• Nem Talált Eredményt

Christoffel functions with power type weights

N/A
N/A
Protected

Academic year: 2022

Ossza meg "Christoffel functions with power type weights"

Copied!
59
0
0

Teljes szövegt

(1)

Christoffel functions with power type weights

Tivadar Danka Vilmos Totik December 20, 2018

Abstract

Precise asymptotics for Christoffel functions are established for power type weights on unions of Jordan curves and arcs. The asymp- totics involve the equilibrium measure of the support of the measure.

The result at the endpoints of arc components is obtained from the corresponding asymptotics for internal points with respect to a differ- ent power weight. On curve components the asymptotic formula is proved via a sharp form of Hilbert’s lemniscate theorem while taking polynomial inverse images. The situation is completely different on the arc components, where the local asymptotics is obtained via a dis- cretization of the equilibrium measure with respect to the zeros of an associated Bessel function. The proofs are potential theoretical, and fast decreasing polynomials play an essential role in them.

Contents

1 Introduction 2

2 Tools 7

2.1 Fast decreasing polynomials . . . 8 2.2 Polynomial inequalities . . . 9

3 The model cases 13

3.1 Measures on the real line . . . 13 3.2 Measures on the unit circle . . . 15

4 Lemniscates 18

4.1 The upper estimate . . . 19 4.2 The lower estimate . . . 20

AMS Classification: 26C05, 31A99, 41A10, 42C05. Key words: Christoffel functions, asymptotics, power type weights, Jordan curves and arcs, Bessel functions, fast decreasing polynomials, equilibrium measures, Green’s functions

Supported by ERC Advanced Grant No. 267055

Supported by NSF DMS-1265375

(2)

5 Smooth Jordan curves 22

5.1 The lower estimate . . . 24

5.2 The upper estimate . . . 29

6 Piecewise smooth Jordan curves 30 7 Arc components 30 7.1 Bessel functions and some local asymptotics . . . 31

7.2 The upper estimate in Theorem 1.1 for one arc . . . 32

7.2.1 Division based on the zeros of Bessel functions . . . . 33

7.2.2 Division based solely on the equilibrium measure . . . 34

7.2.3 Construction of the polynomialsCn . . . 35

7.2.4 Bounds forAn(z) for |z| ≤n−τ . . . 35

7.2.5 Bounds forBn(z) for |z| ≤n−τ . . . 38

7.2.6 The square integral of Cn for|z| ≤n−τ . . . 41

7.2.7 The estimate of Cn(z) for |z|> n−τ . . . 42

7.2.8 Completion of the upper estimate for a single arc . . . 45

7.3 The upper estimate for several components . . . 45

7.4 The lower estimate in Theorem 1.1 on Jordan arcs . . . 46

8 Proof of Theorem 1.1 49 8.1 Proof of Proposition 8.1 . . . 52

8.2 Proof of Proposition 8.2 . . . 53

9 Proof of Theorem 1.2 55

1 Introduction

Christoffel functions have been the subject of many papers, see e.g. [12], [13], [18], and the extended reference lists there. They are intimately con- nected with orthogonal polynomials, reproducing kernels, spectral properties of Jacobi matrices, convergence of orthogonal expansion and even to random matrices, see [5], [13] and [18] for their various connections and applications.

The possible applications are growing, for example recently a new domain recovery technique has been devised that use the asymptotic behavior of Christoffel functions, see [6]; and in the last 4-5 years several important methods for proving universality in random matrix theory were based on them, see [1], [8], [9] and [10]. The aim of the present paper is to complete, to a certain extent, the investigations concerning their asymptotic behavior on Jordan curves and arcs.

Let µ be a finite Borel measure on the plane such that its support is compact and consists of infinitely many points. The Christoffel functions

(3)

associated withµare defined as

λn(µ, z0) = inf

Pn(z0)=1

Z

|Pn|2dµ, (1.1)

where the infimum is taken for all polynomials of degree at mostnthat take the value 1 at z. If pk(z) = pk(µ, z) denote the orthonormal polynomials with respect toµ, i.e. Z

pnpmdµ=δn,m, thenλncan be expressed as

λ−1n (µ, z) = Xn

k=0

|pk(z)|2.

In other words,λ−1(µ, z) is the diagonal of the reproducing kernel Kn(z, w) =

Xn

k=0

pk(z)pk(w)

which makes it an essential tool in many problems. It is easy to see that, with this reproducing kernel, the infimum in (1.1) is attained (only) for

Pn(z) = Kn(z, z0) Kn(z0, z0), see e.g. [20, Theorem 3.1.3]).

The earliest asymptotics for Christoffel functions for measures on the unit circle or on [−1,1] go back to Szeg˝o, see [21, Th. I’, p. 461]. He gave their behavior outside the support of the measure, and for some special cases he also found their behavior at points of (−1,1). The first result for a Jordan arc (a circular arc) was given in [4]. By now the asymptotic behavior of Christoffel functions for measures defined on unions of Jordan curves and arcs Γ is well understood: under certain assumptions we have for points z∈Γ that are different from the endpoints of the arc components of Γ

n→∞lim nλn(µ, z0) = w(z0)

ωΓ(z0), (1.2)

where w is the density of µ with respect to the arc measure sΓ on Γ, and ωΓ is the density of the equilibrium measure (see below) with respect tosΓ. For the most general results see [22] and [24].

What is left, is to decide the asymptotic behavior at the endpoints of the arc components. It turns out that this problem is closely related to the asymptotic behavior away from the endpoints, but for measures of the form dµ(x) =|z−z0|αdsΓ(z),α >−1, and the aim of this paper is to find these

(4)

asymptotic behaviors. When µ is of the just specified form, then we shall show (for the exact formulation see the next section),

n→∞lim n1+αλn(µ, z0) = 1

(πωΓ(z0))α+12α+1Γα+ 1 2

Γα+ 3 2

(1.3)

whenz0 is not the endpoint of an arc component of Γ, while at an endpoint

n→∞lim n2α+2λn(µ, z0) = Γ(α+ 1)Γ(α+ 2) (πM(Γ, z0))2α+2 , whereM(Γ, z0) is the limit ofp

|z−z0Γ(z) asz→z0 along Γ.

This paper uses some basic notions and results from potential theory.

See [2], [3], [16] or [19] for all the concepts we use and for the basic theory.

In particular,νΓ will denote the equilibrium measure of the compact set Γ.

Since the asymptotics reflect the support of the measure, in all such questions a global condition, stating that the measure is not too small on any part of Γ, is needed (for example, ifµis zero on any arc of Γ, then (1.3) does not hold any more). This global condition is the regularity condition from [19]: we say that µ, with support Γ, belongs to theReg class if

sup

Pn

kPnkΓ

kPnkL2(µ)

!1/n

→1

as n → ∞, where the supremum is taken for all polynomials of degree at mostn, and wherekPnkΓ denotes the supremum norm on Γ. The condition says that in then-th root sense theL(µ) andL2(µ)-norms are almost the same. The assumption µ ∈ Reg is a very weak condition – see [19] for several reformulations as well as conditions on the measure µ that implies µ ∈ Reg. For example, if Γ consists of rectifiable Jordan curves and arcs with arc measuresΓ, then any measure dµ(z) =w(z)dsΓ(z) withw(z) >0 sΓ-almost everywhere is regular in this sense.

Actually, it is not even needed that the support Γ of the measure µ be a system of Jordan curves or arcs, the main theorem below holds for any Γ that is a finite union of continua (connected compact sets). However, it is needed thatz0 lies on a smooth arc J of the outer boundary of Γ: the outer boundary of Γ is the boundary of the unbounded connected component of C\Γ. It is known that the equilibrium measure νΓ lives on the outer boundary, and ifJ is a smooth (sayC1-smooth) arc on the outer boundary, then on J the equilibrium measure is absolutely continuous with respect to the arc measure sJ on J: dνΓ(z) = ωΓ(z)dsJ(z). We call this ωΓ the equilibrium density of Γ.

The following theorem describes the asymptotics of the Christoffel func- tion at points that are different from the endpoints of the arc-components/parts of Γ, see Figure 1 for illustration.

(5)

G

z0

Figure 1: A typical position where Theorem 1.1 can be applied Theorem 1.1 Let the support Γ of a measure µ ∈ Reg consist of finitely many continua, and let z0 lie on the outer boundary of Γ. Assume that the intersection ofΓwith a neighborhood of z0 is a C2-smooth arcJ which con- tainsz0in its (one-dimensional) interior. Assume also that in this neighbor- hooddµ(z) =w(z)|z−z0|αdsJ(z), where w is a strictly positive continuous function and α >−1. Then

n→∞lim n1+αλn(µ, z0) = w(z0)

(πωΓ(z0))α+12α+1Γα+ 1 2

Γα+ 3 2

. (1.4) The second main theorem of this work is about the behavior of the Christoffel function at an endpoint, see Figure 2. Ifz0 is an endpoint of a smooth arcJ on the outer boundary of Γ, then atz0 the equilibrium density has a 1/p

|z−z0|behavior (see the proof of Theorem 1.2), and we set M(Γ, z0) := lim

z→z0, z∈Γ

p|z−z0Γ(z). (1.5)

Theorem 1.2 Let Γ and µbe as in Theorem 1.1, but now assume that the intersection of Γ with a neighborhood of z0 is a C2-smooth Jordan arc J with one endpoint atz0. Then

n→∞lim n2α+2λn(µ, z0) = w(z0)

(πM(Γ, z0))2α+2Γ(α+ 1)Γ(α+ 2). (1.6) These results can be used, in particular, if the measure is supported on a finite union of intervals on the real line, in which case the quantitiesωΓ(x) and M(Γ, x) have a rather explicit form. Let Γ = ∪kj=00 [a2j, a2j+1] with disjoint [a2j, a2j+1]. Then the equilibrium density of Γ is (see e.g. [23, (40),

(6)

G

z0

Figure 2: A typical position where Theorem 1.2 can be applied (41)] or [19, Lemma 4.4.1])

ωΓ(x) =

Qk0−1

j=0 |x−λj| πqQ2k0+1

j=0 |x−aj|

, x∈Int(Γ), (1.7)

whereλj are the solutions of the system of equations Z a2k+2

a2k+1

Qk0−1

j=0 (t−λj) qQ2k0+1

j=0 |t−aj|

dt= 0, k= 0, . . . k0−1. (1.8)

It can be easily shown that these λj’s are uniquely determined and there is one λj on every contiguous interval (a2j+1, a2j+2). Now if a is one of the endpoints of the intervals of Γ, saya=aj0, then

M(Γ, a) =

Qk0−1

j=0 |a−λj| πqQ2k0

j=1, j6=j0|a−aj|

. (1.9)

This whole work is dedicated to proving Theorem 1.1 and Theorem 1.2.

Actually, the latter will be a relatively easy consequence of the former one, so the main emphasis will be to prove Theorem 1.1. The main line of reasoning will be the following. We start from some known facts for simple measures like |x|αdx on the real line, and get some elementary results for a model case on the unit circle via a transformation. Then we prove from these simple cases that Theorem 1.1 is true for lemniscate sets, i.e. level sets of polynomials. This part will use the polynomial mapping in question to transform the already known result to the given lemniscate. Then we prove

(7)

the theorem for finite unions of Jordan curves. Recall that a Jordan curve is a homeomorhic image of a circle, while a Jordan arc is a homeomorhic image of a segment. From the point of view of finding the asymptotics of Christoffel functions there is a big difference between arcs and curves: Jordan curves have interior and can be exhausted by lemniscates, so the polynomial inverse image method of [23] is applicable for them, while for Jordan arcs that method cannot be applied. Still, the pure Jordan curve case is used when we go over to a Γ which may have arc components, namely it is used in the lower estimate. The upper estimate is the most difficult part of the proof; there Bessel functions enter the picture, and a discretization technique is developed where the discretization of the equilibrium measure of Γ is done using the zeros of appropriate Bessel functions combined with another discretization based on uniform distribution. Once the case of Jordan curves and arcs have been settled, the proof of Theorem 1.1 will easily follow by approximating a general Γ by a family of Jordan curves and arcs.

2 Tools

In what follows,k · kK denotes the supremum norm on a setK, and sΓ the arc measure on Γ (when Γ consists of smooth Jordan arcs or curves).

We shall rely on some basic notions and facts from logarithmic potential theory. See the books [2], [3], [16] or [17] for detailed discussion.

We shall often use the trivial fact that if µ, ν are two Borel measures, then µ ≤ ν implies λn(µ, x) ≤ λn(ν, x) for all x. It is also trivial that λn(µ, z)≤µ(C) (just use the identically 1 polynomial as a test function in the definition ofλn(µ, z)).

Another frequently used fact is the following: if{nk}is a subsequence of the natural numbers such that nk+1/nk →1 ask→ ∞, then for anyκ >0

lim inf

n→∞ nκλn(µ, x) = lim inf

k→∞ nκkλnk(µ, x), (2.1) and

lim sup

n→∞ nκλn(µ, x) = lim sup

k→∞

nκkλnk(µ, x). (2.2) In fact, sinceλn(µ, x) is a monotone decreasing function of n, fornk≤n≤ nk+1 we have

n nk+1

κ

nκk+1λnk+1(µ, x)≤nκλn(µ, x)≤ n

nk κ

nκkλnk(µ, x), and both claims follow because n/nk and n/nk+1 tend to 1 as n (or nk) tends to infinity.

(8)

2.1 Fast decreasing polynomials

The following lemmas on the existence of fast decreasing polynomials will be a constant tool in the proofs.

Proposition 2.1 Let K be a compact subset on C, Ω the unbounded com- plement of C\K and let z0 ∈ ∂Ω. Suppose that there is a disk in Ω that contains z0 on its boundary. Then, for every γ > 1, there are constants cγ, Cγ, and for every n ∈ N polynomials Sn,z0,K of degree at most n such thatSn,z0,K(z0) = 1, |Sn,z0,K(z)| ≤1 for all z∈K and

|Sn,z0,K(z)| ≤Cγe−ncγ|z−z0|γ, z∈K. (2.3) For details, see [22, Theorem 4.1]. This theorem will often be used in the following form.

Corollary 2.2 With the assumptions of Proposition 2.1 for every 0< τ <

1, there exists constants cτ, Cτ, τ0 > 0 and for every n ∈ N a polynomial Sn,z0,K of degree o(n) such that Sn,z0,K(z0) = 1, |Sn,z0,K(z)| ≤ 1 for all z∈K, and

|Sn,z0,K(z)| ≤Cτe−cτnτ0, |z−z0| ≥n−τ. (2.4) Proof. Let 0< εbe sufficiently small and select γ >1 so that 1−ε−τ γ >

0. Lemma 2.1 tells us that there is a polynomial Pn with deg(Pn) ≤n1−ε such that

|Pn(z)| ≤Cγe−cγn1−(ε+τ γ), |z−z0| ≥n−τ, and this proves the claim withSn,z0,K =Pn.

There is a version of Lemma 2.1 where the decrease is not exponentially small, but starts much earlier than in Lemma 2.1.

Proposition 2.3 Let K be as in Proposition 2.1. Then, for every β < 1, there are constants cβ, Cβ >0, and for everyn= 1,2, . . . polynomials Pn of degree at most n such thatPn(z0) = 1,|Pn(z)| ≤1 for z∈K and

|Pn(z)| ≤Cβe−cβ(n|z−z0|)β, z∈K. (2.5) See [25, Lemma 4].

It will be convenient to use these results when n > 1 is not necessarily integer (formally one has to take the integral part of n, but the estimates will hold with possibly smaller constants in the exponents).

(9)

2.2 Polynomial inequalities

We shall also need some inequalities for polynomials that are used several times in the rest of the paper.

We start with a Bernstein-type inequality.

Lemma 2.4 Let J be a C2 closed Jordan arc and J1 a closed subarc of J not having common endpoint with J. Then, for every D > 0, there is a constant CD, such that

|Pn(z)| ≤CDnkPnkJ, dist(z, J1)≤D/n, holds for any polynomialsPn of degree n= 1,2, . . ..

See [22, Corollary 7.4].

Next, we continue with a Markov-type inequality.

Lemma 2.5 LetK be a continuum. IfQnis a polynomial of degree at most n= 1,2, . . ., then

kQnkK ≤ e

2cap(K)n2kQnkK, (2.6) where cap(K) denotes the logarithmic capacity of K.

In particular, if K has diameter 1, then

kQnkK ≤2en2kQnkK. (2.7) For (2.6) see [15, Theorem 1], and for the last statement note that ifK has diameter 1, then its capacity is at least 1/4 ([16, Theorem 5.3.2(a)]).

Next, we prove a Remez-type inequality.

Lemma 2.6 Let Γbe a C1 Jordan curve or arc, and assume that for every n= 1,2, . . ., Jn is a subarc of Γ, and Jn is a subset of Jn such that

sΓ(Jn\Jn) =o(n−2)sΓ(Jn),

where sΓ denotes the arc-length measure on Γ. Then, for any sequence {Qn}of polynomials of degree at most n= 1,2, . . ., we have

kQnkJn = (1 +o(1))kQnkJn. (2.8) Proof. It is clear from theC1 property that sΓ(Jn)∼diam(Jn) uniformly inJn (meaning that the ratio of the two sides lies in between two positive constants).

Make a linear transformation z → Cz such that, after this transfor- mation, the arc ˜Jn that we obtain from Jn has diameter 1. Under this transformationJn goes into a subset ˜Jn of ˜Jn for which

sJ˜

n( ˜Jn\J˜n) =o(n−2)sJ˜

n( ˜Jn), (2.9)

(10)

and Qn changes into a polynomial ˜Qn of degree at most n. (2.8) is clearly equivalent to its˜-version.

Let M =kQ˜nkJ˜n. By Lemma 2.5, the absolute value of ˜Qn is bounded on ˜Jnby 2en2M, hence ifz, w∈J˜n, then

|Q˜n(z)−Q˜n(w)| ≤2en2M sJ˜

n(zw), (2.10)

where zw is the arc of ˜Jn lying in between z and w. By the assumption (2.9) for everyz∈J˜nthere is aw∈J˜n with

sJ˜

n(zw) =o(n−2)sJ˜

n( ˜Jn) =o(n−2) because sJ˜

n( ˜Jn)∼diam( ˜Jn) = 1. Choose herez ∈J˜n such that |Q˜n(z)|= M. Since |Q˜n(w)| ≤ kQ˜nkJ˜n, we get from (2.10)

M =|Q˜n(z)| ≤ kQ˜nkJ˜n+o(1)M, and the claim follows.

We shall frequently use the following, so called Nikolskii-type inequalities for power type weights. In it we write that a Jordan arc isC1+-smooth if there is aθ >0 such that the arc in question isC1+θ-smooth.

Lemma 2.7 Let J be a C1+-smooth Jordan arc and letJ⊂J be a subarc of J which has no common endpoint with J. Let z0 ∈ J be a fixed point, and for α > −1 define the measure να on J by dνα(u) = |u−z0|αdsJ(u).

Then there is a constantC depending only onα, J andJ such that for any polynomials Pn of degree at most n= 1,2, . . . we have

kPnkJ ≤Cn(1+α)/2kPnkL2α), (2.11) if α≥0, and

kPnkJ ≤Cn1/2kPnkL2α), (2.12) if −1< α <0.

The same is true if dνα(u) = w(u)|u−z0|αdsJ(u) with some strictly positive and continuous w.

Proof. In view of [26, Lemmas 3.8 and Corollary 3.9] (use also that να is a doubling weight in the sense of [26]) uniformly inz∈J we have for large nthe relation

λnα, z)∼να(l1/n(z)),

(11)

whereA∼B means that the ratio lies in between two constants, and where l1/n(z) is the arc of J consisting of those points of z that lie of distance

≤1/n fromz. Ifα≥0, then

να(l1/n(z))≥ c n1+α, while for −1< α <0

να(l1/n(z))≥ c n,

with some positive constantcwhich depends only onα, JandJ. Therefore, we have for allz∈J the inequality

λnα, z)≥ c

n1+α (2.13)

ifα≥0 and

λnα, z)≥ c

n (2.14)

when−1< α <0.

For example, (2.13) means that ifα≥0 and|Pn(z)|= 1 for somez∈J, then necessarily

n1+α c

Z

J|Pn|2α≥1,

which is equivalent to saying that for anyPn andz∈J n1+α

c Z

J|Pn|2α≥ |Pn(z)|2,

and this is (2.11). In a similar manner, (2.12) follows from (2.14).

It is clear that this proof does not change if να is as in the last sentence of the lemma.

Lemma 2.8 If α > −1, then there is a constant Cα such that for any polynomialPn of degree at most n the inequality

kPnk[−1,1]≤Cαn(1+α)/2 Z 1

−1|Pn(x)|2|x|αdx 1/2

(2.15) holds withα= max(1, α).

Proof. We follow the preceding proof, but now bothJ and J agree with [−1,1].

LetJ =J = [−1,1],z0 = 0, ∆n(z) = 1/n2 ifz∈[−1,−1 + 1/n2] orz∈ [1−1/n2,1], and set ∆n(z) =√

1−z2/nifz∈[−1 + 1/n2,1−1/n2]. If now l1/n(z) is the interval [z−∆n(z), z+∆n(z)] intersected with [−1,1], then [26,

(12)

Lemmas 3.8 and Corollary 3.9] state that fordνα(x) =|x−z0|αdx=|x|αdx on [−1,1] we have

λnα, z)∼να(l1/n(z)).

Ifα≥0, then

να(l1/n(z))≥cmin 1

n2, 1 n1+α

, while for −1< α <0

να(l1/n(z))≥ c n2, with some positive constant c. Hence,

λnα, z)≥ c n2 if−1< α≤1, while

λnα, z)≥ c n1+α ifα≥1, from which (2.15) follows exactly as before.

The Nikolskii inequalties can be combined with the following estimate to get an upper bound for the extremal polynomials that produceλn(µ, z).

Lemma 2.9 With the assumptions of Theorem 1.1 we have λn(µ, z0)≤Cn−(α+1)

with some constantC that is independent of n.

Proof. Just use the polynomials Pn from Proposition 2.3 with β = 1/2 and K = Γ. Let δ > 0 be so small that in the δ-neighborhood of z0 we have the dµ(z) = w(z)|z−z0|αdsΓ(z) representation for µ. Outside this δ-neighborhood|Sn,z0|is smaller thanCβexp(−cβ(nδ)1/2), so

Z

|Sn,z0|2dµ≤C Z

e−2cβ(n|t|)1/2|t|αdt+Ce−2cβ(nδ)1/2 ≤Cn−α−1, which proves the claim.

We close this section with the classical Bernstein-Walsh lemma, see [27, p. 77].

Lemma 2.10 Let K⊂Cbe a compact subset of positive logarithmic capac- ity, letΩbe the unbounded component ofC\K, and g the Green’s function of this unbounded component with pole at infinity. Then, for polynomialsPn of degree at most n= 1,2, . . ., we have for any z∈C

|Pn(z)| ≤eng(z)kPnkK.

(13)

3 The model cases

3.1 Measures on the real line

Our first goal is to establish asymptotics for the Christoffel function at 0 with respect to the measure dµ(x) = |x|αdx, x ∈ [−1,1]. We do this by transforming some previously known results.

In what follows, for simpler notations, if dµ(x) =w(x)dx, then we shall write λn(w(x), z) for λn(µ, z).

Proposition 3.1 Forα >−1 we have

n→∞lim n2α+2λn

|x|α

[0,1],0

= Γ(α+ 1)Γ(α+ 2). (3.1) Proof. It follows from [10, (1.10)] or [9, Theorem 4.1] that

n→∞lim n2α+2λn

(1−x)α

[−1,1],1

= 2α+1Γ(α+ 1)Γ(α+ 2), (3.2) from which the claim is an immediate consequence if we apply the linear transformationx→(1−x)/2.

Proposition 3.2 Forα >−1 we have

n→∞lim nα+1λn

|x|α

[−1,1],0

=Lα, (3.3)

where

Lα:= 2α+1Γα+ 1 2

Γα+ 3 2

. (3.4)

Proof. Let us agree that in this proof, whenever we write Pn, Rn etc. for polynomials, then it is understood that the degree is at mostn.

We use that (for continuousf) Z 1

0

f(x)|x|αdx= Z 1

−1

f(x2)|x|2α+1dx. (3.5) Assume first thatP2nis extremal forλ2n

|x|α

[−1,1],0

, i.e. P2n(0) = 1

and Z 1

−1|P2n(x)|2|x|αdx=λ2n

|x|α

[−1,1],0

. Define

R2n(x) = P2n(x) +P2n(−x)

2 .

(14)

Then R2n(0) = 1, and R2n is a polynomial in x2, hence R2n(x) = Rn(x2) with some polynomialRn, for which Rn(0) = 1 and deg(Rn)≤n. Now we have

Z 1

−1|R2n(x)|2|x|αdx= Z 1

−1|Rn(x2)|2|x|αdx = Z 1

0 |Rn(x)|2|x|α−12 dx

≥ λn

|x|α−12

[0,1],0

. With the Cauchy-Schwarz inequality and with the symmetry of the measure

|x|αdx, we have Z 1

−1|R2n(x)|2|x|αdx≤ 1 4

Z 1

−1

|P2n(x)|2+ 2|P2n(x)||P2n(−x)|+|P2n(−x)|2

|x|αdx

≤ 1 2

Z 1

−1|P2n(x)|2|x|αdx +1

2 Z 1

−1|P2n(x)|2|x|αdx

!1/2 Z 1

−1|P2n(−x)|2|x|αdx

!1/2

= Z 1

−1|P2n(x)|2|x|αdx=λ2n

|x|α

[−1,1],0

. Combining these two estimates, we obtain

λn

|x|α−12

[0,1],0

≤λ2n

|x|α

[−1,1],0

. On the other hand, if now Pn is extremal for λn

|x|α−12

[0,1],0

, then λn

|x|α−12

[0,1],0

= Z 1

0 |Pn(x)|2|x|α−12 dx = Z 1

−1|Pn(x2)|2|x|αdx

≥ λ2n

|x|α

[−1,1],0

, therefore we actually have the equality

λn

|x|α−12

[0,1],0

2n

|x|α

[−1,1],0

, (3.6)

from which the claim follows via Proposition 3.1 (see also (2.1) and (2.2) withnk= 2k).

Note also that this proves also that if Pn(x) is the n-degree extremal polynomials for the measure |x|α−12

[0,1], then Pn(x2) is the 2n-degree ex- tremal polynomial for the measure|x|α

[−1,1].

(15)

3.2 Measures on the unit circle

LetµT be the measure on the unit circleT defined bydµT(eit) =wT(eit)dt, where

wT(eit) = |e2it+ 1|α 2α

|e2it−1|

2 , t∈[−π, π). (3.7) We shall prove

n→∞lim nα+1λnT, eiπ/2) = 2α+1Lα (3.8) where Lα is from (3.4), by transforming the measure µT into a measure µ[−1,1]supported on the interval [−1,1] and comparing the Christoffel func- tions for them. With the transformationeit→cost, we have

Z π

−π

f(cost)wT(eit)dt= 2 Z 1

−1

f(x)w[−1,1](x)dx, where

w[−1,1](x) =|x|α. Setdµ[−1,1](x) =w[−1,1](x)dx.

LetPn be the extremal polynomial for λn[−1,1],0) and define Sn(eit) =Pn(cost) 1 +ei(t−π/2)

2

!⌊ηn⌋

ein(t−π/2),

whereη >0 is arbitrary. This Sn is a polynomial of degree 2n+⌊ηn⌋ with Sn(eiπ/2) = 1. For any fixed 0< δ <1

Z π/2+δ

π/2−δ |Sn(eit)|2wT(eit)dt≤

Z π/2+δ

π/2−δ |Pn(cost)|2wT(eit)dt

≤ Z 1

−1|Pn(x)|2w[−1,1](x)dx

n[−1,1],0).

(3.9)

To estimate the corresponding integral over the intervals [−π, π/2−δ] and [π/2 +δ, π], notice that

t∈[−π,π]\[π/2−δ,π/2+δ]max

1 +ei(t−π/2) 2

⌊ηn⌋

=O(qn) (3.10) for someq <1. From Lemma 2.8 we obtain

kPnk[−1,1]≤Cn1+|α|/2kPnkL2[−1,1])≤Cn1+|α|/2, and so

Z π/2−δ

−π

+ Z π

π/2+δ

!

|Sn(eit)|2wT(eit)dt=O(n1+|α|/2qn) =o(n−α−1).

(16)

Therefore, using thisSn as a test polynomial forλdeg(Sn)T, eiπ/2) we con- clude

λdeg(Sn)T, eiπ/2)≤λn[−1,1],0) +o(n−α−1), and so

lim sup

n→∞ 2n+⌊ηn⌋α+1

λ2n+⌊ηn⌋T, eiπ/2)≤lim sup

n→∞ 2 +⌊ηn⌋/nα+1

nα+1λn[−1,1],0)

= (2 +η)α+1Lα, where we used Proposition 3.2 for the measureµ[−1,1].

Since η >0 was arbitrary, lim sup

n→∞

nα+1λnT, eiπ/2)≤2α+1Lα (3.11) follows (see also (2.2)).

Now to prove the matching lower estimate, let S2n(eit) be the extremal polynomial forλ2nT, eiπ/2). Define

Pn(eit) =S2n(eit) 1 +ei(t−π/2) 2

!2⌊ηn⌋

e−(n+⌊ηn⌋)i(t−π/2)

and Pn(cost) =Pn(eit) +Pn(e−it). Note that Pn(cost) is a polynomial in cost of deg(Pn)≤n+⌊ηn⌋ and Pn(0) = 1. With it we have

λdeg(Pn)[−1,1],0)≤ Z 1

−1|Pn(x)|2w[−1,1](x)dx= 1 2

Z π

−π|Pn(cost)|2wT(eit)dt.

(3.12) First, we claim that for every fixed 0< δ <1

|Pn(cost)|2=|Pn(eit)|2+O(qn), t∈[π/2−δ, π/2 +δ],

|Pn(cost)|2=|Pn(e−it)|2+O(qn), t∈[−π/2−δ,−π/2 +δ],

|Pn(cost)|2=O(qn) otherwise,

(3.13)

hold for someq <1. Indeed,

|Pn(cost)|2 =|Pn(eit)+Pn(e−it)|2 ≤ |Pn(eit)|2+2|Pn(eit)||Pn(e−it)|+|Pn(e−it)|2. If we apply Lemma 2.7 to two subarcs (say of length 5π/4) ofTthat contain the upper, resp. the lower half of the unit circle, then we obtain that

kPnkT≤ kS2nkT ≤Cn(1+|α|)/2kS2nkL2T)≤Cn(1+|α|)/2. Therefore (use (3.10))

|Pn(eit)| ≤Cqnn(1+|α|)/2, t∈[−π, π]\[π/2−δ, π/2 +δ].

(17)

These imply (3.13).

Now we have Z π

−π|Pn(cost)|2wT(eit)dt=

Z π/2+δ π/2−δ

+

Z −π/2+δ

−π/2−δ

!

|Pn(cost)|2wT(eit)dt +

Z −π/2−δ

−π

+

Z π/2−δ

−π/2+δ

+ Z π

π/2+δ

!

|Pn(cost)|2wT(eit)dt.

(3.13) tells us that the last three terms areO(qn). For the other two terms we have, again by (3.13),

Z π/2+δ

π/2−δ |Pn(cost)|2wT(eit)dt=

Z π/2+δ

π/2−δ |Pn(eit)|2wT(eit)dt+O(qn)

Z π/2+δ

π/2−δ |S2n(eit)|2wT(eit)dt+O(qn)

≤λ2nT, eiπ/2) +O(qn) and similarly,

Z −π/2+δ

−π/2−δ |Pn(cost)|2wT(eit)dt≤λ2nT, eiπ/2) +O(qn).

Combining these estimates with (3.12), we can conclude λdeg(Pn)[−1,1],0)≤λ2nT, eiπ/2) +O(qn), therefore

lim inf

n→∞ deg(Pn)α+1λdeg(Pn)[−1,1],0)≤lim inf

n→∞ (n+⌊ηn⌋)α+1 λ2nT, eiπ/2) +O(qn)

≤lim inf

n→∞ (1 +⌊ηn⌋/n)α+1 1

2α+1(2n)α+1λ2nT, eiπ/2).

From this, in view of Proposition 3.2 and (2.1), it follows that (1 +η)−(α+1)2α+1Lα≤lim inf

n→∞ λnT, eiπ/2), and upon lettingη→0 we obtain

2α+1Lα ≤lim inf

n→∞ λnT, eiπ/2). (3.14) This and (3.11) verify (3.8).

Finally, let

α(eit) =|eit−i|αdt.

(18)

Let us write|eit−i|α in the form

|eit−i|α=w(eit)wT(eit).

Thenwis continuous in a neighborhood ofeiπ/2 and it has value 1 ateiπ/2. Letτ >0 be arbitrary, and choose 0< δ <1 in such a way that

1

1 +τ ≤w(eit)≤(1 +τ), t∈[π/2−δ, π/2 +δ].

If we now carry out the preceding arguments with this δ and with this µα replacing everywhere µT, then we get that in (3.11) the limsup is at most (1 +τ)2α+1Lα, while in (3.14) the liminf is at least (1 +τ)−12α+1Lα. Since τ >0 can be arbitrarily chosen, this shows that

n→∞lim nα+1λnα, eiπ/2) = 2α+1Lα. (3.15) This result will serve as our model case in the proof of Theorem 1.1.

4 Lemniscates

In this section, we prove Theorem 1.1 for lemniscates.

Letσ ={z∈C:|TN(z)|= 1} be the level line of a polynomial TN, and assume thatσ has no self-intersections. Let deg(TN) =N.

The normal derivative of the Green’s function with pole at infinity of the outer domain to σ at a point z ∈ σ is (see [22, (2.2)]) |TN (z)|/N, and since this normal derivative is 2π-times the equilibrium density of σ (see [14, II.(4.1)] or [17, Theorem IV.2.3] and [17, (I.4.8)]), it follows that the equilibrium density onσ has the form

ωσ(z) = |TN (z)|

2πN . (4.1)

Ifz∈σ, then there arenpointsz1, . . . , zn∈σwith the propertyTN(z) = Tn(zk), and for them (see [22, (2.12)])

Z

σ

XN

i=1

f(zi)

|TN (z)|dsσ(z) =N Z

σ

f(z)|TN (z)|dsσ(z). (4.2) Furthermore, ifg:T→Cis arbitrary, then (see [22, (2.14)])

Z

σ

g(TN(z))|TN (z)|dsσ(z) =N Z

0

g(eit)dt. (4.3) Letz0 ∈σ be arbitrary, and define the measure

σ(z) =|z−z0|αdsσ(z), α >−1, (4.4)

(19)

wheresσ denotes the arc measure onσ. Without loss of generality we may assume thatTN(z0) =eiπ/2. Our plan is to compare the Christoffel functions for the measureµσ with that for the measureµα which is supported on the unit circle and is defined via

α(eit) =|eit−eiπ/2|αdsT(eit), (4.5) and for which the asymptotics of the Christoffel function was calculated in (3.15).

We shall prove that

n→∞lim nα+1λnσ, z0) = Lα

(πωσ(z0))α+1 (4.6) whereLα is taken from (3.4).

4.1 The upper estimate

Let η > 0 be an arbitrary small number, and select a δ > 0 such that for everyz with|z−z0|< δ, we have

1

1 +η|TN (z0)| ≤ |TN (z)| ≤(1 +η)|TN (z0)| 1

1 +η|TN (z0)||z−z0| ≤ |TN(z)−TN(z0)| ≤(1 +η)|TN (z0)||z−z0| (4.7)

(note that TN (z0) 6= 0 because σ has no self-intersections). Let Qn be the extremal polynomial forλnα, eiπ/2), whereµα is from (4.5). DefineRn as

Rn(z) =Qn(TN(z))Sn,z0,L(z),

where Sn,z0,L is the fast decreasing polynomial given by Corollary 2.2 for the lemniscate setLenclosed byσ (and for any fixed 0< τ <1 in Corollary 2.2). Note that Rn is a polynomial of degree nN +o(n) with Rn(z0) = 1.

SinceSn,z0,L is fast decreasing, we have sup

z∈L\{t:|t−z0|<δ}|Sn,z0,L(z)|=O(qnτ0)

for some q < 1 and τ0 > 0. The Nikolskii-type inequality in Lemma 2.7 when applied to two subarcs ofTwhich contain the upper resp. lower part of the unit circle, yields

kQnkT ≤Cn(1+|α|)/2kQnkL2α) ≤Cn(1+|α|)/2. Therefore,

sup

z∈L\{t:|t−z0|<δ}|Rn(z)|=O(qnτ0/2).

(20)

It follows that Z

|z−z0|≥δ|Rn(z)|2|z−z0|αdsσ(z) =O(qnτ0/2). (4.8) Using (4.7), we have

Z

|z−z0|<δ|Rn(z)|2|z−z0|αdsσ(z)

≤ Z

|z−z0|<δ|Qn(TN(z))|2|z−z0|αdsσ(z)

≤ (1 +η)|α|+1

|TN (z0)|α+1 Z

|z−z0|<δ|Qn(TN(z))|2|TN(z)−TN(z0)|α|TN (z)|dsσ(z)

≤ (1 +η)|α|+1

|TN (z0)|α+1 Z

0 |Qn(eit)|2|eit−eiπ/2|αdt

= (1 +η)|α|+1λnα, eiπ/2)

|TN (z0)|α+1 . This and (4.8) imply

λdeg(Rn)σ, z0)≤(1 +η)|α|+1 λnα, eiπ/2)

|TN (eiπ/2)|α+1 +O(qnτ0/2), from which

lim sup

n→∞ deg(Rn)α+1λdeg(Rn)σ, z0)

≤lim sup

n→∞ (nN+o(n))α+1(1 +η)|α|+1λnα, eiπ/2)

|TN (z0)|α+1

= (1 +η)|α|+1 Nα+1

|TN (z0)|α+12α+1Lα,

where we used (3.15). Since η > 0 is arbitrary, we obtain from (4.1) (use also (2.2))

lim sup

n→∞ nα+1λnσ, z0)≤ Nα+1

|TN (z0)|α+12α+1Lα= Lα

(πωσ(z0))α+1. (4.9) 4.2 The lower estimate

Let Pn be the extremal polynomial for λnσ, z0), and let Sn,z0,L be the fast decreasing polynomial given by Corollary 2.2 for the closed lemniscate domainLenclosed by σ(with some fixedτ <1). As before, we obtain from Lemma 2.7

kPnkσ =O(n(1+|α|)/2). (4.10)

(21)

Define Rn(z) =Pn(z)Sn,z0,L(z). Rn is a polynomial of degreen+o(n) and Rn(z0) = 1. Similarly to the previous section, we have

sup

z∈L\{t:|t−z0|<δ}|Rn(z)|=O(qnτ0/2) (4.11) for some q < 1 and τ0 > 0. Since the expression PN

k=1Rn(zk), where {z1, . . . , zN}=TN−1(TN(z)), is symmetric in the variables zk, it is a sum of their elementary symmetric polynomials. For more details on this idea, see [23]. Therefore, there is a polynomial Qn of degree at most deg(Rn)/N = (n+o(n))/N such that

Qn(TN(z)) = XN

k=1

Rn(zk), z∈σ.

We claim that for every z∈σ, we have

|Qn(TN(z))|2≤ XN

k=1

|Rn(zk)|2+O(qnτ0/2). (4.12) Indeed, sinceσ has no self intersection, |zk−zl|cannot be arbitrarily small for distinct k and l. As a consequence, for every z at most one zj belongs to the set{z : |z−z0|< δ} ifδ is sufficiently small, and hence, in the sum

|Qn(TN(z))|2 ≤ XN

k=1

XN

l=1

|Rn(zk)||Rn(zl)|, every term withk6=l isO(qnτ0/2) (use (4.10) and (4.11)).

Now letδ >0 be so small that for everyzwith|z−z0|< δthe inequalities in (4.7) hold. Then (4.2) and (4.12) give (note thatTN(z) =TN(zk) for all k)

Z

σ|Qn(TN(z))|2|TN (z)||TN(z)−TN(z0)|αdsσ(z)

≤O(qnτ0/2) + Z

σ

XN

k=1

|Rn(zk)|2

!

|TN (z)||TN(z)−TN(z0)|αdsσ(z)

=O(qnτ0/2) + Z

σ

XN

k=1

|Rn(zk)|2|TN(zk)−TN(z0)|α

!

|TN (z)|dsσ(z)

=O(qnτ0/2) +N Z

σ|Rn(z)|2|TN(z)−TN(z0)|α|TN (z)|dsσ(z)

≤O(qnτ0/2) + (1 +η)|α|+1|TN (z0)|α+1N Z

|z−z0|<δ|Pn(z)|2|z−z0|αdsσ

≤O(qnτ0/2) + (1 +η)|α|+1|TN (z0)|α+1N λnσ, z0).

Ábra

Figure 1: A typical position where Theorem 1.1 can be applied Theorem 1.1 Let the support Γ of a measure µ ∈ Reg consist of finitely many continua, and let z 0 lie on the outer boundary of Γ
Figure 2: A typical position where Theorem 1.2 can be applied (41)] or [19, Lemma 4.4.1]) ω Γ (x) = Q k 0 −1j=0 | x − λ j | π qQ 2k 0 +1 j=0 | x − a j | , x ∈ Int(Γ), (1.7)
Figure 3: The Γ and the leminscate σ as in the second half of Proposition 5.1
Figure 5: The arc J and the selection of Γ 0 m 8.1 Proof of Proposition 8.1
+2

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

Key words and phrases: Analytic functions, Univalent functions, Starlike functions, Convex functions, Integral operator, Fox- Wright function.. 2000 Mathematics

Abstract: In this note our aim is to present some Jordan-type inequalities for generalized Bessel functions in order to extend some recent results concerning generalized and

In this note our aim is to present some Jordan-type inequalities for generalized Bessel functions in order to extend some recent results concerning generalized and sharp versions of

Key words: Isotonic linear functionals, Jessen’s inequality, Superquadratic functions, Func- tional quasi-arithmetic and power means of Mercer’s type.... Jessen’s Inequality

Key words and phrases: Isotonic linear functionals, Jessen’s inequality, Superquadratic functions, Functional quasi-arithmetic and power means of Mercer’s type.. 2000

In this paper our aim is to show that the idea of using mathematical induction and infinite product representation is also fruitful for Bessel functions as well as for the

Key words: Analytic functions; Univalent functions; Coefficient inequalities and coefficient estimates; Starlike functions; Convex functions; Close-to-convex functions; k-

Key words and phrases: Riemann-Stieltjes integral, Functions of bounded variation, Lipschitzian functions, Integral inequal- ities, ˇ Cebyšev, Grüss, Ostrowski and Lupa¸s