Abstract

We focus on a special class of ideal projectors, subspaces, which possesses two classes of D-invariant polynomial subspaces. The first is a classical type, while the second is a new class. With matrix computation, we discretize this class of ideal projectors into a sequence of Lagrange projectors.

1. Introduction

Polynomial interpolation is to construct a polynomial 𝑝 belonging to a finite-dimensional subspace of 𝔽[𝐱] from a set of data that agrees with a given function 𝑓 at the data set, where 𝔽[𝐱]=𝔽[𝑥1,,𝑥𝑑] denotes the polynomial ring in 𝑑 variables over the field 𝔽. It is important to make the comment that 𝔽 is the real field or the complex field in this paper. Univariate polynomial interpolation has a well-developed theory, while the multivariate one is very problematic since a multivariate interpolation polynomial is determined not only by the cardinal but also by the geometry of the data set, cf. [1, 2].

Recently, more and more people are getting interested in ideal interpolation, which is defined by an ideal projector on 𝔽[𝐱], namely, a linear idempotent operator on 𝔽[𝐱] whose kernel is an ideal, cf. [3]. When the kernel of an ideal projector 𝑃 is the vanishing ideal of certain finite set Ξ in 𝔽𝑑, 𝑃 is a Lagrange projector which provides the Lagrange interpolation on Ξ.

It is well known that an ideal projector can be characterized completely by the range of its dual projector, cf. [47]. Given a finite-rank linear projector 𝑃 on 𝔽[𝐱], the kernel of 𝑃 is ideal if and only if the range of its dual projector is of the form:𝝃Ξ𝛿𝝃𝒬𝝃(𝐷),(1.1) with some finite point set Ξ𝔽𝑑, 𝐷-invariant finite-dimensional polynomial subspace 𝒬𝝃𝔽[𝐱] for each 𝝃Ξ. 𝛿𝝃 denotes the evaluation functional at the point 𝝃, and 𝒬𝝃(𝐷) will be explained in the next section.

In the univariate case, for an integer 𝑛, there is only one 𝐷-invariant polynomial subspace of degree less than 𝑛, which implies that every univariate ideal projector can be viewed as a limiting case of Lagrange projectors, cf. [8]. This prompted de Boor to define the Hermite projector as the pointwise limit of Lagrange projectors and conjecture that every ideal projector in [𝐱] is an Hermite projector in [9].

However, Shekhtman [10] constructed a counterexample to this conjecture for every 𝑑3. In the same paper, Shekthman also showed that the conjecture is true for bivariate complex projectors with the help of Fogarty Theorem (see [11]). Later, using linear algebra, de Boor and Shekhtman [12] reproved the same result. Specifically, Shekhtman [13] completely analyzed the bivariate ideal projectors which are onto the space of polynomials of degree less than 𝑛 over real or complex field and verified the conjecture in this particular case.

In this paper, we consider the converse problem: how to generate a sequence of Lagrange projectors practically such that this sequence of Lagrange projectors converges pointwise to a given Hermite projector. We focus on a special class of ideal projectors, which possesses two classes of 𝐷-invariant polynomial subspaces. Sections 3 and 4 discuss these two classes of 𝐷-invariant subspaces, respectively. In Section 5, we discretize this class of ideal projectors into a sequence of Lagrange projectors. The next section is devoted a preparation for the paper.

2. Preliminaries

In this section, we will introduce some notation used throughout the paper and give some background results. For more details, we refer the reader to [9, 14].

We use to stand for the monoid of nonnegative integers and boldface type for tuples with their entries denoted by the same letter with subscripts, for example, 𝜶=(𝛼1,,𝛼𝑑). For arbitrary 𝜶𝑑, we define 𝜶!=𝛼1!𝛼𝑑!.

Henceforward, ≤ will denote the usual product order on 𝑑, that is, for arbitrary 𝜶, 𝜷𝑑, 𝜶𝜷 if and only if 𝛼𝑖𝛽𝑖,𝑖=1,,𝑑. A finite subset 𝒜𝑑 is lower if for every 𝜶𝒜, 𝟎𝜷𝜶 implies 𝜷𝒜.

A monomial 𝐱𝜶𝔽[𝐱] is a power product of the form 𝑥𝛼11𝑥𝛼𝑑𝑑 with 𝜶𝑑. Thus, a polynomial 𝑝 in 𝔽[𝐱] can be expressed as𝑝=𝜶𝑑̂𝑝(𝜶)𝐱𝜶,(2.1) with ̂𝑝(𝜶)𝔽 nonzero. For a polynomial 𝑝(𝐱) as in (2.1), we write[𝐱][𝐱]𝑝(𝐷)𝔽𝔽𝑓𝜶𝑑𝜕̂𝑝(𝜶)𝜶1𝑓𝜕𝑥𝛼11𝜕𝑥𝛼𝑑𝑑,(2.2) for its associated differential operator. For a finite-dimensional polynomial subspace 𝒬, we define𝒬(𝐷)=span𝔽{𝑞(𝐷)𝑞𝒬}.(2.3) If the polynomial subspace 𝒬 is closed under differentiation, then we call it a 𝐷-invariant polynomial subspace.

Let 𝑃 be a finite-rank ideal projector on 𝔽[𝐱]. The range and kernel of 𝑃 will be described as[𝐱][𝐱][𝐱]ran𝑃={𝑔𝔽𝑔=𝑃𝑓forsome𝑓𝔽},ker𝑃={𝑔𝔽𝑃𝑔=0}.(2.4) Let 𝑃 be the dual projector of 𝑃, then ran𝑃 is the set of interpolation conditions matched by 𝑃. If 𝐪=(𝑞1,,𝑞𝑠) is an 𝔽-basis for ran𝑃 and 𝝀=(𝜆1,,𝜆𝑠) an 𝔽-basis for ran𝑃, then their Gram matrix: 𝝀𝑇𝜆𝐪=𝑖𝑞𝑗1𝑖,𝑗𝑠(2.5) is invertible.

Lemma 2.1 (see [15, page 90]). Let 𝑛 and 𝑚 be arbitrary nonnegative integers satisfying 𝑛𝑚. Then 𝑚𝑟=0(1)𝑚𝑟𝑚𝑟𝑟𝑛=𝑚!,𝑛=𝑚;0,0𝑛<𝑚.(2.6)

3. The First Class of 𝐷-Invariant Polynomial Subspaces

Let 𝜶=(𝛼1,,𝛼𝑑)𝑑 and 𝝆=(𝝆1,𝝆2,,𝝆𝑑)(𝔽𝑑)𝑑 in which𝝆1=𝜌1,1,,𝜌1,𝑑,𝝆2=𝜌2,1,,𝜌2,𝑑,,𝝆𝑑=𝜌𝑑,1,,𝜌𝑑,𝑑(3.1) are 𝔽-linearly independent unit vectors. Then we define a differential operator 𝐷𝜶𝝆𝔽[𝐱]𝔽[𝐱] by the formula: 𝐷𝜶𝝆𝜕𝑓=𝜶1𝑓𝜕𝝆𝛼11𝜕𝝆𝛼𝑑𝑑.(3.2)

Lemma 3.1. Let 𝜶=(𝛼1,,𝛼𝑑)𝑑, 𝝆=(𝝆1,𝝆2,,𝝆𝑑)(𝔽𝑑)𝑑 be as above, and let 𝑝=𝑑𝑖=1𝝆𝑖𝐱𝛼𝑖,(3.3) where denotes the Euclidean product. Then 𝑝(𝐷)=𝐷𝜶𝝆.

Given a lower set 𝔡𝑑 and 𝝆=(𝝆1,𝝆2,,𝝆𝑑)(𝔽𝑑)𝑑 as above, then the subspace:𝒬=span𝔽𝑑𝑖=1𝝆𝑖𝐱𝛼𝑖𝜶𝔡(3.4) is obviously 𝐷-invariant. The following proposition discretizes this functional subspace 𝛿𝝃𝒬(𝐷) into a sequence of evaluation functional subspaces.

Proposition 3.2. Let 𝝆=(𝝆1,𝝆2,,𝝆𝑑)(𝔽𝑑)𝑑 be as above, a nonzero number in 𝔽, and let 𝜶𝑑, 𝝃𝔽𝑑. Then for arbitrary 𝑝𝔽[𝐱]1𝜶1𝟎𝜷𝜶(1)𝜶𝜷1𝜶𝜷𝑝𝝃+𝑑𝑖=1𝛽𝑖𝝆𝑖=𝐷𝜶𝝆𝑝(𝝃)+𝑂(),(3.5) where 𝜶𝜷=𝑑𝑖=1𝛼𝑖𝛽𝑖, and the remainder 𝑂() is a polynomial in .

Proof. For convenience, let 𝜑=𝟎𝜷𝜶(1)𝜶𝜷1𝜶𝜷𝑝𝝃+𝑑𝑘=1𝛽𝑘𝝆𝑘.(3.6) Then we can obtain that for 𝑚=0,,𝜶1: 𝑑𝑚𝜑(0)𝑑𝑚=𝛾11𝜸++𝑑𝟏=𝑚𝜌𝛾1,11,1𝜌𝛾1,𝑑𝑑,1𝜌𝛾𝑑,11,𝑑𝜌𝛾𝑑,𝑑𝑑,𝑑𝜕𝑚𝑝(𝝃)𝜕𝑥𝜸111𝜕𝑥𝜸𝑑𝟏𝑑×𝑚𝛾1,1𝛾1,𝑑,,𝛾𝑑,1,,𝛾𝑑,𝑑𝑑𝑗=1𝛼𝑗𝛽𝑗=0(1)𝛼𝑗𝛽𝑗𝛼𝑗𝛽𝑗𝛽𝑑𝑖=1𝛾𝑖,𝑗𝑗,(3.7) with 𝜸𝑖=(𝛾𝑖,1,,𝛾𝑖,𝑑)𝑑, 1𝑖𝑑.
In the following, we will simplify the right-hand side of the above equality. There are two cases which must be examined.
Case 1 (0𝑚𝜶11). In this case, there must exist some 1𝑗𝑑 such that 𝑑𝑖=1𝛾𝑖,𝑗<𝛼𝑗. By Lemma 2.1, it follows for such 𝑗: 𝛼𝑗𝛽𝑗=0(1)𝛼𝑗𝛽𝑗𝛼𝑗𝛽𝑗𝛽𝑑𝑖=1𝛾𝑖,𝑗𝑗=0.(3.8) Thus, for all 0𝑚𝜶11, 𝑑𝑚𝜑(0)/𝑑𝑚=0.Case 2 (𝑚=𝜶1). In this case, if there exists some 1𝑗𝑑 such that 𝑑𝑖=1𝛾𝑖,𝑗>𝛼𝑗, then there must exist another 1𝑗𝑑, such that 𝑑𝑖=1𝛾𝑖,𝑗<𝛼𝑗. Hence, when 𝑚=𝜶1, 𝑑𝑗=1𝛼𝑗𝛽𝑗=0(1)𝛼𝑗𝛽𝑗𝛼𝑗𝛽𝑗𝛽𝑑𝑖=1𝛾𝑖,𝑗𝑗0(3.9) if and only if for all 1𝑗𝑑, 𝑑𝑖=1𝛾𝑖,𝑗=𝛼𝑗. Combining this with the fact: 𝛼𝑗𝛽𝑗=0(1)𝛼𝑗𝛽𝑗𝛼𝑗𝛽𝑗𝛽𝛼𝑗𝑗=𝛼𝑗!,1𝑗𝑑,(3.10) we can conclude that 𝑑𝜶1𝜑(0)𝑑𝜶1=𝜶1!𝛾1,1++𝛾𝑑,1=𝛼1𝛾1,𝑑++𝛾𝑑,𝑑=𝛼𝑑𝜌𝛾1,11,1𝜌𝛾1,𝑑𝑑,1𝜌𝛾𝑑,11,𝑑𝜌𝛾𝑑,𝑑𝑑,𝑑×𝛼1𝛾1,1,,𝛾𝑑,1𝛼𝑑𝛾1,𝑑,,𝛾𝑑,𝑑𝜕𝜶1𝑝(𝝃)𝜕𝑥𝜸111𝜕𝑥𝜸𝑑𝟏𝑑.(3.11) That is 𝑑𝜶1𝜑(0)𝑑𝜶1=𝜶1!𝐷𝜶𝝆𝑝(𝝃).(3.12) In sum, we have deduced that 𝑑𝑚𝜑(0)𝑑𝑚=0,0𝑚𝜶11;𝜶1!𝐷𝜶𝝆𝑝(𝝃),𝑚=𝜶1,(3.13) which implies that 𝜑=𝜶1𝐷𝜶𝝆𝑝(𝝃)+𝑂𝜶1+1.(3.14)

4. The Second Class of 𝐷-Invariant Polynomial Subspaces

Let 𝐚=(𝑎0,𝑎1,,𝑎𝑛) with 𝑛1 be an 𝑛+1-tuple of positive integers satisfying𝑎0=1,𝑎1>>𝑎𝑛2,(4.1) and let 𝐜1=𝑐1,0,𝑐1,1,,𝑐1,𝑛,,𝐜𝑑=𝑐𝑑,0,𝑐𝑑,1,,𝑐𝑑,𝑛𝔽𝑛+1(4.2) satisfying that 𝑐1,0,𝑐2,0,,𝑐𝑑,0 are not all zero. Then it is clear that𝜏𝜸1,,𝜸𝑑=𝑛𝑗=0𝑎𝑗𝑑𝑖=1𝛾𝑖,𝑗,(4.3) with 𝜸𝑖=(𝛾𝑖,0,𝛾𝑖,1,,𝛾𝑖,𝑛)𝑛+1 defines a map 𝜏(𝑛+1)𝑑.

Proposition 4.1. Let 𝐚=(𝑎0,𝑎1,,𝑎𝑛), 𝐜𝑖=(𝑐𝑖,0,𝑐𝑖,1,,𝑐𝑖,𝑛), 1𝑖𝑑, and the map 𝜏 be as above. Let 𝑞𝑛,𝑚, 𝑚=0,1,,𝑎1 be polynomials defined by 𝑞𝑛,𝑚=𝜏(𝜸1,,𝜸𝑑)=𝑚𝐜𝜸11𝐜𝜸𝑑𝑑𝜸1!𝜸𝑑!𝑥𝜸111𝑥𝜸𝑑1𝑑,(4.4) with 𝜸𝑖=(𝛾𝑖,0,𝛾𝑖,1,,𝛾𝑖,𝑛)𝑛+1, 𝐜𝜸𝑖𝑖=𝑛𝑗=0𝑐𝛾𝑖,𝑗𝑖,𝑗, and let 𝒬 be a polynomial subspace defined by 𝒬=span𝔽𝑞𝑛,𝑚0𝑚𝑎1.(4.5) Then, the following hold:(i)dim𝒬=𝑎1+1;(ii)𝒬 is a 𝐷-invariant polynomial subspace.

Proof. (i) We assume that there exist 𝑘0,𝑘1,,𝑘𝑎11,𝑘𝑎1𝔽 such that 𝑘0𝑞𝑛,0+𝑘1𝑞𝑛,1++𝑘𝑎11𝑞𝑛,𝑎11+𝑘𝑎1𝑞𝑛,𝑎1=0.(4.6) Since there exists some 1𝑖0𝑑 such that 𝑐𝑖0,00, then 𝑥𝑚𝑖0 must belong to the support of 𝑞𝑛,𝑚. Using this together with the definition of 𝜏, we conclude that the degree of 𝑞𝑛,𝑚 is 𝑚.
Specifically, 𝑥𝑎1𝑖0 belongs to the support of 𝑞𝑛,𝑎1, while for all 0𝑚𝑎11, 𝑥𝑎1𝑖0 cannot belong to the support of 𝑞𝑛,𝑚. Therefore 𝑘𝑎1=0, which implies that 𝑘0𝑞𝑛,0+𝑘1𝑞𝑛,1++𝑘𝑎11𝑞𝑛,𝑎11=0.(4.7)
Arguing for 𝑚=𝑎11,𝑎12,,1,0 as for 𝑚=𝑎1, we get 𝑘𝑎11=0,𝑘𝑎12=0,,𝑘1=0,𝑘0=0, successively. That is to say, the family of polynomials 𝑞𝑛,𝑚,0𝑚𝑎1 is 𝔽-linearly independent. So dim𝒬=𝑎1+1.
(ii) If 1𝑚<𝑎𝑛, then 𝜏(𝜸1,,𝜸𝑑)=𝑚 implies that 𝛾𝑖,𝑗=0 with 1𝑖𝑑,1𝑗𝑛. Thus, for all 1𝑚<𝑎𝑛: 𝑞𝑛,𝑚=𝛾1,0++𝛾𝑑,0=𝑚𝑐𝛾1,01,0𝑐𝛾𝑑,0𝑑,0𝛾1,0!𝛾𝑑,0!𝑥𝛾1,01𝑥𝛾𝑑,0𝑑=1𝑐𝑚!1,0𝑥1+𝑐2,0𝑥2++𝑐𝑑,0𝑥𝑑𝑚.(4.8) Consequently, 𝜕𝑞𝑛,𝑚𝜕𝑥𝑖=𝑐𝑖,0𝑞𝑛,𝑚1,1𝑚<𝑎𝑛.(4.9)
Next, let 𝑚 be an arbitrary integer satisfying 𝑎𝑛𝑚𝑎1, and 𝑠 the minimum integer between 1 and 𝑛 such that 𝑚𝑎𝑠0. We claim 𝜕𝑞𝑛,𝑚𝜕𝑥𝑖=𝑐𝑖,0𝑞𝑛,𝑚1+𝑛𝑗=𝑠𝑐𝑖,𝑗𝑞𝑛,𝑚𝑎𝑗,𝑎𝑛𝑚𝑎1.(4.10) This claim together with equality (4.9) immediately means that 𝒬 is 𝐷-invariant.
To prove our claim, we will use induction on the number 𝑛. When 𝑛=1, our claim can be easily verified. Now, assume that our claim is true for 𝑛1. To prove that it holds for 𝑛, let 𝑘 be the maximum nonnegative integer such that 𝑘𝑎𝑛𝑚, that is, 𝑘𝑎𝑛𝑚,(𝑘+1)𝑎𝑛>𝑚(4.11) and 𝑘 the maximum nonnegative integer such that 𝑘𝑎𝑛𝑚1. From the definition of 𝑞𝑛,𝑚, we obtain that 𝑞𝑛,𝑚=𝑘𝑙=0𝜏𝛾1,0,,𝛾1,𝑛1𝛾,,𝑑,0,,𝛾𝑑,𝑛1=𝑚𝑙𝑎𝑛𝑐𝛾1,01,0𝑐𝛾1,𝑛11,𝑛1𝑐𝛾𝑑,0𝑑,0𝑐𝛾𝑑,𝑛1𝑑,𝑛1𝛾1,0!𝛾1,𝑛1!𝛾𝑑,0!𝛾𝑑,𝑛1!×𝑥𝛾1,0++𝛾1,𝑛11𝑥𝛾𝑑,0++𝛾𝑑,𝑛1𝑑𝛾1,𝑛++𝛾𝑑,𝑛=𝑙𝑐𝛾1,𝑛1,𝑛𝑐𝛾𝑑,𝑛𝑑,𝑛𝑐1,𝑛!𝑐𝑑,𝑛!𝑥𝛾1,𝑛1𝑥𝛾𝑑,𝑛𝑑.(4.12) That is 𝑞𝑛,𝑚=𝑘𝑙=01𝑐𝑙!1,𝑛𝑥1++𝑐𝑑,𝑛𝑥𝑑𝑙𝑞𝑛1,𝑚𝑙𝑎𝑛,(4.13) which plays an important role in what follows. At this point, we have two cases to consider.
Case 1 (𝑎𝑛𝑚<𝑎𝑛1). In this case, 𝑠=𝑛 and 0𝑚𝑙𝑎𝑛<𝑎𝑛1 for all 0𝑙𝑘. Hence, 𝑞𝑛1,𝑚𝑙𝑎𝑛=1𝑚𝑙𝑎𝑛!𝑐1,0𝑥1++𝑐𝑑,0𝑥𝑑𝑚𝑙𝑎𝑛,0𝑙𝑘.(4.14) By (4.13), we get 𝑞𝑛,𝑚=𝑘𝑙=01𝑙!𝑚𝑙𝑎𝑛!𝑐1,𝑛𝑥1++𝑐𝑑,𝑛𝑥𝑑𝑙𝑐1,0𝑥1++𝑐𝑑,0𝑥𝑑𝑚𝑙𝑎𝑛.(4.15) So we obtain that 𝜕𝑞𝑛,𝑚𝜕𝑥𝑖=𝑐𝑘𝑖,0𝑙=01𝑙!𝑚1𝑙𝑎𝑛!𝑐1,𝑛𝑥1++𝑐𝑑,𝑛𝑥𝑑𝑙𝑐1,0𝑥1++𝑐𝑑,0𝑥𝑑𝑚1𝑙𝑎𝑛+𝑐𝑘𝑖,𝑛𝑙=11(𝑙1)!𝑚𝑙𝑎𝑛!𝑐1,𝑛𝑥1++𝑐𝑑,𝑛𝑥𝑑𝑙1𝑐1,0𝑥1++𝑐𝑑,0𝑥𝑑𝑚𝑙𝑎𝑛=𝑐𝑖,0𝑞𝑛,𝑚1+𝑐𝑖,𝑛𝑞𝑛,𝑚𝑎𝑛.(4.16)Case 2 (𝑎𝑛1𝑚𝑎1). In this case, 0𝑠𝑛1.
By (4.11) and (4.1), it follows that 𝑚𝑘𝑎𝑛𝑎𝑛1<0. Thus, there exists some 0𝑙0𝑘1 such that 𝑚𝑙0𝑎𝑛𝑎𝑛1𝑙0,𝑚0𝑎+1𝑛𝑎𝑛1<0.(4.17) As a result, we have that 0𝑚𝑙𝑎𝑛<𝑎𝑛1 for all 𝑙0+1𝑙𝑘, which implies that 𝑞𝑛1,𝑚𝑙𝑎𝑛=1𝑚𝑙𝑎𝑛!𝑐1,0𝑥1++𝑐𝑑,0𝑥𝑑𝑚𝑙𝑎𝑛,𝑙0+1𝑙𝑘.(4.18) Therefore, 𝜕𝑞𝑛1,𝑚𝑙𝑎𝑛𝜕𝑥𝑖=𝑐𝑖,0𝑞𝑛1,𝑚1𝑙𝑎𝑛,𝑙0+1𝑙𝑘;0,𝑘+1𝑙𝑘.(4.19)
For all 0𝑙𝑙0, we have 𝑎𝑛1𝑚𝑙𝑎𝑛𝑎1. Then our inductive hypothesis implies that 𝜕𝑞𝑛1,𝑚𝑙𝑎𝑛𝜕𝑥𝑖=𝑐𝑖,0𝑞𝑛1,𝑚1𝑙𝑎𝑛+𝑛1𝑗=𝑠𝑙𝑐𝑖,𝑗𝑞𝑛1,𝑚𝑙𝑎𝑛𝑎𝑗,0𝑙𝑙0,(4.20) where 𝑠𝑙 is the minimal integer between 1 and 𝑛1 such that 𝑚𝑙𝑎𝑛𝑎𝑠𝑙0. It should be noticed that 𝑠0=𝑠.

By (4.19) and (4.20), we can deduce that 𝜕𝑞𝑛,𝑚𝜕𝑥𝑖=𝑐𝑖,0𝑘𝑙=01𝑐𝑙!1,𝑛𝑥1++𝑐𝑑,𝑛𝑥𝑑𝑙𝑞𝑛1,𝑚1𝑙𝑎𝑛+𝑐𝑖,𝑛𝑘𝑙=11𝑐(𝑙1)!1,𝑛𝑥1++𝑐𝑑,𝑛𝑥𝑑𝑙1𝑞𝑛1,𝑚𝑙𝑎𝑛+𝑙0𝑙=01𝑐𝑙!1,𝑛𝑥2++𝑐𝑑,𝑛𝑥𝑑𝑙𝑛1𝑗=𝑠𝑙𝑐𝑖,𝑗𝑞𝑛1,𝑚𝑙𝑎𝑛𝑎𝑗=𝑐𝑖,0𝑞𝑛,𝑚1+𝑐𝑖,𝑛𝑞𝑛,𝑚𝑎𝑛+𝑙0𝑙=01𝑐𝑙!1,𝑛𝑥2++𝑐𝑑,𝑛𝑥𝑑𝑙𝑛1𝑗=𝑠𝑙𝑐𝑖,𝑗𝑞𝑛1,𝑚𝑙𝑎𝑛𝑎𝑗.(4.21)
It remains to show that the last row of the above equality and the right-side hand of (4.10) are equal. More precisely, for each 𝑠0𝑗𝑛1, let 𝑘𝑗 denote the maximum nonnegative integer satisfying 𝑘𝑗𝑎𝑛𝑚𝑎𝑗, that is, 𝑘𝑗𝑎𝑛𝑚𝑎𝑗,𝑘𝑗𝑎+1𝑛>𝑚𝑎𝑗.(4.22) From (4.17), we know 𝑘𝑛1=𝑙0.
Due to (4.13), we observe that 𝑞𝑛,𝑚𝑎𝑗=𝑘𝑗𝑙=01𝑐𝑙!2,𝑛𝑥2++𝑐𝑑,𝑛𝑥𝑑𝑙𝑞𝑛1,𝑚𝑎𝑗𝑙𝑎𝑛,𝑠0𝑗𝑛1.(4.23) Furthermore, 𝑞𝑛,𝑚𝑎𝑗=𝑙0𝑙=01𝑐𝑙!1,𝑛𝑥2++𝑐𝑑,𝑛𝑥𝑑𝑙𝑞𝑛1,𝑚𝑎𝑗𝑙𝑎𝑛,𝑠0𝑗𝑛1.(4.24) Here, we have used the fact that 𝑞𝑛1,𝑚𝑎𝑗𝑙𝑎𝑛=0 for all 𝑠0𝑗𝑛1, 𝑘𝑗+1𝑙𝑙0.
Now, recall that 𝑠𝑙,0𝑙𝑙0 is the minimal integer between 1 and 𝑛1 such that 𝑚𝑙𝑎𝑛𝑎𝑠𝑙0, then 𝑛1𝑗=𝑠0𝑐𝑖,𝑗𝑞𝑛1,𝑚𝑎𝑗𝑙𝑎𝑛=𝑛1𝑗=𝑠𝑙𝑐𝑖,𝑗𝑞𝑛1,𝑚𝑎𝑗𝑙𝑎𝑛,0𝑙𝑙0.(4.25)
According to (4.24), (4.25), and the fact 𝑠=𝑠0, we find that 𝑛1𝑗=𝑠𝑐𝑖,𝑗𝑞𝑛,𝑚𝑎𝑗=𝑛1𝑗=𝑠0𝑐𝑙𝑖,𝑗0𝑙=01𝑐𝑙!1,𝑛𝑥2++𝑐𝑑,𝑛𝑥𝑑𝑙𝑞𝑛1,𝑚𝑎𝑗𝑙𝑎𝑛=𝑙0𝑙=01𝑐𝑙!1,𝑛𝑥2++𝑐𝑑,𝑛𝑥𝑑𝑙𝑛1𝑗=𝑠𝑙𝑐𝑖,𝑗𝑞𝑛1,𝑚𝑎𝑗𝑙𝑎𝑛.(4.26) Consequently, 𝜕𝑞𝑛,𝑚𝜕𝑥𝑖=𝑐𝑖,0𝑞𝑛,𝑚1+𝑛𝑗=𝑠𝑐𝑖,𝑗𝑞𝑛,𝑚𝑎𝑗,𝑎𝑛1𝑚𝑎1.(4.27)

Proposition 4.2. Let 𝐚=(𝑎0,𝑎1,,𝑎𝑛), 𝐜𝑖=(𝑐𝑖,0,𝑐𝑖,1,,𝑐𝑖,𝑛) with 1𝑖𝑑, and let 𝑞𝑛,𝑚 with 0𝑚𝑎1 be as above, a nonzero number in 𝔽. Then for arbitrary 𝑝𝔽[𝐱] and 𝝃𝔽𝑑: 1𝑚!𝑚0𝑟𝑚(1)𝑚𝑟𝑚𝑟𝑝𝝃+𝑛𝑗=0𝑐1,𝑗(𝑟)𝑎𝑗,,𝑛𝑗=0𝑐𝑑,𝑗(𝑟)𝑎𝑗=𝑞𝑛,𝑚(𝐷)𝑝(𝝃)+𝑂().(4.28)

Proof. Applying Taylor Formulas, we obtain 𝑝𝝃+𝑛𝑗=0𝑐1,𝑗(𝑟)𝑎𝑗,,𝑛𝑗=0𝑐𝑑,𝑗(𝑟)𝑎𝑗=𝑘=0𝜸11𝜸++𝑑1=𝑘(𝑟)1𝑖𝑑,0𝑗𝑛𝑎𝑗𝛾𝑖,𝑗𝐜𝜸11𝐜𝜸𝑑𝑑𝜸1!𝜸𝑑!𝜕𝑘𝑝(𝝃)𝜕𝑥𝜸111𝜕𝑥𝜸𝑑1𝑑,(4.29) where 𝐜𝜸𝑖𝑖=𝑛𝑗=0𝑐𝛾𝑖,𝑗𝑖,𝑗, 𝜸𝑖=(𝛾𝑖,0,,𝛾𝑖,𝑛)𝑛+1,1𝑖𝑑.
By means of the map 𝜏 defined as in (4.3), the above equality can be rewritten as 𝑝𝝃+𝑛𝑗=0𝑐1,𝑗(𝑟)𝑎𝑗,,𝑛𝑗=0𝑐𝑑,𝑗(𝑟)𝑎𝑗=𝑚𝑙=0𝜏𝜸1,,𝜸𝑑=𝑙(𝑟)𝑙𝐜𝜸11𝐜𝜸𝑑𝑑𝜸1!𝜸𝑑!𝜕𝜸11++𝜸𝑑1𝑝(𝝃)𝜕𝑥𝜸111𝜕𝑥𝜸𝑑1𝑑+𝑂𝑚+1,(4.30) where 𝑂(𝑚+1) is a polynomial in .
Finally, by Lemma 2.1, we can conclude that 0𝑟𝑚(1)𝑚𝑟𝑚𝑟𝑝𝝃+𝑛𝑗=0𝑐1,𝑗(𝑟)𝑎𝑗,,𝑛𝑗=0𝑐𝑑,𝑗(𝑟)𝑎𝑗=𝑚!𝑚𝜏𝜸1,,𝜸𝑑=𝑚𝐜𝜸11𝐜𝜸𝑑𝑑𝜸1!𝜸𝑑!𝜕𝜸11++𝜸𝑑1𝑝(𝝃)𝜕𝑥𝜸111𝜕𝑥𝜸𝑑1𝑑+𝑂𝑚+1.(4.31)

5. Main Theorem

Let 𝝃(1),,𝝃(𝜇),𝝃(𝜇+1),,𝝃(𝜇+𝜈)𝔽𝑑 be distinct points. For each 1𝑘𝜇, assume that 𝔡(𝑘)𝑑 is a lower set, and 𝝆(𝑘)=(𝝆1(𝑘),,𝝆𝑑(𝑘))(𝔽𝑑)𝑑 satisfies the conditions of Section 3. Likewise, for each 1𝑙𝜈, assume that 𝐚(𝑙)=(𝑎0(𝑙),𝑎1(𝑙),,𝑎(𝑙)𝑛(𝑙)), 𝐜𝑖(𝑙)=(𝑐(𝑙)𝑖,0, 𝑐(𝑙)𝑖,1,,𝑐(𝑙)𝑖,𝑛(𝑙)), 1𝑖𝑑, and 𝑞(𝑙)𝑛(𝑙),𝑚, 0𝑚𝑎1(𝑙) satisfy the conditions of Section 4.

Theorem 5.1. With the notation above, let 𝑃 be an ideal projector with ran𝑃=span𝔽𝛿𝝃(𝑘)𝐷𝜶𝝆(𝑘),𝛿𝝃(𝜇+𝑙)𝑞(𝑙)𝑛(𝑙),𝑚(𝐷)𝜶𝔡(𝑘),1𝑘𝜇,0𝑚𝑎1(𝑙),1𝑙𝜈,(5.1) and let 𝑃 be a Lagrange projector with ran𝑃=span𝔽𝛿𝝃(𝑘)+𝑑𝑖=1𝛼𝑖𝝆𝑖(𝑘),𝛿𝝃(𝜇+𝑙)+𝜙(𝑙)(𝑚)𝜶𝔡(𝑘),1𝑘𝜇,0𝑚𝑎1(𝑙),1𝑙𝜈,(5.2) where 𝔽{0} and 𝜙(𝑙)(𝑚)=𝑛(𝑙)𝑗=0𝑐(𝑙)1,𝑗(𝑚)𝑎𝑗(𝑙),𝑛(𝑙)𝑗=0𝑐(𝑙)2,𝑗(𝑚)𝑎𝑗(𝑙),,𝑛(𝑙)𝑗=0𝑐(𝑙)𝑑,𝑗(𝑚)𝑎𝑗(𝑙).(5.3) Then, the following statements hold:(i)there exists a positive 𝜂𝔽 such thatran𝑃||||=ran𝑃,0<<𝜂;(5.4)(ii)𝑃 is the pointwise limit of 𝑃,0<||<𝜂, as tends to zero.

Proof. Firstly, one can easily verify that 𝛿𝝀=𝝃(𝑘)𝐷𝜶𝝆(𝑘),𝛿𝝃(𝜇+𝑙)𝑞(𝑙)𝑛(𝑙),𝑚(𝐷)𝜶𝔡(𝑘),1𝑘𝜇,0𝑚𝑎1(𝑙)𝔽[𝐱],1𝑙𝜈𝑠,𝝀=𝛿𝝃(𝑘)+𝑑𝑖=1𝛼𝑖𝝆𝑖(𝑘),𝛿𝝃(𝜇+𝑙)+𝜙(𝑙)(𝑚)𝜶𝔡(𝑘),1𝑘𝜇,0𝑚𝑎1(𝑙)𝔽[𝐱],1𝑙𝜈𝑠,(5.5) form 𝔽-bases for ran𝑃 and ran𝑃, respectively, where 𝑠=𝜇𝑘=1#𝔡(𝑘)+𝜈𝑘=1𝑎1(𝑘)+𝜈.(5.6) Provided that the entries of 𝝀 and 𝝀 are arranged in the same order, namely, for arbitrary fixed 1𝑘𝜇, 𝜶𝔡(𝑘), the corresponding entries of 𝝀 and 𝝀 are in the same position, the same as for arbitrary fixed 1𝑙𝜈, 0𝑚𝑎1(𝑙). We denote 𝝀=(𝜆1,,𝜆𝑠) and 𝝀=(𝜆,1,,𝜆,𝑠), respectively.
Let 𝐪=(𝑞1,𝑞2,,𝑞𝑠) be an 𝔽-basis for ran𝑃, For convenience, we introduce two 𝑠×𝑠 matrices: 𝝀𝑇𝜆𝐪=𝑖𝑞𝑗1𝑖,𝑗𝑠,𝝀𝑇𝜆𝐪=,𝑖𝑞𝑗1𝑖,𝑗𝑠(5.7) and for arbitrary 𝑓𝔽[𝐱], vectors: 𝝀𝑇𝜆𝑓=𝑖𝑓1𝑖𝑠,𝝀𝑇𝜆𝑓=,𝑖𝑓1𝑖𝑠.(5.8)
For arbitrary 𝑝𝔽[𝐱], we have the following facts according to equality (3.5) and (4.28). (1)For fixed 1𝑘𝜇 and 𝜶𝔡(𝑘), (𝐷𝜶𝝆(𝑘)𝑝)(𝝃(𝑘)) can be linearly expressed by 𝑝𝝃(𝑘)+𝑑𝑖=1𝛽𝑖𝝆𝑖(𝑘)𝜷𝔡(𝑘){𝑂()},(5.9) since 𝔡(𝑘) is lower, and moreover, the linear combination coefficient of each 𝑝(𝝃(𝑘)+𝑑𝑖=1𝛽𝑖𝝆𝑖(𝑘)) is independent of 𝑝𝔽[𝐱].(2) For fixed 1𝑙𝜈 and 0𝑚𝑎1(𝑙), (𝑞(𝑙)𝑛(𝑙),𝑚(𝐷)𝑝)(𝝃(𝜇+𝑙)) can be linearly expressed by 𝑝𝝃(𝜇+𝑙)+𝜙(𝑙)(𝑟)0𝑟𝑚{𝑂()}.(5.10) Also, the linear combination coefficient of each 𝑝(𝝃(𝜇+𝑙)+𝜙(𝑙)(𝑟)) is independent of 𝑝𝔽[𝐱].
In brief, we can conclude that there exists a nonsingular matrix 𝑇 such that 𝝀𝑇𝝀𝐪𝑇𝑓𝝀=𝑇𝑇𝐪𝝀𝑇𝑓=𝝀𝑇𝐪𝝀𝑇𝑓+𝐸𝝐,(5.11) where each entry of [𝐸𝝐] has the same order as . As a consequence, the linear systems: 𝝀𝑇𝐪𝝀𝐱=𝑇𝝀𝑓,𝑇𝐪𝐱=𝝀𝑇𝑓(5.12) are equivalent, namely, they have the same set of solutions.(i)From (5.11), it follows that each entry of matrix 𝝀𝑇𝐪 converges to its corresponding entry of matrix 𝝀𝑇𝐪 as tends to zero, which implies that lim0𝝀det𝑇𝐪𝝀=det𝑇𝐪.(5.13) Since det(𝝀𝑇𝐪)0, there exists 𝜂>0 such that 𝝀det𝑇𝐪||||0,0<<𝜂.(5.14) Notice that (5.11) directly leads to 𝝀rank(𝑇𝐪)=rank(𝝀𝑇𝐪): ran𝑃=span𝔽||||𝐪,0<<𝜂,(5.15) follows, that is, 𝐪 forms an 𝔽-basis for ran𝑃. Since 𝐪 is also an 𝔽-basis for ran𝑃, we have ran𝑃=ran𝑃||||,0<<𝜂.(5.16)(ii) Suppose that ̃𝐱 and 𝐱𝟎 be the unique solutions of nonsingular linear systems: 𝝀𝑇𝐪𝐱=𝝀𝑇𝝀𝑓,(5.17)𝑇𝐪𝐱=𝝀𝑇𝑓,(5.18) respectively, where 𝑓𝔽[𝐱] and 0<||<𝜂. It is easy to see that𝑃̃𝑓=𝐪𝐱,𝑃𝑓=𝐪𝐱𝟎.(5.19) Notice that, as 0, 𝑃 is the pointwise limit of 𝑃 if and only if 𝑃𝑓 is the coefficient-wise limit of 𝑃𝑓 for all 𝑓𝔽[𝐱]. Therefore, it is sufficient to show that for every 𝑓𝔽[𝐱], the solution vector of system (5.17) converges to the one of system (5.18) as tends to zero, namely, lim0̃𝐱=𝐱𝟎.(5.20) By (5.11), the linear system: 𝝀𝑇𝐪𝝀𝐱=𝑇𝑓(5.21) can be rewritten as 𝝀𝑇𝐪+𝐸𝝀𝐱=𝑇𝑓+𝝐.(5.22) Since system (5.21) is equivalent to system (5.17), ̃𝐱 is also the unique solution of it. Consequently, using the perturbation analysis of the sensitivity of linear systems (see, e.g., [16, page 80]), we have ̃𝐱𝐱𝟎𝝀𝑇𝐪1𝝐𝐸𝐱𝟎+𝑂2.(5.23) Since each component of vector 𝝐𝐸𝐱𝟎 has the same order as , it follows that lim0̃𝐱𝐱𝟎=0, or, equivalently, lim0̃𝐱=𝐱𝟎, which completes the proof of the theorem.

6. Conclusions

The main work of this paper is to discretize a class of ideal projectors with two special classes of 𝐷-invariant polynomial subspaces into a sequence of Lagrange projectors.

Acknowledgment

This work was supported by the National Natural Science Foundation of China nos. 11101185 and 11171133.