx[\xi,\xi,\ldots,\xi;f]=\frac{1}{(m!)^{2}}\left[\frac{d}{dx}\left(\frac{d^{2}}{dx^{2}}+1^{2}\right)\left(\frac{d^{2}}{dx^{2}}+2^{2}\right)\cdots\left(\frac{d^{2}}{dx^{2}}+m^{2}\right)f\right]_{x\rightarrow\xi}
  1. 22.

    Formula (54) is the trigonometric analogue of Gauss's classical numerical integration formula,

11f(x)𝑑x=i=1maif(gi)+R[t]\int_{-1}^{1}f(x)dx=\sum_{i=1}^{m}\alpha_{i}f\left(\zeta_{i}\right)+R[t] (56)

wheregi,i=1,2,,m\zeta_{i},i=1,2,\ldots,mare the roots, all real, distinct and contained in (1,1-1,1), of the polynomial

P(x)=m!(2m)!dmdxm(x21)mP(x)=\frac{m!}{(2m)!}\cdot\frac{d^{m}}{dx^{m}}\left(x^{2}-1\right)^{m}

and whose remainder cancels out on any polynomial of degree2m12m-1. Formula (56) is relative to the particular case (21), (21') and to obtain it it is sufficient to take the functionL(fx)L(f\mid x)(LagrangeHermite interpolation polynomial) relative to double knotsgi,i=1,2,,m\zeta_{i},i=1,2,\ldots,mBy virtue of Theorem 10 the remainder is of simple form and we have (n=2m1n=2m-1 ),

R[xn+1]=R[P2]=11P2𝑑x=22m+1(m!)4(2m+1)[(2m)!]2R\left[x^{n+1}\right]=R\left[P^{2}\right]=\int_{-1}^{1}P^{2}dx=\frac{2^{2m+1}(m!)^{4}}{(2m+1)[(2m)!]^{2}}

The remainder is therefore of the form

R[f]=22m+1(m!)4(2m+1)[(2m)!]2[x1,x2,,x2m+1;f]R[f]=\frac{2^{2m+1}(m!)^{4}}{(2m+1)[(2m)!]^{2}}\left[\xi_{1},\xi_{2},\ldots,\xi_{2m+1};f\right] (57)

the function being continuous on[1,1][-1,1]and having a continuous derivative on(1,1)(-1,1). The pointsxi(1,1),i=1,2,,2m+1\xi_{i}\in(-1,1),i=1,2,\ldots,2m+1are distinct.

Existence and continuity of the derivative of the functionffin the study of the simplicity of the rest of the formulas (54) and (56) are imposed by the particular method by which we obtained this simplicity. It can be shown that the hypothesis of the existence of the derivative is superfluous, which we will effectively show below for Gauss's formula.
23. Let us consider a linear functional of the form

R[f]=i=1pi=0j=0ki1ci,jf(j)(Withi),R[f]=\sum_{i=1}^{p}\sum_{\begin{subarray}{c}i=0\\ j=0\end{subarray}}^{k_{i}-1}c_{i,j}f^{(j)}\left(z_{i}\right), (58)

wheren+2k1,k2,,kp1,k1+k2++kp=mn+2,With1<With2<<Withpn+2\geqq k_{1},k_{2},\ldots,k_{p}\geqq 1,k_{1}+k_{2}+\ldots+k_{p}=m\geqslant n+2,z_{1}<z_{2}<\ldots<z_{p}are points of the intervalANDEandci,jc_{i,j}, are independent coefficients of the functionff. Definition spaceffof the functional is formed by the functions whose derivative of the max order(k11,k21,,kp1)\left(k_{1}-1,k_{2}-1,\ldots,k_{p}-1\right)exists and is continuous onANDE. We assume that the functions (18) and (19) belong tof𝑓\underset{f}{f}and form regular systems of max order(k1,k2,,kp)\left(k_{1},k_{2},\ldots,k_{p}\right).

Whetherx1x2xmx_{1}\leqq x_{2}\leqq\ldots\leqq x_{m}pointsWithiz_{i}counted with their respective multiplicity orders. The functional (58) can also be written in the form

R[f]=R1[f]+i=1m1mi[xi,xi+1,,xi+n+1;f]R[f]=R_{1}[f]+\sum_{i=1}^{m-1}\mu_{i}\left[x_{i},x_{i+1},\ldots,x_{i+n+1};f\right]

wheremi\mu_{i}are independent coefficients of the functionff.
R1[f]R_{1}[f]is an expression analogous to (58), but where only the values ​​of the function appearffand its successive derivatives on the firstn+1n+1knotsx1,x2,,xn+1x_{1},x_{2},\ldots,x_{n+1}(distinct or not). If one of these latter nodes is repeated bykktimes, inR1[f]R_{1}[f]only the value of the function and its primes appear linearly (possibly with zero coefficients)k1k-1derivatives at this point.

If we observe that in the divided divergence (27) (whereWithiz_{i}are distinct) the coefficients off(ki1)(Withi),i=1,2,,pf^{\left(k_{i}-1\right)}\left(z_{i}\right),i=1,2,\ldots,pare always different from zero, we see that the coefficientsmi\mu_{i}and the linear functionalR𝟏[f]R_{\mathbf{1}}[f]are completely determined by the linear functional (58).

For the linear functional (58) to be zero on the functions (19), it is necessary and sufficient thatR1[f]R_{1}[f]be identically null. The condition is obviously sufficient (formula (23)). It is also necessary because the coefficients of can be successively nullifiedR1[f]R_{1}[f], choosing forffa convenient linear combination of functions (19).

From here the formula first resultsR1[f]=R[L(x1,x2,,xn+1;fx)]R_{1}[f]=R\left[L\left(x_{1},x_{2},\ldots,x_{n+1};f\mid x\right)\right]and then

Lemma 3. - For the linear functional (58) to be zero on the functions (19), it is necessary and sufficient that it be of the form

R[f]=i=1mn1mi[xi,xi+1,,xi+n+1;f]R[f]=\sum_{i=1}^{m-n-1}\mu_{i}\left[x_{i},x_{i+1},\ldots,x_{i+n+1};f\right] (59)

where the coefficientsmi\mu_{i}are well-determined and independent of functiontt
From here we deduce
THEOREM 11. - If : 11^{\circ}. functions (18) and (19) form completely regular systems (I) on the intervalAND,2E,2^{\circ}. the linear functional (58) is zero on the functional (19),33^{\circ}In expression (59) of this linear functional, the coefficientsmi\mu_{i}are of the same sign (all0\geqq 0or all0\leqq 0 ), 44^{\circ}. assumingx1x2xmx_{1}\leqq x_{2}\leqq\ldots\leqq x_{m}, we have

i=1m1mi(xi+n+1mxi)0,\sum_{i=1}^{m-1}\mu_{i}\left(x_{i+n+1}^{m}-x_{i}\right)\neq 0,

the linear functional (58) is of the simple form.
We assume herem>n+2m>n+2. The condition44^{\circ}means that at least one of the coefficientsmi\mu_{i}is0\neq 0and at the same time the nodesxi,xi+1,,xi+n+1x_{i},x_{i+1},\ldots,x_{i+n+1}, corresponding to such a coefficient, are not all confused. The proof of Theorem 11 follows easily. Indeed, for a convex function all the terms of the sum (59) are of the same sign and at least one is0\neq 0.

The result is also valid form=n+2m=n+2, suppressing the condition in the theorem33^{\circ}.

It is easy to see that the conditionn+2k1,k2,,kpn+2\geqq k_{1},k_{2},\ldots,k_{p}is essential. In particular, this condition is satisfied by the linear functional (59). However, in the case when the condition is not satisfied, the linear functional (58) may not be of the indicated form and therefore Theorem 11 may not hold.
24. In the particular case (21), (21') we can give more complete results. In this case we can distinguish convexities from successive ordersn=1,0,1,n=-1,0,1,\ldotsand the notion of simplicity of a linear functional is related to its degree of accuracy.

It is said that the linear functionalR[f]R[f](or the corresponding approximation formula that has this remainder) has the degree of accuracy (integer)n1n\geqq-1ifR[xi]=0,i=0,1,,n,R[xn+1]0R\left[x^{i}\right]=0,i=0,1,\ldots,n,R\left[x^{n+1}\right]\neq 0Here we putn=1n=-1ifR[1]0R[1]\neq 0andn=n=\inftyifR[xi]=0R\left[x^{i}\right]=0fori=0,1,i=0,1,\ldotsThe degree of accuracy (finite or not) is always well determined. In what follows we consider only linear functionals having a finite degree of accuracy and which are defined, in particular, on any polynomial. For such a linear functional to have a finite degree of accuracy, it is necessary and sufficient that it is not zero on any polynomial. For example, the linear functional (58), assumed non-identically zero (more precisely with coefficientsci,jc_{i,j}not all zero), has a finite degree of accuracy. Indeed, without restricting generality, it can be assumed that one of the coefficientsci,ki1,i=1,2,,pc_{i,k_{i}-1},i=1,2,\ldots,pis0\neq 0Either, for fixing ideas,cr,kr10c_{r},k_{r-1}\neq 0. It can then be easily seen thatR[1xWithri=1p(xWithi)ki]0R\left[\frac{1}{x-z_{r}}\prod_{i=1}^{p}\left(x-z_{i}\right)^{k_{i}}\right]\neq 0.

For a linear functional to be of simple form, it is necessary for it to have a finite degree of accuracy.

We will prove
Theorem 12. - Assumingx1x2xn+3x_{1}\leqq x_{2}\leqq\ldots\leqq x_{n+3}, because the linear junction

Rf]=m1[x1,x2,,xn+2;f]+m2[x2,x3,,xn+3;f],R\mid f]=\mu_{1}\left[x_{1},x_{2},\ldots,x_{n+2};f\right]+\mu_{2}\left[x_{2},x_{3},\ldots,x_{n+3};f\right], (60)

(the coefficientsm1,m2\mu_{1},\mu_{2}being independent of the functionff) to be of simple form, it is necessary and sufficient that one of the conditions:
11^{\circ}. The nodesxix_{i}they are not all confused andm1=m20\mu_{1}=-\mu_{2}\neq 0.
2.(xn+2x1)m1+(xn+3x2)m20,m1m20,2^{\circ}.\left(x_{n+2}-x_{1}\right)\mu_{1}+\left(x_{n+3}-x_{2}\right)\mu_{2}\neq 0,\quad\mu_{1}\mu_{2}\geqq 0,\quadto be verified.
From the condition22^{\circ}it also follows that the nodes are not all confused. Moreover, if the firstn+2n+2respectively the lastn+2n+2nodes are confused, the coefficientm2\mu_{2}respectively the coefficientm1\mu_{1}is0\neq 0.

To prove the theorem it is necessary and sufficient to verify that in the cases11^{\circ}and22^{\circ}of the statement, the functional is of simple form, while in the other possible cases it is not of simple form. These possible cases are the following:
33^{\circ}. The nodesxix_{i}they are all confused.
44^{\circ}. The nodesxix_{i}are not all confused andm1m20\mu_{1}\mu_{2}\geqq 0,

(xn+2x1)m1+(xn+3x2)m2=0.\left(x_{n+2}-x_{1}\right)\mu_{1}+\left(x_{n+3}-x_{2}\right)\mu_{2}=0.

55^{\circ}. The nodesninnuthey are all confused andm1m2<0,m1+m20\mu_{1}\mu_{2}<0,\mu_{1}+\mu_{2}\neq 0
We will examine each of the 5 cases .
11^{\circ}In this case expression (60) can be written asm2(xn+3x1)[x1,x2,,xn+3;f]\mu_{2}\left(x_{n+3}-x_{1}\right)\left[x_{1},x_{2},\ldots,x_{n+3};f\right]It is of the degree of accuracyn+1n+1and is of simple form, by virtue of Theorem 8.
22^{\circ}The property follows from Theorem 11.
33^{\circ}Based on the definition of the differences divided by not all distinct nodes, expression (60) is of the form(m1+m2)[x1,x1,,x1n2;f]\left(\mu_{1}+\mu_{2}\right)[\underbrace{x_{1},x_{1},\ldots,x_{1}}_{n-2};f]Then the linear functional is:3the3^{\prime o}. or identically null, so it is not of the simple form,3′′3^{\prime\prime}. Or has the degree of accuracynn, but it cancels out on the function|xx1|(xx1)n+1\left|x-x_{1}\right|\left(x-x_{1}\right)^{n+1}which is convex of ordernn, so it is not of simple form.
44^{\circ}At least one of the coefficientsm1,m2\mu_{1},\mu_{2}is zero and the linear functional (60) is:44^{\prime\circ}. or identically null,4′′′4^{\prime\prime\prime}or in the form of33^{\circ}previous. In this case too, the functional is not of simple form.
55^{\circ}The degree of accuracy isnnand we can write withWith1<With2<<Withpz_{1}<z_{2}<\ldots<z_{p}the distinct nodes,klk_{l}being the order of multiplicity ofWithiz_{i}We have1k1,k2,,kpn+2,k1+k2++kp=n+31\leqq k_{1},k_{2},\ldots,k_{p}\leqq n+2,k_{1}+k_{2}+\ldots+k_{p}=n+3Let's consider the functions

ψ1=(xl1|xl1|2)n+2ψ2=(xl2+|xl2|2)n+2\psi_{1}=-\left(\frac{x-\lambda_{1}-\left|x-\lambda_{1}\right|}{2}\right)^{n+2}\quad\psi_{2}=\left(\frac{x-\lambda_{2}+\left|x-\lambda_{2}\right|}{2}\right)^{n+2} (61)

which are non-concave of the ordernnand belong to the definition set𝔽\mathbb{F}of the linear functional (60), as this set was defined in no.𝟐𝟑\mathbf{23}Indeed, the functions (61) have (everywhere) continuous derivatives of ordern+1n+1We will calculate onR[ψ1]R\left[\psi_{1}\right]and onR[ψ2]R\left[\psi_{2}\right], assuming thatl1(With1,With2)\lambda_{1}\in\left(z_{1},z_{2}\right)andl2(Withp1,Withp)\lambda_{2}\in\left(z_{p-1},z_{p}\right)It is unnecessary to reproduce this calculation in detail here. We have

R[ψ1]=i=1k11Mi(l1With1)n+2i,(With1<l1<With2)\displaystyle R\left[\psi_{1}\right]=\sum_{i=1}^{k_{1}-1}M_{i}\left(\lambda_{1}-z_{1}\right)^{n+2-i},\quad\left(z_{1}<\lambda_{1}<z_{2}\right)
R[ψ2]=i=1kp1Ni(Withpl2)n+2i,(Withp1<l2<Withp),\displaystyle R\left[\psi_{2}\right]=\sum_{i=1}^{k_{p}-1}N_{i}\left(z_{p}-\lambda_{2}\right)^{n+2-i},\quad\left(z_{p-1}<\lambda_{2}<z_{p}\right),

where

Mk11=(n+2k11)m1[i=2p1(WithiWith1)ki](WithpWith1)kp1,Nkp1=(n+2kp1)m2[l=2p1(WithpWithi)ki](WithpWith1)k11,M_{k_{1}-1}=\frac{\binom{n+2}{k_{1}-1}\mu_{1}}{\left[\prod_{i=2}^{p-1}\left(z_{i}-z_{1}\right)^{k_{i}}\right]\left(z_{p}-z_{1}\right)^{k_{p-1}}},\quad N_{k_{p}-1}=\frac{\binom{n+2}{k_{p}-1}\mu_{2}}{\left[\prod_{l=2}^{p-1}\left(z_{p}-z_{i}\right)^{k_{i}}\right]\left(z_{p}-z_{1}\right)^{k_{1}-1}},

the other coefficientsMi,NiM_{i},N_{i}, independent ofl1\lambda_{1}andl2\lambda_{2}, having values ​​that are unnecessary to calculate here.

We note thatMk11,Nkp1M_{k_{1}-1},N_{k_{p}-1}are different from zero and of the same sign asm1,m2\mu_{1},\mu_{2}respectively. It is then seen that we can find al1\lambda_{1}close enough toWith1z_{1}and al2\lambda_{2}close enough toWithpz_{p}so that we haveR[ψ1].R[ψ2]<0R\left[\psi_{1}\right].R\left[\psi_{2}\right]<0. From an observation made in no. 10 it follows that the linear functional (60) cannot be of simple form.

Theorem 12 is completely proven.

The construction of functions (61) depends, to some extent, on the space (7. If this space is more restricted, e.g. if it contains only indefinitely differentiable functions onANDE, we must replace the functions (61) by other convenient ones. We can avoid this modification by criteria analogous to those studied below (see no. 30).
25. Remaining in the particular case (21), (21'), ifR[f]R[f]is a linear functional defined on(,R[f]=R[f]\left(\mathcal{F},R^{*}[f]=R\left[f^{\prime}\right]\right.is a linear functional defined on the set\mathscr{F}^{*}of continuous and differentiable functions whose derivative belongs to𝔽\mathbb{F}It is easy to see that ifR[f]R[f]is of degree of accuracyn(1),R[f]n(\geqq-1),R^{*}[f]is of degree of accuracyn+1n+1.

We also have
Theorem 13. - Under the previous assumptions, becauseR[f]R[f]to be of simple form, it is necessary and sufficient thatR[f]R^{*}[f]to be of simple form.

The proof is immediate. It suffices to observe that the derivative of a convex function of ordernnis a convex function of ordern1n-1and that all the primitives of such a function are convex functions of ordern+1n+1.
26. To make an application, let's consider the numerical quadrature formula

abf(x)𝑑x=i=0k1aif(i)(a)+i=0l1bif(i)(c)+i=0m1cif(i)(b)+R[f]\int_{a}^{b}f(x)dx=\sum_{i=0}^{k-1}\alpha_{i}f^{(i)}(a)+\sum_{i=0}^{l-1}\beta_{i}f^{(i)}(c)+\sum_{i=0}^{m-1}\gamma_{i}f^{(i)}(b)+R[f] (62)

whereffis a continuous function on[a,b][a,b]having the derivatives written continuous anda<c<ba<c<b.

Let us assume that the remainder of formula (62) is zero on any polynomial of degreen1=k+l+m10n-1=k+l+m-1\geq 0. Then the formula falls into the category of those studied at no. 20. The numbersk,l,mk,l,mcan be zero, which means that the corresponding sum (hence the pointa,ca,corbbcorresponding) does not occur in the second term of formula (62).

Particular cases of formula (62) have been studied by various authors and in particular by K. Petr [10, 11], GN Watson [27], N. Obreschkoff [9]. The method of these authors is different from the one presented here.

By virtue of Theorem 10, the remainder is of simple form ifllis even, in particular so ifl=0l=0We will find this result below with the help of Theorems 12 and 13.

It is easily seen thatR[f]R[f]has a finite degree of accuracy which is equal ton1n-1or withnn. Linear functionalR[f]=R[f]R^{*}[f]=R\left[f^{\prime}\right]is of the form (58), with nodes not all confused, their total number beingn+2n+2ifl=0l=0andn+3n+3ifl>0l>0We can now discuss the simplicity of the remainder with the help of Theorems 12 and 13.
R[f]R[f]is of degree of accuracynnif and only if

P(c)=ab(xa)k(xc)l(bx)m𝑑x=0P(c)=\int_{a}^{b}(x-a)^{k}(x-c)^{l}(b-x)^{m}dx=0 (63)

This algebraic equation (of degreell) incchas no real root in (a,ba,b) (in fact on the entire real axis) ifllis even, and has only one real rootcc^{*}which is in(a,b)(a,b)ifllis odd. This result is obtained by noting that the derivative equationP(c)=0P^{\prime}(c)=0is of the same shape.R[f]R[f]is therefore of degree of accuracynnif and only ifllis odd andc=cc=c^{*}.

Theorem 11 shows us that ifl=0,R[t]l=0,R^{*}[t]is of degree of accuracynnand is of simple form. ThereforeR[f]R[f]is of degree of accuracyn1n-1and of the simple form. It is also seen that ifllis odd andc=cc=c^{*}, it is of accuracy gradennand it is of simple form.

To study the other possible cases, the coefficients must be calculatedm1,m2\mu_{1},\mu_{2}of the formula (60) corresponding toR[f]R^{*}[f]Some calculations, which we will not reproduce in detail, give us
m1=(1)l+mk!(ca)l+1(bc)mak1,m2=m!(ba)k(bc)l+1cm1\mu_{1}=(-1)^{l+m}k!(c-a)^{l+1}(b-c)^{m}\alpha_{k-1},\mu_{2}=-m!(b-a)^{k}(b-c)^{l+1}\gamma_{m-1}, where

ak1=(1)l(k1)!(ca)l(ba)mab(xa)k1(xc)l(bx)m𝑑x,(k>0)cm1=(1)m1(m1)!(ba)k(bc)lab(xa)k(xc)l(bx)m1𝑑x,(m>0)a1=1,c1=1,(0!=1)\begin{gathered}\alpha_{k-1}=\frac{(-1)^{l}}{(k-1)!(c-a)^{l}(b-a)^{m}}\int_{a}^{b}(x-a)^{k-1}(x-c)^{l}(b-x)^{m}dx,\quad(k>0)\\ \gamma_{m-1}=\frac{(-1)^{m-1}}{(m-1)!(b-a)^{k}(b-c)^{l}}\int_{a}^{b}(x-a)^{k}(x-c)^{l}(b-x)^{m-1}dx,\quad(m>0)\\ \alpha_{-1}=1,\gamma_{-1}=-1,\quad(0!=1)\end{gathered}

Applying Theorem 12, we see that ifl>0l>0and if the restR[f]R[f]is of degree of accuracyn1n-1, it is of simple form if and only ifm1m2>0\mu_{1}\mu_{2}>0This condition is checked ifllis an even number.

Ifllis odd andk>0k>0, exists in(a,b)(a,b)a valuec1c_{1}his/herccand only one for whichm1=0\mu_{1}=0and ifm>0m>0a valuec2c_{2}his/herccand only one for whichm2=0\mu_{2}=0.

wec1<c<c2c_{1}<c^{*}<c_{2}To prove the first inequality it is sufficient to observe that for the polynomial (63) we have

P(a)>0,P(c1)=ab(xa)k1(xc)l+1(bx)m𝑑x>0P(a)>0,\quad P\left(c_{1}\right)=\int_{a}^{b}(x-a)^{k-1}\left(x-c^{*}\right)^{l+1}(b-x)^{m}dx>0

The second inequality is proved in the same way.
It is immediately seen that ifc1<c<c2c_{1}<c<c_{2}HAVEm1m2<0\mu_{1}\mu_{2}<0, and ifcc1c\leqq c_{1}orc2cc_{2}\leqq cHAVEm1m20\mu_{1}\mu_{2}\geqq 0The results remain even ifk=0k=0TAKINGc1=ac_{1}=a, and ifm=0m=0taking thenc2=bc_{2}=b.

RESTR[f]R[f]of formula (62) is therefore of simple form only in the following three cases:
1.l1^{\circ}.lodd,c=cc=c^{*}.
22^{\circ}. llodd,a<cc1a<c\leqq c_{1}orc2c<bc_{2}\leqq c<b.
33^{\circ}.llabout.

In case11^{\circ}the rest has the form

R[f]=K[x1,x2,,xk+l+m+2;f]R[f]=K^{*}\left[\xi_{1},\xi_{2},\ldots,\xi_{k+l+m+2};f\right]

and in cases22^{\circ}and33^{\circ}is of the form

R[f]=K[x1,x2,,xk+l+m+1;f]R[f]=K\left[\xi_{1},\xi_{2},\ldots,\xi_{k+l+m+1};f\right]

wherexi\xi_{i}are, each time, distinct points of the interval (a,ba,b) and

K=ab(xa)k(xc)l+1(xb)m𝑑x,K=ab(xa)k(xc)l(xb)m𝑑xK^{*}=\int_{a}^{b}(x-a)^{k}(x-c)^{l+1}(x-b)^{m}dx,\quad K=\int_{a}^{b}(x-a)^{k}(x-c)^{l}(x-b)^{m}dx

In the "symmetric" casek=mk=m, we havec=12(a+b)c^{*}=\frac{1}{2}(a+b)andc1+c2=a+bc_{1}+c_{2}=a+b
In the case of l=1l=1, we have

c1=(m+1)a+kbm+k+1,c=(m+1)a+(k+1)bm+k+2,c2=ma+(k+1)bm+k+1c_{1}=\frac{(m+1)a+kb}{m+k+1},c^{*}=\frac{(m+1)a+(k+1)b}{m+k+2},c_{2}=\frac{ma+(k+1)b}{m+k+1}

It can be shown that in all cases of simplicity of remainder, simplicity also occurs if the function is assumed to be continuous only on[a,b][a,b], having on the pointsa,c,ba,c,bthe derivatives that actually appear in the second member of formula (62). The hypothesis of continuity of the derivative of the ordermax(k1,l1,m1)\max(k-1,l-1,m-1)was imposed only by the definition we adopted for differences divided by multiple nadas and by the criterion we relied on to demonstrate the simplicity of the remainder.
27. In the particular case (21), (21'), we will resume, specifying and completing it, a criterion we have already given [15].

Whether

fn+1,l=(xl+|xl|2)n\varphi_{n+1,\lambda}=\left(\frac{x-\lambda+|x-\lambda|}{2}\right)^{n} (64)

wherennis a natural number. This is a non-concave function of ordernnfor anythingxxIts derivative of the orderkkexists if0kn10\leqq k\leqq n-1and is continuous for anyxxWe have, moreover,

fn+1,l(k)=n!(nk)!fn+1k,l(0kn1).\varphi_{n+1,\lambda}^{(k)}=\frac{n!}{(n-k)!}\varphi_{n+1-k,\lambda}\quad(0\leqq k\leqq n-1). (65)

Whethernna natural number and divide the finite and closed interval[a,b][a,b]inm>2nm>2nequal parts by points

li=a+ih,i=0,1,,m,h=bam(l0=a,lm=b)\lambda_{i}=a+ih,i=0,1,\ldots,m,h=\frac{b-a}{m}\quad\left(\lambda_{0}=a,\lambda_{m}=b\right) (66)

We denote with
Dji[f]=[li,li+1,,li+j;f],i=0,1,,mj,j=0,1,,mD_{j}^{i}[f]=\left[\lambda_{i},\lambda_{i+1},\ldots,\lambda_{i+j};f\right],\quad i=0,1,\ldots,m-j,j=0,1,\ldots,m(67) the divided (ordinary) differences of the function//on (66) consecutive points,

Let's consider the functions

ψm=fm+Qm\psi_{m}=f_{m}+Q_{m} (68)

where

fm=(n+1)hi=0mn1Dn+1i[f]fn+1,li+n\displaystyle f_{m}=(n+1)h\sum_{i=0}^{m-n-1}D_{n+1}^{i}[f]\varphi_{n+1,\lambda_{i+n}} (69)
Qm=(1)nn!hn{r=0n(1)rf(lr)[i=0r(1)i(n+1ri)(xli+n)n]}\displaystyle Q_{m}=\frac{(-1)^{n}}{n!h^{n}}\left\{\sum_{r=0}^{n}(-1)^{r}f\left(\lambda_{r}\right)\left[\sum_{i=0}^{r}(-1)^{i}\binom{n+1}{r-i}\left(x-\lambda_{i+n}\right)^{n}\right]\right\} (70)

The function (68) is continuous and has a continuous derivative of ordern1n-1(so in any ordern1\leqq n-1) for anyxxIt reduces to a polynomial of degreennin each of the intervals[li,li+1],i=0,1,,m1\left[\lambda_{i},\lambda_{i+1}\right],i=0,1,\ldots,m-1. Pm\Psi_{m}is what I once called an elementary function of ordernn.

We have shown [15] that ifffis continuous on{a,b}\{a,b\}, the string{ψm}\left\{\psi_{m}\right\}converges uniformly over the entire interval[a,b][a,b]byffformcm\rightarrow{}^{c}We will complete this convergence property for the case when the functionffis differentiable a certain number of times.
28. Before stating and proving Theorem 14, which we will establish below, it is necessary to make some preliminary calculations.

Recurrence formulaDji[f]=1jh{Dj1i+1[f]Dj1i[f])}\left.D_{j}^{i}[f]=\frac{1}{jh}\left\{D_{j-1}^{i+1}[f]-D_{j-1}^{i}[f]\right)\right\}allows us to establish various relations between the divided differences (67). Thus we have

(n+1)hDn+1i[f]=(1)n+1kk!n!hnkj=0n+1k(1)j(n+1kj)Dki+j|f|(n+1)hD_{n+1}^{i}[f]=\frac{(-1)^{n+1-k}k!}{n!h^{n-k}}\sum_{j=0}^{n+1-k}(-1)^{j}\binom{n+1-k}{j}D_{k}^{i+j}|f| (71)

Herekkis a whole such that0kn+10\leq k\leq n+1For what follows it will suffice to assume that0kn10\leqq k\leqq n-1.

Taking into account formula (71), function (69) becomes

im=(1)n+1kk!n!hnkr=0mkDkr[f][(1)ri=rn1+kr(1)i(n+1kri)fn+1,li+n]i_{m}=\frac{(-1)^{n+1-k}k!}{n!h^{n-k}}\sum_{r=0}^{m-k}D_{k}^{r}[f]\left[(-1)^{r}\sum_{i=r-n-1+k}^{r}(-1)^{i}\binom{n+1-k}{r-i}\varphi_{n+1,\lambda_{i+n}}\right] (72)

wherefn+1,li+n=0\varphi_{n+1,\lambda_{i+n}}=0fori<0i<0and fori>ji>jifx[lj+n,lj+n+1]x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right], j=n,n+1,,mn1j=-n,-n+1,\ldots,m-n-1.

To simplify, we introduce the notations

Pi,in,in=(1)ri=inin(1)i(n+1kri)(xli+n)nP_{i,u,v}=(-1)^{r}\sum_{i=u}^{u}(-1)^{i}\binom{n+1-k}{r-i}\left(x-\lambda_{i+n}\right)^{n} (73)

Taking into account (73), we find

fm=0, for x[l0,ln]\displaystyle f_{m}=0,\text{ pentru }x\in\left[\lambda_{0},\lambda_{n}\right]
fm=(1)n+1kk!n!hnk[r=0nkDkr[f]Pr,0,rr=f+1nkDkr[f]Pr,j+1,r+\displaystyle f_{m}=\frac{(-1)^{n+1-k}k!}{n!h^{n-k}}\left[\sum_{r=0}^{n-k}D_{k}^{r}[f]P_{r,0,r}-\sum_{r=f+1}^{n-k}D_{k}^{r}[f]P_{r,j+1,r}+\right.

+r=n+1kn+1k+jDkr[j]Pr,rn1+k,j]\left.+\sum_{r=n+1-k}^{n+1-k+j}D_{k}^{r}[j]P_{r,r-n-1+k,j}\right], forx[lj+n,lj+n+1]x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right],

j=0,1,,nk1,\displaystyle j=1,\ldots,n-k-1,
fm=(1)n+1kk!n!hnk[r=0nkDkr[f]Pk,0,r+r=n+1k2n2k+1Dkr[f]Pr,rn1+k,nk],\displaystyle f_{m}=\frac{(-1)^{n+1-k}k!}{n!h^{n-k}}\left[\sum_{r=0}^{n-k}D_{k}^{r}[f]P_{k,0,r}+\sum_{r=n+1-k}^{2n-2k+1}D_{k}^{r}[f]P_{r,r-n-1+k,n-k}\right],
for x[l2nk,l2nk+1],\displaystyle\text{ pentru }x\in\left[\lambda_{2n-k},\lambda_{2n-k+1}\right],
fm=(1)n+1kk!n!hnk[r=0nkDkr[f]Pr,0,r+r=n+1kjDkr[f]Pr,rn1+k,r+\displaystyle f_{m}=\frac{(-1)^{n+1-k}k!}{n!h^{n-k}}\left[\sum_{r=0}^{n-k}D_{k}^{r}[f]P_{r,0,r}+\sum_{r=n+1-k}^{j}D_{k}^{r}[f]P_{r,r-n-1+k,r+}\right.
+k=j+1n+1k+jDkr[f]Pr,rn1+k],forx[lj+n,lj+n+1],\displaystyle\left.+\sum_{k=j+1}^{n+1-k+j}D_{k}^{r}[f]P_{r,r-n-1+k}\right],\operatorname{pentru}x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right],
j=nk+1,nk+2,,mn1.\displaystyle j=n-k+1,n-k+2,\ldots,m-n-1.

To put the polynomial (70) in a convenient form, we will apply the transformation formula

r=0ncrf(lr)=(1)kk!hkr=0nkDkr[f][s=0r(k+s1s)crs]+\displaystyle\quad\sum_{r=0}^{n}c_{r}f\left(\lambda_{r}\right)=(-1)^{k}k!h^{k}\sum_{r=0}^{n-k}D_{k}^{r}[f]\left[\sum_{s=0}^{r}\binom{k+s-1}{s}c_{r-s}\right]+
+(1)nr=n+1kn(1)r(nr)!hnrrDnrr[f][s=0r(nr+ss)crs].\displaystyle+(-1)^{n}\sum_{r=n+1-k}^{n}(-1)^{r}(n-r)!h^{n-r-r}D_{n-r}^{r}[f]\left[\sum_{s=0}^{r}\binom{n-r+s}{s}c_{r-s}\right].
Let's take
cr=(1)ri=0r(1)i(n+1ri)(xli+n)n,r=0,1,,nc_{r}=(-1)^{r}\sum_{i=0}^{r}(-1)^{i}\binom{n+1}{r-i}\left(x-\lambda_{i+n}\right)^{n},r=0,1,\ldots,n

If we take into account the well-known formula (see, e.g., E. Netto [8])
we deduce

s=0l(1)s(s+as)(bts)=(ba1t)\sum_{s=0}^{l}(-1)^{s}\binom{s+a}{s}\binom{b}{t-s}=\binom{b-a-1}{t}
s=0r(k+s1s)crs=(1)ri=0r(1)i[s=0ri(1)s(k+s1s)(n+1ris)](xli+n)n==Pr,0,rs=0r(nr+ss)crs=(1)ri=0r(1)i(rri)(xli+n)n\begin{gathered}\sum_{s=0}^{r}\binom{k+s-1}{s}c_{r-s}=(-1)^{r}\sum_{i=0}^{r}(-1)^{i}\left[\sum_{s=0}^{r-i}(-1)^{s}\binom{k+s-1}{s}\binom{n+1}{r-i-s}\right]\left(x-\lambda_{i+n}\right)^{n}=\\ =P_{r,0,r}\\ \sum_{s=0}^{r}\binom{n-r+s}{s}c_{r-s}=(-1)^{r}\sum_{i=0}^{r}(-1)^{i}\binom{r}{r-i}\left(x-\lambda_{i+n}\right)^{n}\end{gathered}

where, finally,

Qm=(1)nkk!n!hnkr=0nkDkr(f]Pr,0,r+\displaystyle Q_{m}=\frac{(-1)^{n-k}k!}{n!h^{n-k}}\sum_{r=0}^{n-k}D_{k}^{r}(f]P_{r,0,r}+
+r=n+1kn{(nr)!n!hrDnrr[f][i=0r(1)i(rri)(xli+n)n]}.\displaystyle+\sum_{r=n+1-k}^{n}\left\{\frac{(n-r)!}{n!h^{r}}D_{n-r}^{r}[f]\left[\sum_{i=0}^{r}(-1)^{i}\binom{r}{r-i}\left(x-\lambda_{i+n}\right)^{n}\right]\right\}. (74)

We will now calculate the derivatives of the function (68). Note that

i=0r(1)i(rri)(xli+n)nk=(1)rr!hrDrn[(xt)nk]\sum_{i=0}^{r}(-1)^{i}\binom{r}{r-i}\left(x-\lambda_{i+n}\right)^{n-k}=(-1)^{r}r!h^{r}D_{r}^{n}\left[(x-t)^{n-k}\right]

where we considerxxas a parameter andttas the variable of the polynomial(xt)nk(x-t)^{n-k}whose divided difference is calculated on the nodesln,ln+1,,ln+r\lambda_{n},\lambda_{n+1},\ldots,\lambda_{n+r}. However, the difference divided by the orderrrof a polynomial of degreer1r-1is identically zero. It follows that the derivative of the orderkkof the second sum of the second term of formula (74), vanishes. It is seen in the same way thatPr,rn1+k,r(k)=0P_{r,r-n-1+k,r}^{(k)}=0forrn+1kr\geqq n+1-k.

So we have

ψm(k)=\displaystyle\psi_{m}^{(k)}= (1)nkk!n!nnkr=0nkDkr[f]Pr,0,r(k) for x[l0,ln]\displaystyle\frac{(-1)^{n-k}k!}{n!n^{n-k}}\sum_{r=0}^{n-k}D_{k}^{r}[f]P_{r,0,r}^{(k)}\quad\text{ pentru }x\in\left[\lambda_{0},\lambda_{n}\right]
ψm(k)=\displaystyle\psi_{m}^{(k)}= (1)n+1kk!n!hnk[r=j+1nkDhr[f]Pr,j+1,r(k)+r=n+1kn+1k+iDkr[/]Pr,in1+k,j(k)]\displaystyle\frac{(-1)^{n+1-k}k!}{n!h^{n-k}}\left[-\sum_{r=j+1}^{n-k}D_{h}^{r}[f]P_{r,j+1,r}^{(k)}+\sum_{r=n+1-k}^{n+1-k+i}D_{k}^{r}[/]P_{r,i-n-1+k,j}^{(k)}\right]
for x[lj+n,lj+n+1],j=0,1,,nk1\displaystyle\text{ pentru }x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right],j=1,\ldots,n-k-1
ψm(k)=\displaystyle\psi_{m}^{(k)}= (1)n+kkk!n!hnkr=j+1nk+jDkr[f]Pr,rn1+k,j(k)\displaystyle\frac{(-1)^{n+k-k}k!}{n!h^{n-k}}\sum_{r=j+1}^{n-k+j}D_{k}^{r}[f]P_{r,r-n-1+k,j}^{(k)}
for x[lj+n,lj+n+1],j=nk,nk+1,,mn1\displaystyle\text{ pentru }x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right],j=n-k,n-k+1,\ldots,m-n-1

We will also need some convenient delimitations of derivatives of the orderkkof the polynomials (73) that intervene in these formulas.

For0srnk,x[l0,l2nk]0\leqq s\leqq r\leqq n-k,x\in\left[\lambda_{0},\lambda_{2n-k}\right]HAVE

|Pr,s,r(k)|n!(nk)!i=0r(n+1kri)|xli+n|nk==n!(nk)!i=0r(n+1ki)|xlr+ni|nkn!(nk)!i=0r(n+1ki)[maxx[l0,l2nk]|xlr+ni|nk]n!hnk(nk)!(2nk)nk(2n+1k1)\begin{gathered}\left|P_{r,s,r}^{(k)}\right|\leqq\frac{n!}{(n-k)!}\sum_{i=0}^{r}\binom{n+1-k}{r-i}\left|x-\lambda_{i+n}\right|^{n-k}=\\ =\frac{n!}{(n-k)!}\sum_{i=0}^{r}\binom{n+1-k}{i}\left|x-\lambda_{r+n-i}\right|^{n-k}\leqq\\ \leqq\frac{n!}{(n-k)!}\sum_{i=0}^{r}\binom{n+1-k}{i}\left[\max_{x\in\left[\lambda_{0},\lambda_{2n-k}\right]}\left|x-\lambda_{r+n-i}\right|^{n-k}\right]\leqq\\ \leqq\frac{n!h^{n-k}}{(n-k)!}(2n-k)^{n-k}\left(2^{n+1-k}-1\right)\end{gathered}

So if we put

M=k!(nk)!(2nk)nk(2n+1k1)M=\frac{k!}{(n-k)!}(2n-k)^{n-k}\left(2^{n+1-k}-1\right) (75)

we have, in particular,

|Pr,0,r(k)|n!hnkk!M, for 0rnk,x[l0,ln]|Pr,j+1,r(k)n!hnkk!M, for j+1rnk,x[lj+11,lj+n+1]j=0,1,,nk1\begin{gathered}\left|P_{r,0,r}^{(k)}\right|\leqq\frac{n!h^{n-k}}{k!}M,\quad\text{ pentru }0\leqq r\leqq n-k,\quad x\in\left[\lambda_{0},\lambda_{n}\right]\\ \left\lvert\,P_{r,j+1,r}^{(k)}\leqq\frac{n!h^{n-k}}{k!}M\right.,\quad\text{ pentru }j+1\leqq r\leqq n-k,\quad x\in\left[\lambda_{j+11},\lambda_{j+n+1}\right]\\ \quad j=0,1,\ldots,n-k-1\end{gathered}

Forj+1rn+1k+j,x[lj+n,lj+n+1],j=0,1,;nn1j+1\leqq r\leqq n+1-k+j,x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right],j=0,1,\ldots;n-n-1,
we have

|Pr,rn1+k,j(k)|n!(nk)!i=rn1+k(n+1kri)|xli+n|nk\displaystyle\left|P_{r,r-n-1+k,j}^{(k)}\right|\leqq\frac{n!}{(n-k)!i=r-n-1+k}\binom{n+1-k}{r-i}\left|x-\lambda_{i+n}\right|^{n-k}\leqq
n!(nk)!i=0n+1k+jr(n+1ki)x[lj+nlj+n+1]|xlr+k1+i|nk\displaystyle\leqq\frac{n!}{(n-k)!}\sum_{i=0}^{n+1-k+j-r}\binom{n+1-k}{i}_{x\in\left[\lambda_{j+n^{\prime}}\lambda_{j+n+1}\right]}\left|x-\lambda_{r+k-1+i}\right|^{n-k}\leqq
n!hnkn+1k+jr(nk)!i=0(n+1ki)(n+2k+jri)nk\displaystyle\leqq\frac{n!h^{n-k}n+1-k+j-r}{(n-k)!}\sum_{i=0}\binom{n+1-k}{i}(n+2-k+j-r-i)^{n-k}\leqq
n!hnk(nk)!(n+2k+jr)nki=0n+1k+ir(n+1ki)<\displaystyle\leqq\frac{n!h^{n-k}}{(n-k)!}\left(n+2-k+j-r^{\prime}\right)^{n-k}\sum_{i=0}^{n+1-k+i-r}\binom{n+1-k}{i}<
<n!hnk(nk)!(n+1k)nk(2n+1k1)n!hnkk!M.\displaystyle<\frac{n!h^{n-k}}{(n-k)!}(n+1-k)^{n-k}\left(2^{n+1-k}-1\right)\leqq\frac{n!h^{n-k}}{k!}M.

Better delimitations can be found. I have given delimitations of this kind in another paper [15]. For what follows it is sufficient to note that the number (75) is independent ofmm(and ofjj).
29. I will demonstrate the act.

THEOREM 14. - Given a natural numbernnand the wholekkso that0kn10\leqq k\leqq n-1, if the function\midadmitstheoderived from the orderkkcontinue at the intersectionIIof the tinnitus and closed interval[a,b][a,b]with an open interval,
the series of derivatives of orderk{ψm(k)}k\left\{\psi_{m}^{(k)}\right\}, of the functions (68) converges uniformly to the derivative of orderk,f(l)k,f^{(l)}, of the function ! formm\rightarrow\inftyand on any closed subinterval ofII.

The zeroth derivative of a function coincides with the function itself.
The conclusion of the statement means that the convergence is uniform on[a,b][a,b]\left[a^{\prime},b^{\prime}\right]\subseteq[a,b]iff(k)f^{(k)}is continuous on[a,b]\left[a^{\prime},b^{\prime}\right]and if, in addition,a<aa<a^{\prime}, is continuous and on an interval[a′′,a)\left[a^{\prime\prime},a^{\prime}\right)witha<a′′<aa<a^{\prime\prime}<a^{\prime}, and ifb<bb^{\prime}<bis continuous and on an interval (b,b′′b^{\prime},b^{\prime\prime}] withb<b′′<bb^{\prime}<b^{\prime\prime}<b.

To demonstrate, we will delimit the differencef(k)ψm(k)f^{(k)}-\psi_{m}^{(k)}.

If in his expressionψm(k)\psi_{m}^{(k)}we replace all divided differencesDkr[f]D_{k}^{r}[f]by 1, the functionfm(k)f_{m}^{(k)}the polynomial also cancels out identicallyQm(k)Q_{m}^{(k)}it reduces to

(1)nkk!n!hnkr=0nkPr,0,r(k)=\displaystyle\frac{(-1)^{n-k}k!}{n!h^{n-k}}\sum_{r=0}^{n-k}P_{r,0,r}^{(k)}=
=\displaystyle= (1)nkk!(nk)!hnkr=0nk(1)r[i=0r(1)i(n+1kri)(xli+n)nk]=\displaystyle\frac{(-1)^{n-k}k!}{(n-k)!h^{n-k}}\sum_{r=0}^{n-k}(-1)^{r}\left[\sum_{i=0}^{r}(1-)^{i}\binom{n+1-k}{r-i}\left(x-\lambda_{i+n}\right)^{n-k}\right]=
=\displaystyle= (1)nkk!(nk)!hnki=0nk(1)i[i=0nk(1)r(n+1kri)](xli+n)nk=\displaystyle\frac{(-1)^{n-k}k!}{(n-k)!h^{n-k}}\sum_{i=0}^{n-k}(-1)^{i}\left[\sum_{i=0}^{n-k}(-1)^{r}\binom{n+1-k}{r-i}\right]\left(x-\lambda_{i+n}\right)^{n-k}=
=\displaystyle= k!(nk)!hnki=1nk(1)i(nki)(xli+n)nk=\displaystyle\frac{k!}{(n-k)!h^{n-k}}\sum_{i=1}^{n-k}(-1)^{i}\binom{n-k}{i}\left(x-\lambda_{i+n}\right)^{n-k}=
=\displaystyle= k!(nk)!i=0nk(1)i(nki)(nki)nk=k!\displaystyle\frac{k!}{(n-k)!}\sum_{i=0}^{n-k}(-1)^{i}\binom{n-k}{i}(n-k-i)^{n-k}=k!

In this calculation we have taken into account an observation already made on the divided differences of a polynomial. It is seen that the expression is independent ofxxand so one can take (e.g.)x=l2nkx=\lambda_{2n-k}.

It follows that the differencef(k)ψm(k)f^{(k)}-\psi_{m}^{(k)}is obtained fromψm(k)\psi_{m}^{(k)}replacingDkr[f]byf(k)k!Dkr[f],r=0,1,,mkD_{k}^{r}[f]\operatorname{prin}\frac{f^{(k)}}{k!}-D_{k}^{r}[f],\quad r=0,1,\ldots,m-k.

Taking into account the calculations made in the previous section, we have

|f(k)ψm(k)|Mr=0nk|f(k)k!Dkr[j]|, for x[l0,ln]|f(k)ψm(k)|Mr=j+1n+1k+j|f(k)k!Dkr[j]|, for x[land+n,lj+n+1]j=0,1,,mn1\begin{gathered}\left|f^{(k)}-\psi_{m}^{(k)}\right|\leq M\sum_{r=0}^{n-k}\left|\frac{f^{(k)}}{k!}-D_{k}^{r}[j]\right|,\text{ pentru }x\in\left[\lambda_{0},\lambda_{n}\right]\\ \left|f^{(k)}-\psi_{m}^{(k)}\right|\leq M\sum_{r=j+1}^{n+1-k+j}\left|\frac{f^{(k)}}{k!}-D_{k}^{r}[j]\right|,\text{ pentru }x\in\left[\lambda_{y+n},\lambda_{j+n+1}\right]\\ j=0,1,\ldots,m-n-1\end{gathered}

Be it now[a,b]\left[a^{\prime},b^{\prime}\right]a closed subinterval ofIILet us first assume thata<a<b<ba<a^{\prime}<b^{\prime}<band thena<a′′<a,b<b′′<ba<a^{\prime\prime}<a^{\prime},b^{\prime}<b^{\prime\prime}<b, derivative of the orderk,f(k)k,f^{(k)}being continuous on[a′′,b′′]\left[a^{\prime\prime},b^{\prime\prime}\right]Let us denote byohk(d)\omega_{k}(\delta)the oscillation modulus of1k!f(k)\frac{1}{k!}f^{(k)}on the interval[a′′,b′′]\left[a^{\prime\prime},b^{\prime\prime}\right].

Let's take the natural numbermmbig enough for us to have

m>max(2n,n(ba)aa′′,bab′′b,2(ba)b′′a)m>\max\left(2n,\frac{n(b-a)}{a^{\prime}-a^{\prime\prime}},-\frac{b-a}{b^{\prime\prime}-b^{\prime}},\frac{2(b-a)}{b^{\prime\prime}-a^{\prime}}\right) (76)

and let's put

j0=aahn,j1=b′′ahn1,h=bam,j_{0}=\left\lfloor\frac{a^{\prime}-a}{h}\right\rfloor-n,\quad j_{1}=\left\lfloor\frac{b^{\prime\prime}-a}{h}\right\rfloor-n-1,\quad h=\frac{b-a}{m},

wherea\lfloor\alpha\rfloormeans the largest integer less than or equal toa\alpha.

We then have0j0+1,j0j10\leq j_{0}+1,j_{0}\leqq j_{1}and[a,b][lj0+n,lj1+n+1][lj0+1,lj1+n+1][a′′¯,b′′]\left[a^{\prime},b^{\prime}\right]\subseteq\left[\lambda_{j_{0}+n},\lambda_{j_{1}+n+1}\right]\subseteq\subseteq\left[\lambda_{j_{0}+1},\lambda_{j_{1}+n+1}\right]\subseteq\left[\overline{a^{\prime\prime}},b^{\prime\prime}\right].

Ifj0jj1,j+1rn+1k+jj_{0}\leqq j\leqq j_{1},j+1\leqq r\leqq n+1-k+j, the nodes of the divided differenceDkr[f]D_{k}^{r}[f]I am in the range[lj+1,lj+n+1][a′′,b′′]\left[\lambda_{j+1},\lambda_{j+n+1}\right]\subseteq\left[a^{\prime\prime},b^{\prime\prime}\right], wheref(k)f^{(k)}is continuous. There is then a pointx\xiso that

Dkr[t]=1k!f(k)(x),x[lj+1,lj+n+1]D_{k}^{r}[t]=\frac{1}{k!}f^{(k)}(\xi),\quad\xi\in\left[\lambda_{j+1},\lambda_{j+n+1}\right]

and it turns out that

|f(k)k!Dkr[f]|ohk(nh), for x[lj+n,lj+n+1].\left|\frac{f^{(k)}}{k!}-D_{k}^{r}[f]\right|\leqq\omega_{k}(nh),\text{ pentru }x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right].

So we have

|f(k)ψinn(k)|(n+1k)Mohk(nh), for x[lj+n,lj+n+1]\left|f^{(k)}-\stackrel{{\scriptstyle i}}_{nn}^{(k)}\right|\leqq(n+1-k)M_{\omega_{k}}(nh),\text{ pentru }x\in\left[\lambda_{j+n},\lambda_{j+n+1}\right]
j=j0,j011,,j1j=j_{0},j_{0}-1-1,\ldots,j_{1}

so, all the more so,

|f(k)ψm(k)|(n+1k)Mohk(nh), for x[a,b]\left|f^{(k)}-\psi_{m}^{(k)}\right|\leqq(n+1-k)M\omega_{k}(nh),\text{ pentru }x\in\left[a^{\prime},b^{\prime}\right] (77)

which, based on the well-known properties of the oscillation modulus, of continuous functions, proves the theorem in this case.

It is easy to see that the delimitation (77) is also valid in the other possible cases. The modifications that need to be made to the proof are the following:

Ifb=bb^{\prime}=b, the term is deletedbab′′b\frac{b-a}{b^{\prime\prime}-b^{\prime}}, in the second member of formula (76).

Ifa=aa^{\prime}=a, the term is deletedn(ba)aa′′\frac{n(b-a)}{a^{\prime}-a^{\prime\prime}}in the second member of formula (76) and it is observed that forj<0j<0NUMBERrris subject to the condition that0rnk0\leqq r\leqq n-k. Divided difference nodesDkr[/]D_{k}^{r}[/]I am then in the range[l¯0,ln]\left[\bar{\lambda}_{0},\lambda_{n}\right].

Theorem 14 is proved.
30. We can now return to the study of the simplicity criteria of linear functionals.

Whether[a,b][a,b]a finite and closed interval and consider the non-ascending sequence ofk+1k+1partial intervals[a0,b0][a1,b1][ak,bk]\left[a_{0},b_{0}\right]\supseteqq\left[a_{1},b_{1}\right]\supseteq\ldots\supseteq\left[a_{k},b_{k}\right], wherea0=a,b0=ba_{0}=a,b_{0}=b.

Whetherk\bigodot_{k}function space//which admit continuous derivatives of the orderiion[ai,bi]\left[a_{i},b_{i}\right]fori=0,1,,ki=0,1,\ldots,kand let's consider the norm

f=i=0kmaxx[ai,bi]|f(i)|\|f\|=\sum_{i=0}^{k}\max_{x\in\left[a_{i},b_{i}\right]}\left|f^{(i)}\right| (78)

of this space.

we

THEOREM 15. - Find the given natural numbernnand the wholekkso that0hn10\leqq h\leqq n-1, if the linear functionalR[f]R[f]is:11^{\circ}. defined on@k@_{k},22^{\circ}. degree of accuracyn,3n,3^{\circ}. bounded with respect to the norm (78) becauseR[f]R[f]to be of simple form is necessary and sufficient for us to have

R[xn+1]R[fn+1,l]0, for l[a,b],R\left[x^{n+1}\right]R\left[\varphi_{n+1,\lambda}\right]\geqq 0,\text{ pentru }\lambda\in[a,b], (79)

where the functionsfn+1,n\varphi_{n+1,n}are defined by formula (64).
Let us note that the polynomials and functionsfn+1,l\varphi_{n+1,\lambda}, belong to the spaceandke_{k}.
The condition is necessary. Indeed,xn+1x^{n+1}is convex andfn+1,l\varphi_{n+1,\lambda}is non-concave of the ordernnThe property results from formula (31).

The condition is also sufficient. By hypothesis, we have

|R[f]|Af,t𝒞k,|R[f]|\leqq A\|f\|,\quad t\in\mathscr{C}_{k},

AAbeing a number independent of the functionffandf\|f\| norma (78).
Vom demonstra întîi că R[φn+1,λ]R\left[\varphi_{n+1,\lambda}\right] este o funcție continuă de λ\lambda pe [a,b][a,b]. Intr-adevăr, avem

φn+1,λφn+1,λ|n|λλ(ba)n1,x[a,b]\varphi_{n+1,\lambda}-\varphi_{n+1,\lambda^{\prime}}|\leqq n|\lambda-\lambda^{\prime}\mid(b-a)^{n-1},x\in[a,b]

deci și (0in1)(0\leqq i\leqq n-1),

|φn+1,λ(i)φn+1,λ(i)|=n!(ni)!|φn+1i,λφn+1i,λ|\displaystyle\left|\varphi_{n+1,\lambda}^{(i)}-\varphi_{n+1,\lambda^{\prime}}^{(i)}\right|=\frac{n!}{(n-i)!}\left|\varphi_{n+1-i,\lambda}-\varphi_{n+1-i,\lambda^{\prime}}\right|\leqq
n!(ni1)!|λλ|(ba)n1i,x[a,b].\displaystyle\leq\frac{n!}{(n-i-1)!}\left|\lambda-\lambda^{\prime}\right|(b-a)^{n-1-i},\quad x\in[a,b].

Avem deci
|R[φn+1,λ]R[φn+1,λ]|=|R[φn+1,λφn+1,λ]|Aφn+1,λφn+1,λ\left|R\left[\varphi_{n+1,\lambda}\right]-R\left[\varphi_{n+1,\lambda^{\prime}}\right]\right|=\left|R\left[\varphi_{n+1,\lambda}-\varphi_{n+1,\lambda^{\prime}}\right]\right|\leqq A\left\|\varphi_{n+1,\lambda}-\varphi_{n+1,\lambda^{\prime}}\right\|.
Dar,

φn+1,λφn+1,λ[i=0kn!(ni1)!(ba)n1i]|λλ|\left\|\varphi_{n+1,\lambda}-\varphi_{n+1,\lambda^{\prime}}\right\|\leqq\left[\sum_{i=0}^{k}\frac{n!}{(n-i-1)!}(b-a)^{n-1-i}\right]\left|\lambda-\lambda^{\prime}\right|

de unde proprietatea rezultă fără nici o dificultate.
Prin ipoteză, R[xn+1]0R\left[x^{n+1}\right]\neq 0, deci R[φn+1,λ]R\left[\varphi_{n+1,\lambda}\right] nut schimbă de semn cînd λ\lambda parcurge intervalul [a,b][a,b]. Să ne reamintim că o funcție convexă de ordinul nn pe [a,b][a,b] are o derivată continuă de toate ordinele n1\leq n-1 pe (a,b)(a,b). Dacă deci fekf\in e_{k} este convex de ordinul nn, în virtutea teoremei 14 şirul {R[ψm]}\left\{R\left[\psi_{m}\right]\right\} tinde către R[f]R[f] pentru mm\rightarrow\infty. Dar, pe baza formulelor (68) - (70), avem R[ψm]=R[f]=(n+1)hmn1Dn+1i[f]R[φn+1,λi+n]R\left[\psi_{m}\right]=R[f]=(n+1)h^{m-n-1}D_{n+1}^{i}[f]R\left[\varphi_{n+1,\lambda_{i+n}}\right] şi din (79) rezultă că dacă ff este convex de ordinul nn, avem

R[xn+1]R[f]0.R\left[x^{n+1}\right]R[f]\geqq 0. (80)

Rămîne să demonstrăm că în această formulă egalitatea nu poate avea loc. Am dat această demonstraţie în altă parte [15], aşa că mu mai revenim aici asupra ei.

Se deduce că pentru orice funcție fkf\in\mathbb{C}_{k} convexă de ordinul nn semmul >> este valabil în (80), deci că R[/]0R[/]\neq 0.

Yeorema 15 este deci demonstrată.
31. Fie R[/]R[/] o funcţională liniară definită pe eke_{k} şi mărginită in raport cu norma (78). Să presupunem că 0kn10\leqq k\leqq n-1 şi că a=a0==a1==ak,b=b0=b1==bka=a_{0}==a_{1}=\ldots=a_{k},b=b_{0}=b_{1}=\ldots=b_{k}. Atunci, după E. Ya. Remez [22], dacă R[f]R[f] este de gradul de exactitate nn, avem

R[f]=abf(μ)(x)𝑑αμ(x)R[f]=\int_{a}^{b}f^{(\mu)}(x)d\alpha_{\mu}(x) (81)

unde μ\mu este un întreg, kμn+1k\leqq\mu\leqq n+1 si αμ\alpha_{\mu} o functie cu variaţia mărginită care, pentru μλ\mu\leqq\lambda, verifică egalitatea αμ(a)=αμ(b)=0\alpha_{\mu}(a)=\alpha_{\mu}(b)=0. Reprezentarea (81) este valabilă dacă derivata a μa,f(μ)\mu-a,f^{(\mu)} este continuă pe [a,b][a,b]. E. Ya. Remez a demonstrat [22] şi formulele

αμ+1(x)=abαμ(x)𝑑x,kμn\displaystyle\alpha_{\mu+1}(x)=-\int_{a}^{b}\alpha_{\mu}(x)dx,\quad k\leqq\mu\leqq n (82)
R[f]=abf(μ)(x)αμ1(x)𝑑x,k+1μn+1\displaystyle R[f]=-\int_{a}^{b}f^{(\mu)}(x)\alpha_{\mu-1}(x)dx,\quad k+1\leqq\mu\leqq n+1 (83)

In particular, funcția de λφn+1,λ\lambda\varphi_{n+1,\lambda} admite o derivată continuă de ordinul n1n-1 pe [a,b][a,b]. Avem deci, ținînd cont de (80), (81),

R[φn+1,λ]=\displaystyle R\left[\varphi_{n+1,\lambda}\right]= n!λb(xλ)𝑑αn1(x)=n!λbαn1(x)𝑑x=\displaystyle n!\int_{\lambda}^{b}(x-\lambda)d\alpha_{n-1}(x)=-n!\int_{\lambda}^{b}\alpha_{n-1}(x)dx=
=n!aλαn1(x)𝑑x=n!αn(λ)\displaystyle=n!\int_{a}^{\lambda}\alpha_{n-1}(x)dx=-n!\alpha_{n}(\lambda)

Din (83) rezultă deci că dacă ff are o derivată de ordinul n+1n+1 continux pe [a,b][a,b], avem reprezentarea

R[f]=1n!abR[φn+1,λ]f(n+1)(x)𝑑xR[f]=\frac{1}{n!}\int_{a}^{b}R\left[\varphi_{n+1,\lambda}\right]f^{(n+1)}(x)dx (84)
  1. 32.

    Să reluăm formula (56) a lui Gauss. Am stabilit formula (57) sub ipoteza continuitătii funcţiei ff pe [1,1][-1,1] si a derivatei sale pe (1,1)(-1,1). Insă în cazul acesta functionala liniară R[f]R[f] este mărginită pe spatiul e0e_{0} al functiilor continue pe [1,1][-1,1], in raport cu norma max |f||f|.

Formula (57) este, în particular, adevărată pentru funcțiile f=φ2m,λf=\varphi_{2m,\lambda} care sînt neconcave de ordinul 2m12m-1. Se deduce că R[φ2m,λ]0R\left[\varphi_{2m,\lambda}\right]\geqq 0 pentru λ[1,1]\lambda\in[-1,1] și, aplicînd teorema 15, rezultă că formula (57) este adevărată sub singura ipoteză a continuitătii functiei f pe intervalul [a,b][a,b].

§4.

  1. 83.

    Vom examina în acest §, fără a întra în prea multe detalii, cazul cînd funcţionala liniară R[f]R[f] nu este de forma simplă.

O funcţională liniară R1[f]R_{1}[f] definită pe \mathscr{F} se numește o majorantă simplă a lui R[f]R[f] dacă : 11^{\circ}. ea este de forma simplă, 22^{\circ}. avem R1[f]>R[f]R_{1}[f]>R[f] pentru orice funcţie convexă fff\in\sqrt{f}

Avem atunci

TEOREMA 16. - Dacă functionala liniară R[f]R[f] definită pe ff admite majorantă simplă, avem

R[f]=K[ξ1,ξ2,,ξn+2;f]K[ξ1,ξ2,,ξn+2;f]R[f]=K\left[\xi_{1},\xi_{2},\ldots,\xi_{n+2};f\right]-K^{\prime}\left[\xi_{1}^{\prime},\xi_{2}^{\prime},\ldots,\xi_{n+2}^{\prime};f\right] (85)

unde : 1.K,K1^{\circ}.K,K^{\prime} sint numere diferite de zero și independente de functia f,2f,2^{\circ}. punctele ξiE,i=1,2,n+2\xi_{i}\in E,i=1,2,\ldots n+2 pe de o parte și punctele ξiE,i==1,2,,n+2\xi_{i}\in E,i==1,2,\ldots,n+2 pe altă parte, sînt distincte (ele pot depinde, în general, de functia ff ).

Intr-adevăr, fie R1[f]R_{1}[f] o majorantă simplă a lui R[f]R[f]. Avem R[f]==R1[f]{R1[f]R[f]}R[f]==R_{1}[f]-\left\{R_{1}[f]-R[f]\right\}, unde functionalele liniare R1[f],R1[f]R_{1}[f],R_{1}[f]- - R[f]R[f] sînt de forma simplă.

Să considerăm o funcțională liniară R[f]R[f] definită pe 𝔽\mathbb{F} şi de forma (85) indicată în teorema 16. Dacă constantele K,KK,K^{\prime} sînt de semne contrare, R[f]R[f] este de forma simplă. Este deci destul să examinăm cazul cînd K,KK,K^{\prime} sînt (diferite de zero şi) de acelaşi semn. Fără să restrîngem generalitatea, putem atunci presupune că ei sînt pozitivi. Avem atunci

L e m a 4. - Dacă functionala liniară R[f]R[f]is defined on the space (f and if it is of the form (85), indicated in Theorem 16, for any function /\inwith the bounded divided difference,
the representation (85) is valid for anytt\in\mathscr{F}(so also for the elements1ini1ui f^\widehat{f}which do not have bounded divided difference).

It is easy to see that Lemma 4 is a consequence of the following:
Lemma 5. - If:11^{\circ}. R[f] is a linear functional defined onf{f},2.K,K2^{\circ}.K,K^{\prime}are two positive numbers,
for anyf𝔽f\in\mathbb{F}whose divided difference is not bounded, one can findn+2n+2distinct pointsxiAND,i=1,2,,n+2\xi_{i}\in E,i=1,2,\ldots,n+2andn+2n+2distinct pointsxiAND,i=1,2,,n+2\xi_{i}^{\prime}\in E,i=1,2,\ldots,n+2, so that we have (85).

Let us assume, for the sake of clarity, that the divided difference of the functionffis not bounded above, By virtue of Theorem 4, if the divided difference of this function takes the valuemm, it will take any value greater thanmm. Be it thenmma value taken from the difference

12 - Mathematics studies and research
[x1,x2,,xn+2;f]\left[\xi_{1}^{\prime},\xi_{2}^{\prime},\ldots,\xi_{n+2}^{\prime};f\right]a divided difference that takes a value>mKR[f]K>\frac{mK-R[f]}{K^{\prime}}and[x1,x2,,xn+2;t]\left[\xi_{1},\xi_{2},\ldots,\xi_{n+2};t\right]a divided difference that takes the value

1K{K[x1,x2,,xn+2;f]+R[f]}>m.\frac{1}{K}\left\{K^{\prime}\left[\xi_{1}^{\prime},\xi_{2}^{\prime},\ldots,\xi_{n+2}^{\prime};f\right]+R[f]\right\}>m.

Formula (85) results.
The same procedure is followed if the divided difference of the functionffit is not bounded inferiorly.

Lemma 5 is therefore proven.
We recall that the notion of divided difference, of simplicity of linear functionals and the spaces considered are in the sense of § 1.

It is clear that instead of simple majorities we can use simple minorities. Linear functionalR1[f]R_{1}[f]defined on\mathscr{F}is called a simple minor forR[t]R[t]if it is of simple form and ifR1[t]<<R[f]R_{1}[t]<<R[f]for any concave functionff.

To be able to put a linear functionalR[f]R[f]in the form ( 85 ), it is therefore sufficient to know a simple majorant (or a minorant). For example, the linear functional (58), which cancels out on the functions (19) and which can therefore be put in the form (59), has as a simple majorant the linear functional

i=1mn1|mi|+mi2[xi,xi+1,,xi+n+1;f]+m[x1,x2,,xn+2;f],\sum_{i=1}^{m-n-1}\frac{\left|\mu_{i}\right|+\mu_{i}}{2}\left[x_{i},x_{i+1},\ldots,x_{i+n+1};f\right]+\mu\left[x_{1}^{\prime},x_{2}^{\prime},\ldots,x_{n+2}^{\prime};f\right],

wherem\muis a positive number andxi,n+2x_{i}^{\prime},n+2distinct points of the intervalANDE. All linear functionals of the form (58) can therefore be put into the form (85), indicated in Theorem 16.
34. If the linear functionalR[f]R[f]is of the form (85), the differenceKK=R[fn+1]K-K^{\prime}=R\left[f_{n+1}\right]has a perfectly determined value. Suppose thatK,KK,K^{\prime}are positive. We can then replaceK,KK,K^{\prime}byK+e,K+eK+\varepsilon,K^{\prime}+\varepsilonrespectively,e\varepsilonbeing an arbitrary positive number. Indeed, if we have (85) for af𝔽f\in\mathbb{F}given, we haveR[f]=R1[f]R2[f]R[f]=R_{1}[f]-R_{2}[f], whereR1[f]==K[x1,x2,,xn+2;f]+e[x1,x2,,xn+2;f],R2[f]=K[x1,x2,,xn+2;f]++e[x1,x2,,xn+2;f]R_{1}[f]==K\left[\xi_{1},\xi_{2},\ldots,\xi_{n+2};f\right]+\varepsilon\left[x_{1},x_{2},\ldots,x_{n+2};f\right],R_{2}[f]=K^{\prime}\left[\xi_{1}^{\prime},\xi_{2}^{\prime},\ldots,\xi_{n+2}^{\prime};f\right]++\varepsilon\left[x_{1},x_{2},\ldots,x_{n+2};f\right], wherexix_{i}saintn+2n+2distinct points of the intervalANDEWe can look atR1[f],R2[f]R_{1}[f],R_{2}[f]as linear functionals defined on\mathscr{F}. Then they are of the simple form. The stated property results from observing thatR1[fn+1]=K+e,R2[fn+1]=K+eR_{1}\left[f_{n+1}\right]=K+\varepsilon,R_{2}\left[f_{n+1}\right]=K^{\prime}+\varepsilon.

IfRf]R\|f]is of the form (85) but is not of simple form, the coefficientsK,KK,K^{\prime}, assumed positive, have some lower bounds whose values ​​are of interest especially whenR[f]R[f]is the remainder of an approximation formula. In this sense we will examine an important particular case in the next no.
35. Let us suppose again that we are in the particular case (21), (21') and consider a linear functionalR[f]R[f]defined and bounded in spaceandke_{k}considered at No. 30. We have

THEOREM 17. - If:11^{\circ}. linear functionalR[f]R[f]is defined onk\bigodot_{k}, limited in relation to the norm (78) and the degree of accuracynn, withR[xn+1]>0,2R\left[x^{n+1}\right]>0,2^{\circ}. A is the upper edge ofR[f]R[f]for the functionsfkf\in\mathbb{C}_{k}whose difference divided by the ordern+1n+1remains contained in[0,1][0,1]andB=AR[xn+1]B=A-R\left[x^{n+1}\right],
for anythinge>0\varepsilon>0, linear functionalR[f]R[f]is of the form (85), indicated by Theorem 16, whereK=A+e,K=B+eK=A+\varepsilon,K^{\prime}=B+\varepsilon.

From the demonstration it will follow thatA,BA,Bare finite.
We haveA>0A>0because, in particular,xn+1x^{n+1}has its difference divided by the ordern+1n+1contained in[0,1][0,1]We have obviouslyB0B\geq 0.

If we consider the functions (69), by the formulaRm[f]=R[fm]R_{m}[f]=R\left[f_{m}\right]we define a linear functional which, formm\rightarrow\infty, tends towardsR[f]R[f]for anythingfandkf\in e_{k}Let's put

Rm+[f]=(n+1)hi=1m1Dn+1i[f]R[fn+1,li+n]+|R[fn+1,li+n]|2R_{m}^{+}[f]=(n+1)h\cdot\sum_{i=1}^{m-1}D_{n+1}^{i}[f]\frac{R\left[\varphi_{n+1,\lambda_{i+n}}\right]+\left|R\left[\varphi_{n+1,\lambda_{i+n}}\right]\right|}{2} (86)

and let's note withandke_{k}^{*}his/her subsetandke_{k}made up of the functionsffwhich have their differences divided by the ordern+1n+1bounded. Moreover, any function defined on[a,b][a,b], having its difference divided by the ordern+1n+1bordered, belongs tok\ell_{k}Let us note that the functionx,R[fn+1,x]x,R\left[\varphi_{n+1,x}\right]being continuous on[a,b][a,b], the string with positive terms

{(n+1)hi=1mn1R[fn+1;li+n]+|R[fn+1,li÷n]|2}\left\{(n+1)h\sum_{i=1}^{m-n-1}\frac{R\left[\varphi_{n+1;\lambda_{i+n}}\right]+\left|R\left[\varphi_{n+1,\lambda_{i\div n}}\right]\right|}{2}\right\} (87)

tend, formm\rightarrow\infty, towards a finite and well-determined limit equal to

A=(n+1)abR[fn+1,x]+|R|fn+1,x]2𝑑xA=(n+1)\int_{a}^{b}\frac{\left.R\left[\varphi_{n+1,}x\right]+|R|\varphi_{n+1},x\right]\mid}{2}dx (88)

It follows that the sequence (87) is bounded. Iff𝒞kf\in\mathcal{C}_{k}^{*}, the string{Rm+[f]}\left\{R_{m}^{+}[f]\right\}is also bounded. One can extract from this sequence a partial sequence convergent to the functionalR+[f]R^{+}[f]It is easy to see that the functionalR+[f]R^{+}[f]thus defined onandke_{k}^{*}is linear and vanishes on any polynomial of degreennBut we haveRm+[f]>0,Rm[f]Rm[f]R_{m}^{+}[f]>0,R_{m}^{\star}[f]\geqq R_{m}[f]iffandkf\in e_{k}is convex, soR+[f]0,R+[f]R[f]R^{+}[f]\geqq 0,R^{+}[f]\geqq R[f]iffkf\in\mathbb{C}_{k}^{*}is convex. It immediately follows that ife\varepsilonis a positive number andx1,x2,,xn+2,n+2x_{1},x_{2},\ldots,x_{n+2},n+2fixed points of the intervalANDE, the linear functionalR1[f]=R+[f]+e[x1,x2,,xn+2;f]R_{1}[f]=R^{+}[f]+\varepsilon\left[x_{1},x_{2},\ldots,x_{n+2};f\right]is a simple majority ofR[f]R[f]It is easy to see thatR1[xn+1]=A+eR_{1}\left[x^{n+1}\right]=A+\varepsilon, whereAAis given by formula (88).

It remains to be proven that the numberAA, given by formula (88), coincides with the upper edge ofR[f]R[f]iffftraverses the set of functions whose difference divided by the ordern+1n+1remains contained in[0,1][0,1]Ifffis such a function, it is clear thatRm[f]R_{m}[f]does not exceed the general (corresponding) term of the sequence ( 87 ). Taking the limit, it follows thatR[f]R[f]don't blink -
step onAA. Be it nowe\varepsilonan arbitrary positive number. Let us take into account the continuity of the functionx,R[fn+1,x]x,R\left[\varphi_{n+1,x}\right], therefore by the continuity and non-negativity of the functionx,12{R[fn+1,x]+R[fn+1,x]}x,\frac{1}{2}\left\{R\left[\varphi_{n+1,x}\right]+\mid R\left[\varphi_{n+1,x}\right]\right\}and note that the points at which a function continues on[a,b][a,b]cancels out, formstheoclosed set. It follows that we can find a finite numberkkof disjoint intervals[ai,bi],i=1,2,,k\left[\alpha_{i},\beta_{i}\right],i=1,2,\ldots,k, belonging to(a,b)(a,b)and so that the functionR[fn+1,x]R\left[\varphi_{n+1,x}\right]to be nonnegative on these intervals and so that we have

(n+1)i=1kaibiR[fn+1,x]𝑑x>Ae2(n+1)\sum_{i=1}^{k}\int_{a_{i}}^{\beta_{i}}R\left[\varphi_{n+1,x}\right]dx>A-\frac{\varepsilon}{2} (89)

We can assumea<a1<b1<a2<b2<<ak<bk<ba<\alpha_{1}<\beta_{1}<\alpha_{2}<\beta_{2}<\ldots<\alpha_{k}<\beta_{k}<b. EitherM=max(n+1)|R[fn+1,x]| and let's choose the points ai,bii=1,2,,kM=\max_{(n+1)\left|R\left[\varphi_{n+1,x}\right]\right|\text{ si să alegem punctele }\alpha_{i}^{\prime},\quad\beta_{i}^{\prime}\text{, }}i=1,2,\ldots,kso that we havea<a1<a1,bk<bk<b,bi1<bi1<<ai<ai,i=2,3,,ka<\alpha_{1}^{\prime}<\alpha_{1},\beta_{k}<\beta_{k}^{\prime}<b,\beta_{i-1}<\beta_{i-1}^{\prime}<<\alpha_{i}^{\prime}<\alpha_{i},\quad i=2,3,\ldots,kand as

Mi=1k(aiai+bibi)<e2.M\sum_{i=1}^{k}\left(\alpha_{i}-\alpha_{i}^{\prime}+\beta_{i}^{\prime}-\beta_{i}\right)<\frac{\varepsilon}{2}. (90)

Be it nowffa function whose derivative of the ordern+1n+1exists and is continuous on[a,b][a,b], this derivative reducing to(n+1)(n+1)  ! on the intervals[ai,bi],i=1,2,,k\left[\alpha_{i},\beta_{i}\right],i=1,2,\ldots,k, to 0 on the intervals[a,aj],[bk,b]\left[a,\alpha_{j}^{\prime}\right],\left[\beta_{k}^{\prime},b\right],[bi1,ai],i=2,3,,k\left[\beta_{i-1}^{\prime},\alpha_{i}\right],i=2,3,\ldots,kand one linear function on each of the intervals[ai,ai],[bi,bi],i=1,2,,k\left[\alpha_{i}^{\prime},\alpha_{i}\right],\left[\beta_{i},\beta_{i}^{\prime}\right],i=1,2,\ldots,kFunctionffconsidered to belong toandke_{k}and the formula (29) of the average shows us that the difference divided by the ordern+1n+1remains contained in[0,1][0,1]Taking into account the representation ( 85 ), we have for this function

R[f]=(n+1)i=1kaibiR[fn+1,x]𝑑x+\displaystyle R[f]=(n+1)\sum_{i=1}^{k}\int_{a_{i}}^{\beta_{i}}R\left[\varphi_{n+1,x}\right]dx+
+(n+1){i=1kaiaiR[fn+1,x]f(n+1)(x)(n+1)!𝑑x+i=1kbibiR[fn+1,x]f(n+1)(x)(n+1)!𝑑x}.\displaystyle+(n+1)\left\{\sum_{i=1}^{k}\int_{a_{i}^{\prime}}^{a_{i}}R\left[\varphi_{n+1,x}\right]\frac{f^{(n+1)}(x)}{(n+1)!}dx+\sum_{i=1}^{k}\int_{\beta_{i}}^{\beta_{i}^{\prime}}R\left[\varphi_{n+1,x}\right]\frac{f^{(n+1)}(x)}{(n+1)!}dx\right\}. (91)

But

(n+1)|i=1kaiai+i=1kbibiR[fn+1,x]f(n+1)(x)(n+1)!𝑑x|(n+1){i=1kaiai+i=1kbib|R[fn+1,x]|dx}Mi=1k(aiai+bibi)<e2\begin{gathered}(n+1)\left|\sum_{i=1}^{k}\int_{a_{i}}^{a_{i}}+\sum_{i=1}^{k}\int_{\beta_{i}}^{\beta_{i}^{\prime}}R\left[\varphi_{n+1,x}\right]\frac{f^{(n+1)}(x)}{(n+1)!}dx\right|\leqq\\ \leqq(n+1)\left\{\sum_{i=1}^{k}\int_{a_{i}^{\prime}}^{a_{i}}+\sum_{i=1}^{k}\int_{\beta_{i}}^{\beta^{\prime}}\left|R\left[\varphi_{n+1,x}\right]\right|dx\right\}\leqq M\sum_{i=1}^{k}\left(\alpha^{i}-\alpha_{i}^{\prime}+\beta_{i}^{\prime}-\beta_{i}\right)<\frac{\varepsilon}{2}\end{gathered}

Taking into account (89), (92), from formula (91) it follows thatR[f]>AeR[f]>A-\varepsilon. NumberAAis therefore the upper bound indicated in the statement of the theorem.

Theorem 17 is therefore proven.
In this theorem we assumedR[xn+1]>0R\left[x^{n+1}\right]>0. Otherwise, so ifR[xn+1]<0R\left[x^{n+1}\right]<0, the property and the proof are analogous. In this caseB>0,A0B>0,A\geqq 0.

In casesA=0¯A=\overline{0}orB=0B=0, functionalR[f]R[f]is of simple form.
It is easy to show that iffCkf\in C_{k}has a derivative of ordern+1n+1continue on[a,b][a,b], we have

R[f]=1(n+1)!{Af(n+1)(x)Bf(n+1)(or)},x,or[a,b]R[f]=\frac{1}{(n+1)!}\left\{Af^{(n+1)}(\xi)-Bf^{(n+1)}(\eta)\right\},\quad\xi,\eta\in[a,b]

Iffϵandkf\epsilon e_{k}^{*}and ifddis the upper bound of the absolute value of the difference divided by the ordern+1n+1his/hertt, we have the delimitation

|R[f]|(A+B)d|R[f]|\leqq(A+B)d
  1. 36.

    There are other forms in which a linear functional can be putR[f]R[f], so the remainder of a linear approximation formula. These expressions are of interest especially whenR[f]R[f]it is not of simple form.

Let us suppose that we are in the particular case (21), (21') and let us suppose thatR[f]R[f]is a linear functional defined and of degree of accuracynnonf^\widehat{f}Let us consider a decomposition of the form

R[f]=R1[f]+{R[f]R1[f]}R[f]=R_{1}[f]+\left\{R[f]-R_{1}[f]\right\} (93)

whereR1[t]R_{1}[t]is a linear functional defined on(t(\tauand where the linear functional (also defined on(f)R[f]R1[f](f)R[f]-R_{1}[f]has a degree of accuracyn+p>nn+p>n. Then ifR1[f]R_{1}[f]andR[f]R1[f]R[f]-R_{1}[f]are of simple form, we have

R[f]=R[xn+1][x1,x2,,xn+2;f]+K[x1,x2,,xn+p+2;f]R[f]=R\left[x^{n+1}\right]\left[\xi_{1},\xi_{2},\ldots,\xi_{n+2};f\right]+K\left[\xi_{1}^{\prime},\xi_{2}^{\prime},\ldots,\xi_{n+p+2}^{\prime};f\right] (94)

whereK0K\neq 0is independent of the functionffandxi,xi\xi_{i},\xi_{i}are groups ofn+2n+2 resp. n+p+2n+p+2distinct points inANDE.

Without claiming to make a general theory here, we will show, by two examples, how a representation of the form (94) can actually be found for the rest of certain approximation formulas.
37. Let us consider Hardy's quadrature formula,

06f(x)𝑑x=0,28[f(0)+f(6)]+1,62[f(1)+f(5)]+2,2f(3)+R[f]\int_{0}^{6}f(x)dx=0,28[f(0)+f(6)]+1,62[f(1)+f(5)]+2,2f(3)+R[f]

The degree of accuracy of the remainderR[f]R[f]is 5. A simple calculation shows us thatR[f6,3]=8150>0,R[f6,5]=17150<0R\left[\varphi_{6,3}\right]=\frac{81}{50}>0,\quad R\left[\varphi_{6,5}\right]=-\frac{17}{150}<0and so, by virtue of Theorem 15, the remainder is not of simple form.

To putR[f]R[f]in the form (94) it is advantageous to first consider the linear functionalR[f]=R[f]R^{*}[f]=R\left[f^{\prime}\right], which we have already considered in the previous §. Indeed, it is enough to find a decomposition of the form (94) for this linear functional. The corresponding decomposition forR[f]R[f], it results immediately.

we
R[f]=63[0,0,1,1,3,3,5,5;f]+190,8[0,1,1,3,3,5,5,6;f]R^{*}[f]=-63[0,0,1,1,3,3,5,5;f]+190,8[0,1,1,3,3,5,5,6;f]-

63[1,1,3,3,5,5,6,6;f]-63[1,1,3,3,5,5,6,6;f]

Whether

R1[f]=m1[0,0,1,1,3,3,5,5;f]+m2[0,1,1,3,3,5,5,6;f]+\displaystyle R_{1}[f]=\mu_{1}[0,0,1,1,3,3,5,5;f]+\mu_{2}[0,1,1,3,3,5,5,6;f]+
+m3[1,1,3,3,5,5,6,6;f]\displaystyle+\mu_{3}[1,1,3,3,5,5,6,6;f] (95)

where
we have thenm1+m2+m3=64,8\mu_{1}+\mu_{2}+\mu_{3}=64,8.

R[f]R1[f]=6(63+m1)[0,0,1,1,3,3,5,5,6;f]R^{*}[f]-R_{1}[f]=6\left(63+\mu_{1}\right)[0,0,1,1,3,3,5,5,6;f]- (96)
6(63+m3)[0,1,1,3,3,5,5,6,6;f]-6\left(63+\mu_{3}\right)[0,1,1,3,3,5,5,6,6;f] (97)

which has a degree of accuracy>6>6The linear functionals
(95), (97) are of simple form ifm1=m3,m2\mu_{1}=\mu_{3},\mu_{2}are nonnegative. We thus find the following expression for the remainder in Hardy's formula,

R[f]=9700{6![x1,x2,,x2;f]5(63+m1)6488![or1,or2,,or9;f]}R[f]=\frac{9}{700}\left\{6!\left[\xi_{1},\xi_{2},\ldots,\xi_{2};f\right]-\frac{5\left(63+\mu_{1}\right)}{648}8!\left[\eta_{1},\eta_{2},\ldots,\eta_{9};f\right]\right\}

whereffis continuous on[0,6],xi[0,6],\xi_{i}there are 7 distinct points andori\eta_{i}There are 9 distinct points in the interval (0.6).

From the particular method of demonstration it follows that in this formula we have0m132,40\leqq\mu_{1}\leqq 32,4.

Ifffhas a continuous derivative of order 8 on ( 0.6 ), we have

R[f]=9700{f(6)(x)5(63+m1)648f(8)(or)},x,or(0,6)R[f]=\frac{9}{700}\left\{f^{(6)}(\xi)-\frac{5\left(63+\mu_{1}\right)}{648}f^{(8)}(\eta)\right\},\quad\xi,\eta\in(0,6)

If we put in this formulam1=95\mu_{1}=\frac{9}{5}, we find the well-known remainder [24]

R[f]=9700{f(6)(x)12f(8)(or)},x,or(0,6)R[f]=\frac{9}{700}\left\{f^{(6)}(\xi)-\frac{1}{2}f^{(8)}(\eta)\right\},\quad\xi,\eta\in(0,6) (98)

But we can also takem1=0\mu_{1}=0and then we find

R[f]=9700{f(6)(x)3572f(8)(or)},x,or(0,6)R[f]=\frac{9}{700}\left\{f^{(6)}(\xi)-\frac{35}{72}f^{(8)}(\eta)\right\},\quad\xi,\eta\in(0,6)

where the coefficient3572\frac{35}{72}is smaller than the corresponding coefficient12\frac{1}{2}from formula (98).
38. As a second application, let us take Weddle's quadrature formula,
6
06f(x)𝑑x=0,3[f(0)+f(2)+f(4)+f(6)]+1,5[f(1)+f(5)]+1,8f(3)+R[f]\int_{0}^{6}f(x)dx=0,3[f(0)+f(2)+f(4)+f(6)]+1,5[f(1)+f(5)]+1,8f(3)+R[f], whose remainder is still of accuracy degree 5. We haveR[f6,3]=310>0R\left[\varphi_{6,3}\right]=\frac{3}{10}>0,R[f6,4]=1330<0R\left[\varphi_{6,4}\right]=-\frac{13}{30}<0, so the remainder is not of simple form. Proceeding as in the previous example, we have

518R[t]=3[0,0,1,1,2,2,3,3;t]4[0,1,1,2,2,3,3,4;t]++4[1,2,2,3,3,4,4,5;t]4[2,3,3,4,4,5,5,6;t]3[3,3,4,4,5,5,6,6;t]\begin{gathered}\frac{5}{18}R^{*}[t]=-3[0,0,1,1,2,2,3,3;t]-4[0,1,1,2,2,3,3,4;t]+\\ +4[1,2,2,3,3,4,4,5;t]-4[2,3,3,4,4,5,5,6;t]-\\ -3[3,3,4,4,5,5,6,6;t]\end{gathered}

and we take

518R1[f]=m1[0,0,1,1,2,2,3,3;f]+m2[0,1,1,2,2,3,3,4;f]+\displaystyle\frac{5}{18}R_{1}[f]=\mu_{1}[0,0,1,1,2,2,3,3;f]+\mu_{2}[0,1,1,2,2,3,3,4;f]+
+m3[1,1,2,2,3,3,4,4;f]+m4[1,2,2,3,3,4,4,5;f]+\displaystyle+\mu_{3}[1,1,2,2,3,3,4,4;f]+\mu_{4}[1,2,2,3,3,4,4,5;f]+ (99)
+m3[2,2,3,3,4,4,5,5;f]+m2[2,3,3,4,4,5,5,6;f]+\displaystyle+\mu_{3}[2,2,3,3,4,4,5,5;f]+\mu_{2}[2,3,3,4,4,5,5,6;f]+
+m1[3,3,4,4,5,5,6,6;f]\displaystyle+\mu_{1}[3,3,4,4,5,5,6,6;f]
2(m1+m2+m3)+m4=10\displaystyle 2\left(\mu_{1}+\mu_{2}+\mu_{3}\right)+\mu_{4}=-10

and then we have

518[R[f]\displaystyle-\frac{5}{18}\left[R^{*}[f]\right. R1[f]]=16(m1+3)[0,0,1,1,2,2,3,3,4,4;f]+\displaystyle\left.-R_{1}[f]\right]=16\left(\mu_{1}+3\right)[0,0,1,1,2,2,3,3,4,4;f]+ (101)
+20(2m1+m2+10)[0,1,1,2,2,3,3,4,4,5;f]+\displaystyle+20\left(2\mu_{1}+\mu_{2}+10\right)[0,1,1,2,2,3,3,4,4,5;f]+
+16(3m1+2m2+m3+17)[1,1,2,2,3,3,4,4,5,5;f]+\displaystyle+16\left(3\mu_{1}+2\mu_{2}+\mu_{3}+17\right)[1,1,2,2,3,3,4,4,5,5;f]+
+20(2m1+m2+10)[1,2,2,3,3,4,4,5,5,6;f]+\displaystyle+20\left(2\mu_{1}+\mu_{2}+10\right)[1,2,2,3,3,4,4,5,5,6;f]+
+16(m1+3)[2,2,3,3,4,4,5,5,6,6;f]\displaystyle+16\left(\mu_{1}+3\right)[2,2,3,3,4,4,5,5,6,6;f]

Let's takem1=3,m2=2,m3=m4=0\mu_{1}=-3,\mu_{2}=-2,\mu_{3}=\mu_{4}=0Then (100) is true.

R[f]=1140{6![x1,x2,,x7;f]+158![or1,or2,,m3;f]}R[f]=-\frac{1}{140}\left\{6!\left[\xi_{1},\xi_{2},\ldots,\xi_{7};f\right]+\frac{1}{5}8!\left[\eta_{1},\eta_{2},\ldots,\mu_{3};f\right]\right\}

function//and the pointsxi,ori\xi_{i},\eta_{i}verifying the same conditions as in the previous example (no. 37). If the functionffhas a continuous derivative of order 8 on(0,6)(0,6), we have

R[f]=1140{f(6)(x)+15f(8)(or)},x,or(0,6)R[f]=-\frac{1}{140}\left\{f^{(6)}(\xi)+\frac{1}{5}f^{(8)}(\eta)\right\},\quad\xi,\eta\in(0,6)

In the well-known formula [24],

R[t]=1140{f(6)(x)+910f(8)(or)}x,or(0,6)R[t]=-\frac{1}{140}\left\{f^{(6)}(\xi)+\frac{9}{10}f^{(8)}(\eta)\right\}\quad\xi,\eta\in(0,6)

the coefficient of the 8th order derivative is 4.5 times greater in absolute value.

Let us also observe that if, in addition to (100), we also have

20m1+9m2+2m3=96,20\mu_{1}+9\mu_{2}+2\mu_{3}=-96, (102)

we can write

518{R[f]R1[f]}=400(m1+3)[0,0,1,1,2,2,3,3,4,4,5,5;f]+\displaystyle-\frac{5}{18}\left\{R^{*}[f]-R_{1}[f]\right\}=400\left(\mu_{1}+3\right)[0,0,1,1,2,2,3,3,4,4,5,5;f]+
+120(18m1+5m2+74)[0,1,1,2,2,3,3,4,4,5,5,6;f]+\displaystyle+120\left(18\mu_{1}+5\mu_{2}+74\right)[0,1,1,2,2,3,3,4,4,5,5,6;f]+ (103)
+400(m1+3)[1,1,2,2,3,3,4,4,5,5,6,6;f]\displaystyle+400\left(\mu_{1}+3\right)[1,1,2,2,3,3,4,4,5,5,6,6;f]

If we takem1=5111,m2=411,m3=m4=0\mu_{1}=-\frac{51}{11},\mu_{2}=-\frac{4}{11},\mu_{3}=\mu_{4}=0, the equalities (100), (102) are verified and the linear functionals (99), (103) are of simple form. For the restR[t]R[t]of Weddle's formula we obtain

R[f]=1140{6![x1,x2,,x7;f]61181510![or1,or2,,or11;f]}R[f]=-\frac{1}{140}\left\{6!\left[\xi_{1},\xi_{2},\ldots,\xi_{7};f\right]-\frac{61}{1815}10!\left[\eta_{1},\eta_{2},\ldots,\eta_{11};f\right]\right\}

whereffis continuous on[0,6],xi[0,6],\xi_{i}there are 7 distinct points andori\eta_{i}there are 11 distinct points of the interval(0,6)(0,6)

If the functionffhas a continuous derivative of order 10 on ( 0.6 ), we have

R[f]={1140f(6)(x)618115f(10)(or)},x,or(0,6).R[f]=-\left\{\frac{1}{140}f^{(6)}(\xi)-\frac{61}{8115}f^{(10)}(\eta)\right\},\quad\xi,\eta\in(0,6).
1959

Related Posts