• Nem Talált Eredményt

Asymptotics of Christoffel functions on arcs and curves

N/A
N/A
Protected

Academic year: 2022

Ossza meg "Asymptotics of Christoffel functions on arcs and curves"

Copied!
45
0
0

Teljes szövegt

(1)

Asymptotics of Christoffel functions on arcs and curves

Vilmos Totik

1

November 11, 2013

1Supported by NSF DMS0968530 and by ERC No. 267055

(2)

Contents

1 Results 1

2 Upper estimate for Christoffel functions 4 2.1 Part I: Γ0is a Jordan curve . . . 4 2.2 Part II: Γ0 is a Jordan arc . . . 25 3 The lower estimate for the Christoffel functions in Theorem 1.1

for positive weights 28

4 Proof of Theorems 1.1 and 1.2 36

5 Appendix 38

References 40

(3)

Abstract

For a system of smooth Jordan curves and arcs asymptotics for Christoffel func- tions is established. A separate new method is developed to handle the upper and lower estimates. In the course to the upper bound a theorem of Widom on the norm of Chebyshev polynomials is generalized.

(4)

1 Results

Letν be a finite Borel measure on the plane with compact support consisting of infinitely many points. The Christoffel functions associated withν are defined as

λn(z, ν) = inf

Pn(z)=1

Z

|Pn|2dν,

where the infimum is taken for all polynomials of degree at most n that take the value 1 atz.

Christoffel functions are closely related to orthogonal polynomials (for a survey see [13] by P. Nevai and [21] by B. Simon), to statistical physics (see e.g.

[15] by L. Pastur), to universality in random matrix theory (see e.g. the recent breakthrough [10] by D. Lubinsky, as well as [2],[22],[24]), to spectral theory (see e.g. [20], [21] by B. Simon and [1] by Breuer, Last and Simon) and to several other fields in mathematics. For the role and various use of Christoffel functions see [4], [6], [20], and particularly [13] by P. Nevai and [21] by B. Simon.

In this paper we consider asymptotics of Christoffel functions on smooth (C1+α) Jordan curves and arcs. Recall that a Jordan curve is the homeomorphic image of the unit circle while a Jordan arc is the homeomorphic image of [−1,1].

Thus, a Jordan arc has two endpoints. The asymptotics of Christoffel functions onC2-Jordan curves was established in [25] with a systematic use of polynomial inverse images of the unit circle (lemniscates). The idea of that paper was that many things can be carried over to lemniscates from the unit circle, and a system ofC2 Jordan curves can be well approximated (in a very specific sense) by lemniscates both from the inside and from the outside. This method does not work for arcs, and, in fact, except for the case when the set is a subset of the real line, no result has been known regarding Christoffel function asymptotics for arcs. In this paper we develop a method that handles both Jordan curves and arcs. We emphasize that we need a new method (actually very different ones) for both the upper and lower estimates, for previous methods do not work in either cases.

Thus, let Γ be the union of finitely manyC1+α,α >0, smooth Jordan curves and arcs lying exterior to one another, and letsΓ=sbe the arc measure on Γ.

Let Γk,k= 0,1, . . . , k0 be the disjoint components of Γ: Γ =∪kk=00 Γk. We call those Γk that are Jordan arcs the arc-components of Γ. Since we need C1+α smoothness just to have higher smoothness than C1, we may and shall always assume 0< α <1.

Our main theorem is

Theorem 1.1 LetΓbe a system ofC1+α-smooth Jordan arcs and curves lying exterior to one another, let z0 ∈ Γ be a point on Γ that is different from the endpoints of the arc components of Γ, and assume that Γ is C2-smooth in a neighborhood of z0. Assume that dν = wdsΓ is a measure on Γ with density w (with respect to the arc measure sΓ) which is continuous on Γ and positive

(5)

sΓ-almost everywhere. Then

nlim→∞n(z0, ν) = dν(z0) dµΓ

, (1.1)

where µΓ denotes the equilibrium measure of Γ, and on the right-hand side dν(z)/dµΓ is the Radon-Nikodym derivative of ν with respect toµΓ.

For the concepts from potential theory (like equilibrium measure, logarithmic capacity, Greens’ function etc.) see e.g. [5], [9], [17], or [19].

In the case that we are considering the equilibrium measureµΓis absolutely continuous with respect to the arc measuresΓ on Γ: dµΓ(t) =ωΓ(t)dsΓ(t) with a Cα-continuous density function ωΓ (see Proposition 2.2), and with it (1.1) takes the form

nlim→∞n(z0, ν) = w(z0)

ωΓ(z0). (1.2)

TheC1+α-smoothness could be replaced by piecewiseC1+α-smoothness with- out cusps, in which case Γ could have corners, and then the result is claimed for z0 which is not an endpoint or a corner (at endpoints and at corners the order of the Christoffel function is no longer 1/n, see [28]).

The global positivity and continuity ofwwas assumed only to have an easy formulation, the proof actually gives a much more general result. To this end we recall the class Reg from [23]: a measureν with support Γ is said to be in theRegclass if

nlim→∞

sup

Pn

kPnkΓ kPnkL2(ν)

1/n

= 1, (1.3)

where the supremum is taken for all polynomials of degree at mostn, and where kPnkΓstands for the supremum norm ofPnon Γ. This is not the standard defi- nition of theRegclass (which is in terms of the leading coefficients of orthogonal polynomials), but it is equivalent to it, see [23, Theorem 3.4.3,(v)]. See [23] for several other equivalent formulations and for general criteria implyingν ∈Reg.

We only mention here thatν ∈Regis a very weak global assumption onν, e.g.

it holds ifdν(t)/dsΓ(t)>0sΓ-almost everywhere. Therefore, Theorem 1.1 is a special case of

Theorem 1.2 LetΓbe a system ofC1+α-smooth Jordan arcs and curves lying exterior to one another, let z0 ∈ Γ be different from the endpoints of the arc components of Γ and assume that Γ is C2-smooth in a neighborhood of z0. Assume that dν = wdsΓ+dνsing is a measure on Γ with density w and with singular part νsing (with respect to the arc measure sΓ) which is in the Reg class. Then, if w is continuous at z0 andz0 is a Lebesgue-point for νsing, we have

nlim→∞n(z0, ν) = w(z0)

ωΓ(z0), (1.4)

whereωΓ denotes the density of the equilibrium measureµΓ (with respect tosΓ).

(6)

The Lebesgue-point property ofνsing mentioned in the statement is

νsing({ζ |ζ−z0| ≤τ}) =o(τ) asτ→0. (1.5) Let us mention that some kind of global condition likeν ∈Reg is needed, e.g. ifν vanishes on a subarc of Γ, then (1.4) is necessarily false because then

lim inf

n→∞n(z0, ν)> w(z0)

ωΓ(z0). (1.6)

We shall give a detailed proof for Theorem 1.1, the proof of Theorem 1.2 follows by simple changes. During the proof of the upper estimate in Theorem 1.1 we shall also verify (see Proposition 2.4)

Theorem 1.3 LetΓbe a finite union of disjointC1+αJordan curves and arcs.

Then there is a constant C and for every n= 1,2, . . . there are monic polyno- mials Pn(z) =zn+· · · of degreensuch that

kPnkΓ≤Ccap(Γ)n, wherecap(Γ) denotes the logarithmic capacity ofΓ.

This should be compared to the fact (see e.g. [17, Theorem 5.5.4]) that for any nand monic polynomialPn(z) =zn+· · · we have

kPnkΓ ≥cap(Γ)n.

Thus, the theorem says that on unions of smooth curves and arcs this theoretical lower bound can be achieved for every n disregarding a constant factor. For C2+αcurves and arcs this follows from deep results of Widom [30]. Let us also mention that if there are at least two components, or Γ is a single smooth arc, then the better estimate

kPnkΓ= (1 +o(1))cap(Γ)n

is impossible for alln([27], [26]). It is a delicate problem (connected with simul- taneous Diphantine approximation of the harmonic measures of the components of Γ) how close one can get by the norm of monic polynomials of degree nto the theoretical lower bound cap(Γ)n, see the papers [26] and [30].

First we shall deal with Theorem 1.1 in the special case whenwis continuous and positive on Γ. The general case of Theorem 1.1 will follow from this via a simple argument. The proof of the upper and lower estimates are distinctively different. The upper estimate will be obtained by a careful discretization of the equilibrium measure. That part of the proof holds at every Lebesgue-point ofν (Lebesgue-point with respect to arc-measure) and the localC2 property is not needed there. The lower estimate will be reduced to the case when there are no arc-components of Γ.

(7)

2 Upper estimate for Christoffel functions

In this section we establish that lim sup

n→∞n(z0, ν)≤dν(z0) dµΓ

. (2.1)

We need the concept of Lebesgue-point of a measure on Γ. Thus, letν be a Borel-measure on Γ and dν(t) =w(t)ds(t) +dνsing its decomposition into its absolutely continuous and singular parts with respect to arc measure s =sΓ. We say that z0 ∈ Γ, which is not an endpoint of an arc-component of Γ, is a Lebesgue-point forν (with respect to arc measure) if for everyε >0 there is a ρ >0 such that if 0≤τ ≤ρthen

Z

|ζz0|≤τ|w(ζ)−w(z0)|ds(ζ)≤ετ (2.2) and

νsing({ζ |ζ−z0| ≤τ})≤ετ . (2.3) Since the derivative ofνsingwith respect tosΓis 0sΓ-almost everywhere (see [18, Theorem 7.13]), standard proof shows thatsΓ-almost every point is a Lebesgue- point forν.

The main theorem of this part of the paper is

Theorem 2.1 Let Γ be a finite union of disjoint C1+α Jordan curves or arcs lying exterior to one another, andν a Borel measure onΓ. Ifz0∈Γ is not an endpoint of an arc-component of Γ and z0 is a Lebesgue-point (with respect to arc measuresΓ)ofν, then

lim sup

n→∞n(ν, z0)≤ dν(z0) dµΓ

.

For the proof of Theorem 2.1, let, as before, Γk be the disjoint components of Γ with Γ0 being the one containing z0. There is a change in the argument when Γ0is a Jordan arc as opposed to the case when it is a Jordan curve. First we consider the latter case, and return to the arc case after we have presented the proof for curves.

2.1 Part I: Γ

0

is a Jordan curve

Without loss of generality we may assume that z0 = 0 and that the real line is the tangent line to Γ at 0. Then in a neighborhood of 0 the curve Γ0 has a parametrization t+iγ(t) with γ ∈ Cα and γ(0) = 0, γ(0) = 0; hence

(t)| ≤C|t|α,|γ(t)| ≤C|t|1+α.

(8)

LetθkΓk), and for annconsider the integersnk = [θkn]. Divide each Γk intonk arcsIjk, (for eachkthe number of suchj’s isnk), each having equal weightθk/nk with respect toµΓ, i.e. µΓ(Ijk) =θk/nk. Then

θk

nk − 1 n

=|µΓ(Ijk)−1/n| ≤C/n2. (2.4) Let

ξjk= 1 µΓ(Ijk)

Z

Ijk

u dµΓ(u) (2.5)

be the center of mass with respect toµΓ. Simple argument shows that on Γ0we can choose theIj0’s so that the real part of one of theξ0j’s lying close to 0 is 0, sayℜξ00= 0. Indeed, since Γ0is a closed curve, the aforementioned subdivision can be started from any point on Γ0, i.e. if P ∈Γ0 is any point then there is a unique subdivision σP such that P is one of the division points. Take now any subdivisionσ, and in that subdivision let 0 lie in the subarc bc, with, say,b ℜb ≤ 0,ℜc ≥0 (recall that at 0 the x-axis is tangent to Γ), and let the two neighboring arcs of that subdivision beabb andcdb with ℜa <0,ℜc >0. Call a the left endpoint of ab. Now ifb P is moving on Γ0 froma toc in a continuous manner, then the subarcI(P) inσP for whichP is its left endpoint moves from abb to cd. Since the first one lies in the negative half-planeb ℜz ≤0, while the latter lies in the positive half-planeℜz≥0, in the first case the center of mass lies inℜz <0, while in the second case it lies in ℜz >0. Therefore, there will be a moment for which the center of mass of I(P) lies on the imaginary axis, and thenσP is the required subdivision, and we selectI(P) asI00. It then easily follows that ξ00 lies closest to 0 among theξjk’s.

Consider now the polynomial

Rn(z) =Y

j,k

(z−ξjk) (2.6)

of degree at most n. We claim that the polynomial

Pn(z) =Rn(z)/(z−ξ00) (2.7) verifies Theorem 2.1. We prove this via a series of propositions.

In what followsA∼B means that the ratioA/Bis bounded away from zero and infinity.

Proposition 2.2 dµΓ(t) =ωΓ(t)ds(t)with a positive density functionωΓwhich is Cα-smooth away from the endpoints of the arc-components ofΓ. If E is an endpoint of an arc-component ofΓ, thenωΓ(z)∼1/|z−E|1/2 around E.

This is a standard result. When Γ consists of one component which is a Jordan curve it immediately follows from the Kellogg-Warschawski theorem (see [16,

(9)

E=a1k b =a1k 2k

b =a2k 3k

b =a3k 4k

I

1k

I

2k

I

3k

x

1k

x

2k

x

3k

Figure 1: The choice of the intervalsIjk and of the pointsξjk

Theorem 3.6]). When Γ consists of several components, we could not find it in the appropriate form in the literature, hence we will present a proof in the Appendix to this paper. Actually, the proof gives that around an endpointE of an arc-component of Γ the functionωΓ(t)|t−E|1/2 is a positive Lipαfunction.

Since away from endpoints of arc-components of Γ the densityωΓis bounded away from 0 and infinity, it follows that away from the endpoints we have s(Ijk) ∼ 1/n, and if akj, bkj are the endpoints of the arc Ijk, then in this case

jk−akj| ∼1/n, |ξjk−bkj| ∼1/nand|ξjk−ξik| ∼ |j−i|/n.

Proposition 2.3 If E is an endpoint of an arc-component of Γ, say E ∈ I1k andI1k, I2k, . . .follow one another in this order onΓ, then|ξjk−E| ∼(j/n)2and s(Ijk)∼j/n2 in a neighborhood of E. Furthermore, if the endpoints of the arc Ijk areakj, bkj then

jk−akj| ∼ |ξjk−bkj| ∼s(Ijk)∼j/n2, (2.8) and

jk−ξki| ∼ |j2−i2|

n2 . (2.9)

See Figure 1.

Proof. LetIjk be the arcadkjbkj withakj lying closer toE. Then, by Proposition 2.2,

k

nk = Z

c

Ebkj

ωΓ(t)ds(t)∼ Z

c

Ebkj |t−E|1/2ds(t),

(10)

and since|t−E| ∼s(cEt) we can continue this as Z

c

Ebkj

s(cEt)1/2ds(t)∼s(Ebdkj)1/2∼ |E−bkj|1/2.

Therefore, |E−bkj| ∼ (j/n)2 and s(I1k) ∼ 1/n2 follow because θk/nk ∼ 1/n.

Sinceakj =bkj1, we also get forj ≥2 the relation|E−akj| ∼(j/n)2. Therefore, forj≥2

θk

nk

= Z

d

akjbkj

ωΓ(t)ds(t)∼ Z

d

akjbkj

((j/n)2)1/2ds(t)∼s(Ijk)(n/j), which, in view again ofθk/nk∼1/n, givess(Ijk)∼j/n2.

Since ξjk lies close to Ijk, |ξjk −E| ∼ (j/n)2 is immediate for j ≥ 2. To prove it for j = 1 we may assume temporarily (i.e. just for the proof of this relation) that E = 0 and R+ is the half-tangent to the arc Γk of Γ. Let the orthogonal projection of the arc I1k onto the real line be [0, d]. Then, as we have just seen, d ∼ 1/n2, and ℜξ1k is the center of mass of a measure ρ(t)dt on [0, d] for whichρ(t)∼t1/2. Elementary estimate shows then thatℜξ1k/d is bounded away from 0 and infinity (no matter how smalldis), which combined with diam(I1k)∼1/n2 yields the desired estimate|ξ1k| ∼(1/n)2.

The same argument verifies (2.8), while (2.9) follows from the other state- ments in the proposition: for example ifi < j ≤2i,i6=j then

kj −ξik| ∼s(ab kibkj) = Xj τ=i

s(Iτk)∼ Xj τ=i

(τ /n2)∼(j2−i2)/n2, while ifj >2ithen (use also the preceding relation withj= 2i)

kj −ξik| ∼ |E−ξjk| ∼j2/n2∼(j2−i2)/n2.

Proposition 2.4 For the polynomials (2.6) we have

kRnkΓ ≤Ccap(Γ)n (2.10) with some C independent ofn.

This almost proves Theorem 1.3, the only problem is that the degree ofRn

is P

kkn], which may be smaller than n but at most by k0. To have exact degreen one should divide some of the Γk’s into not [θkn] but [θkn] + 1 parts so as to get totallynarcs, and proceed as below.

(11)

Proof. By Frostman’s theorem (see [17, Theorem 3.3.4]) Z

log|z−t|dµΓ(t) = log cap(Γ), z∈Γ. (2.11) Note that (with log+= max(0,log))

Z

log+|z−t|dµΓ(t)≤log+diam(Γ),

hence Z

|log|z−t||dµΓ(t)≤2 log+diam(Γ)−log cap(Γ). (2.12) Now we write in view of (2.11)

nlog cap(Γ) = X

j,k

n− 1 µΓ(Ijk)

! Z

Ijk

log|z−t|dµΓ(t)

+ X

j,k

1 µΓ(Ijk)

Z

Ijk

log|z−t|dµΓ(t) = Σ1+ Σ2. (2.13) Here, by (2.4) and (2.12),

1| ≤X

j,k

O(1) Z

Ijk

log|z−t|µΓ(t)

=O(1). (2.14) Therefore, to prove the claim we have to show that on Γ

log|Rn(z)| −Σ2=X

j,k

1 µΓ(Ijk)

Z

Ijk

log

z−ξjk z−t

ωΓ(t)ds(t)≤C. (2.15) The proof uses the idea of [19, Theorem VI.4.2]. It is more involved around endpoints of arc-components of Γ, so we give it only there. Thus, let z lie in an arc Ijl0 that lies around an endpoint E of an arc-component Γl of Γ, on which, say, the arcs Ijl are following each other in the order I1l, . . . , Ijl0, ... with I1l containingE. zand (j0, l) will always have this meaning below. We consider the sum

X

(j,k)6=(j0,l)

1 µΓ(Ijk)

Z

Ijk

log

z−ξjk z−t

ωΓ(t)ds(t) =: X

(j,k)6=(j0,l)

Lj,k(z), (2.16)

and prove that it is uniformly bounded (both from below and above). Note that this sum differs from the one on the right of (2.15) in one term (the term with

(12)

E=a1l

z a b

j +10

j +10 j0

j0

I

l

x

l

I

l

x

l

Figure 2: The position ofz, a, b

integral overIjl0 is missing), and we shall actually show that not just the sum, but also the sum consisting of the absolute values|Lj,k|is uniformly bounded,

i.e. X

(j,k)6=(j0,l)

|Lj,k(z)|=O(1). (2.17) First we verify that the individual terms Lj,k(z) in (2.16) are uniformly bounded. This is clear for k6=l (i.e. whenIjk is on a different component of Γ thanz) or for k=lbutj 6=j0±1 (thej=j0term is not in the sum), for then in the integrand

|z−ξkj| ∼dist{Ijl0, Ijk} ∼ |z−t| for allt∈Ijk.

So let j = j0±1, say j = j0+ 1. Then we know from Proposition 2.3 that

|z−ξlj0+1| ∼s(Ijl0+1)∼j0/n2, and from Propositions 2.2 and 2.3 thatωΓ(t)≤ Cn/j0 onIjl0+1. LetIjl0+1 be the arcab, see Figure 2. Clearlyb

Lj0+1,l(z) = 1 µΓ(Ijl0+1)

Z

Ijl

0 +1

log

z−ξjl0+1 z−t

ωΓ(t)ds(t)

≤ Cnn j0

Z

Ijl

0 +1

log|z−ξjl0+1|+ log 1

|a−t|

ds(t). (2.18) HereZ

Ijl

0 +1

log 1

|a−t|ds(t)≤ Z

Ijl

0 +1

log C0

s(at)b ds(t) =s(Ijl0+1)(logC0+1−logs(Ijl0+1)).

Therefore, the integral on the right of (2.18) equals s(Ijl0+1) log|z−ξlj0+1|

s(Ijl0+1) +O s(Ijl0+1)

≤Cs(Ijl0+1)≤Cj0

n2.

(13)

If we substitute this into (2.18) then we obtain the boundedness of Lj0+1,l(z) from above. Its boundedness from below is clear since for z∈Ijl0,t∈Ijl0+1 we

have

z−ξjl0+1 z−t

≥c >0 (2.19)

by (2.8).

The casej =j0−1 is similar providedj0−1>1, but for j0−1 = 1, we must proceed somewhat differently, for then ωΓ(t)≤ Cn/j0 is no longer true on I1l. In this case (i.e. when Ijl01 = I1l) we have µ(I1l) ∼ 1/n ∼ s(I1l)1/2,

|z−t| ∼s(zt), sob

L1,l ≤ C s(ab)b 1/α

Z b

ab

logCs(ab)b

s(zt)b s(at)b 1/2ds(t),

and the right-hand side will be shown to be bounded from above in the proof of (2.24)–(2.25) (the boundedness from below ofL1,l follows again from (2.19)).

These prove the uniform boundedness of the individual termsLj,k, (j, k)6= (j0, l).

It follows from Proposition 2.3 that there is anM such that if either k6=l ork=lbut |j−j0| ≥M then forz∈Ijl0 andt∈Ijk we have

ξjk−t z−ξkj ≤1

2

(a closer look at the proof of Propositions 2.2 and 2.3 reveals thatM = 4 suffices for large n, but we do not need the best value ofM).

Thus, in this case for the integrands inLj,k(z) we get (use that log|1−u|= ℜlog(1−u) with any local branch of the log)

log

z−ξjk z−t

=−log

1− ξkj −t z−ξjk

=ℜξjk−t z−ξjk +O

ξjk−t z−ξkj

2

.

Therefore, for such j andkwe have

|Lj,k(z)|= 1 µΓ(Ijk)

Z

Ijk

O

ξkj −t z−ξjk

2

dµΓ(t) =O s(Ijk)2

jk−ξlj0|2

!

, (2.20)

because the integral Z

Ijkℜξjk−t

z−ξjkΓ(t) =ℜ Z

Ikj

ξkj −t z−ξjkΓ(t)

(14)

vanishes by the choice ofξjk.

The expression on the right of (2.20) is bounded by a constant timess(Ijk)2 when k 6= l or k = l but Ijl is far from E (say farther than a fixed constant δ >0), and fork=l andIjl close toE (say for|ξkj −E| ≤δ) it is at most (see Proposition 2.3) a constant times

s(Ij)2

|(j/n)2−(j0/n)2|2 ∼ (j/n2)2

|(j/n)2−(j0/n)2|2 = j2

|j2−j02|2.

All in all, if we take into account the uniform boundedness of the termsLj,k

we obtain that the sum in (2.17) is at most X

|jj0|≤M, j6=j0

|Lj,l| + X

|jj0|>M

|Lj,l|+ X

j,k, k6=l

|Lj,k|

≤ (2M)C+C X

|jj0|>M

j2

|j2−j02|2+CX

j,k

s(Ijk)2≤C.

To complete the proof of the proposition we have to show that the additional term

1 µΓ(Ijl0)

Z

Ijl

0

log

z−ξjl0 z−t

ωΓ(t)ds(t) (2.21)

in (2.15) is also bounded from above (from below we cannot claim boundedness forz can be very close toξlj0). As before, we get from Proposition 2.3 that for j0>1 this term is at most

Cn Z

Ij0l

logCs(Ijl0) s(zt)b

! j02 n2

1/2

ds(t),

which, withIjl0=:ab, equalsb Cn2

j0

s(ab) log(Cs(b ab))b −s(zb) logb s(zb)b −s(caz) logs(caz) +s(ab)b

. (2.22) Now we use for 0≤x≤y≤1 the inequality

−2

e(x+y)≤xlogx+ylogy−(x+y) log(x+y)≤0, (2.23) which is immediate from the concavity of log and from the fact that on the interval (0,1) the minimum oftlogtis−1/e. Apply (2.23) withs(zb),b s(caz) in place ofx, y (in which casex+y=s(ab)) to continue (2.22) asb

≤Cn2 j0

s(ab) log(Cs(b ab))b −s(ab) logb s(ab) +b O(s(ab))b

≤Cn2 j0

s(ab)b ≤C,

(15)

E=a

b z

w

I

1l

x

l1

Figure 3: The choice ofw

where, in the last step we used that by Proposition 2.3,s(ab) =b s(Ijl0)∼j0/n2. This gives the required estimate for (2.21) whenj0>1.

Whenj0= 1 thenEis an endpoint of the arcIjl0, e.g. E=a. In that caseωΓ

is not bounded onIjl0, so we have to proceed differently than before. Similarly as above, now we have withs(ab) =b s(Ijl0)∼1/n2Γ(I1l)∼1/n∼s(ab)b 1/2the bound

C s(ab)b 1/2

Z b

ab

logCs(ab)b

s(zt)b s(at)b 1/2ds(t) =:I (2.24) for the expression in (2.21). Recall that z lies on the arc abb = Eb, and letc w be the midpoint on the arcEzc in the sense thats(dEw) =s(wz), see Figure 3.c Now we split the integral in (2.24) over abb into three parts: the integrals over zb,b wzc andEw. For the first we have (use that the antiderivative ofd t1/2logt is 2t1/2logt−4t1/2)

Z b

zb

logCs(ab)b

s(zt)b s(cEt)1/2ds(t)≤ Z

b

zb

logCs(ab)b

s(zt)b s(zt)b 1/2ds(t)

= 2 log(Cs(ab))s(b zb)b 1/2−2s(zb)b 1/2logs(zb) + 4s(b zb)b 1/2≤Cs(ab)b 1/2 because, for anyC0> e2 (by the monotonicity ofx1/2log 1/xon (0, e2)),

2s(zb)b 1/2logC0s(ab)b

s(zb)b ≤2s(ab)b 1/2logC0s(ab)b

s(ab)b = 2s(ab)b 1/2logC0. The integral overwzc can be similarly handled. Finally, for the integral over Ewdwe have the bound

Z c

Ew

logCs(ab)b

s(dEw)s(cEt)1/2ds(t) ≤ logCs(ab)b

s(dEw)2s(dEw)1/2≤logCs(ab)b

s(ab)b 2s(ab)b 1/2

= 2(logC)s(ab)b 1/2.

(16)

Substituting all these into (2.24) we get

I≤C, (2.25)

and with this the upper boundedness of (2.21) forj0= 1, as well.

Proposition 2.5 For the polynomials from (2.7) we have

|Pn(z)| ∼ncap(Γ)n (2.26)

uniformly in nandz∈I00, in particular

|Pn(0)| ∼ncap(Γ)n. (2.27) Forz∈Γ\I00

|Pn(z)| ≤Ccap(Γ)n 1

|z| (2.28)

with some C independent ofn.

Proof. Let z ∈ I00, i.e. with the notations of the preceding proof we have l=i0= 0. By the proof of Proposition 2.4 (see in particular (2.11)–(2.14)) and (2.17)) we have uniformly innandz∈I00

log|Pn(z)| −nlog cap(Γ) + 1 µΓ(I00)

Z

I00

log|z−t|ωΓ(t)ds(t) =O(1). (2.29) Now use that |z−t|= (1 +o(1))s(zt) (forb z−t ∼0) to get with I00=:abb for the last term in (2.29)

1 µΓ(I00)

Z

I00

log|z−t|ωΓ(t)ds(t) = 1 µΓ(I00)

Z

I00

log

(1 +o(1))s(zt)b

ωΓ(t)ds(t)

= o(1) + 1 µΓ(I00)

Z

I00

logs(zt)ωb Γ(t)ds(t).

Here we need that for t∈I00 Proposition 2.2 yields

Γ(t)−ωΓ(0)| ≤C|t|α≤Cnα to continue the preceding estimates as

= o(1) +ωΓ(0)(1 +O(nα)) µΓ(I00)

Z

I00

logs(zt)ds(t)b

(17)

= o(1) +ωΓ(0)(1 +O(nα))

µΓ(I00) (s(caz) logs(az) +c s(zb) logb s(zb)b −s(ab))b

= o(1) +ωΓ(0)(1 +O(nα)) µΓ(I00)

s(ab) logb s(ab) +b O(s(ab))b ,

where, in the last step we used again (2.23) with s(caz) ands(zb) playing theb role ofx, y (note that then x+y=s(ab)).b

Since

ωΓ(0)

µΓ(I00)s(ab) = 1 +b O(nα) and

logs(ab) =b O(1)−logn

are also true (the latter one follows from the first one in view of µΓ(I00) = (1 +o(1))/n), finally we can conclude

1 µΓ(I00)

Z

I00

log|z−t|ωΓ(t)ds(t) =O(1)−logn.

This and (2.29) prove (2.26).

The claim (2.28) follows immediately from Proposition 2.4, for|z−ξ00| ∼ |z| whenz∈Γ\I00.

Label the pointsξj0around 0 in such a way that, as their real part increases, they follow each other in the order

· · ·<ℜξ02<ℜξ01<0 =ℜξ00<ℜξ01<ℜξ20<· · ·.

We may also assume that this labeling is such that the range ofj includes all integers in [−τ n, τ n] for someτ >0.

Proposition 2.6 For all j we have

ξj0− j nωΓ(0)

≤C

|j| n

1+α

. (2.30)

Proof. Enough to prove this for |ξj0| ≤ δ with some small δ > 0 (otherwise the discussion below gives|j| ≥cδnand then the statement is obvious).

Recall that the real line is the tangent line to Γ at 0 and in a neighborhood of 0 the curve Γ has a parametrization t+iγ(t) withγ ∈ Cα and γ(0) = 0, γ(0) = 0,|γ(t)| ≤C|t|α,|γ(t)| ≤C|t|1+α.

(18)

Letu=t+iγ(t)∈Γ. Then ds(u) =p

1 + (γ(t))2dt=dt+O(|u|)dt. (2.31) Let a, bbe the endpoints of I00, ℜa <ℜb. We know that|a|,|b| ∼1/n (this is immediate from the facts thatℜξ00= 0, the equilibrium densityωΓis continuous and positive at 0, andds(u)∼dtby (2.31)). We can write

0 = ℜξ00= n0

θ0

Z

I00ℜu ωΓ(u)ds(u) =n0

θ0

Z

I00ℜu ωΓ(0)ds(u) +O

n1 nnα1

n

= n0

θ0

Z b a

Γ(0)dt+O(n1α)

= n0ωΓ(0) θ0

1

2((ℜb)2−(ℜa)2) +O(n1α),

from which it follows that (note n∼n0,ℜb− ℜa∼1/n)

|ℜb+ℜa|=O(n1α).

Now lett0=ℜ(a+b)/2 +iγ(ℜ(a+b)/2). (2.31) implies

s(tc0b) =ℜb− ℜt0+O(n1α); s(atc0) =ℜt0− ℜa+O(n1α) and hence

s(tc0b)−s(atc0) =O(n1α).

Therefore, ifξ0j is the midpoint of the arc Ij0 with respect to arc length, then

|t0−ξ00| ≤Cn1α. Since

|t0−ξ00| ≤ |t0|+|ξ00| ≤Cn1α

is also true, finally we obtain |ξ00−ξ00| ≤Cn1α. Note that by the definition ofξ0j and theCα-smoothness ofωΓ we also have

µΓ(aξd00) = 1

Γ(ab) +b O(nα) µΓ(ξc00b) = 1

Γ(ab) +b O(nα)

Sinceξj00j are geometric quantities defined in terms ofωΓ andsΓ, the same argument can be given for allj and we obtain

j0−ξ0j| ≤Cn1α, |ξj| ≤δ, (2.32)

(19)

and (withaj, bj being the endpoints ofIj0) µΓ(adjξ0j) = 1

Γ(adjbj) +O(nα) µΓ(ξd0jbj) = 1

Γ(adjbj) +O(nα).

These latter imply forj6= 0, say for j >0, jθ0

n0

= µ

j1

[

l=0

Il0

!

Γ(ξd00ξ0j) +O(n1α)

= Z ξ0j

ξ00

ωΓ(0)ds(u) +O(|ξ0j|1+α) +O(n1α)

= Z ξ0j

ξ00

ωΓ(0)dt+O(|ξ0j|1+α) = (ℜξ0j− ℜξ00Γ(0) +O(|ξ0j|1+α)

= (ξ0j−ξ00Γ(0) +O(|ξ0j|1+α),

which, in view of|ξ00| ≤Cn1α,|ξ0j| ≤Cj/nand (2.4) implies

ξ0j− j nωΓ(0)

≤C

|j| n

1+α .

The argument for negativej is just the same. Finally, this inequality combined with (2.32) gives (2.30).

Fix a large integer numberM and a smallρ >0, so small that evenραM is small. Let

Qn(z) = Y

M3<|j|≤ρn

(z−ξ0j). (2.33)

Proposition 2.7 . For z∈Γ,|z| ≤M/nwe have

Qn(z)

Qn(0) −1

CM ρα+ 1 M

(2.34) with a C that depends only onΓ.

Proof.

log Qn(z)

Qn(0)

= X

M3<|j|≤ρn

log 1− z

ξj0 ,

(20)

and on applying (2.30) this can be written as X

M3<jρn

log

1− z

j/nωΓ(0) +O((j/n)1+α) 1− z

−j/nωΓ(0) +O((j/n)1+α)

= X

M3<jρn

log

1 +O(|z|(j/n)1+α) +O(|z|2) (j/n)2

= X

M3<jρn

O

|z|(j/n)α1+ |z|2 (|j|/n)2

=O

(M/n)n1α(ρn)α+(M/n)2 M3/n2

,

from which the claim follows.

The key statement in the proof of Theorem 2.1 is

Proposition 2.8 Let Γδ be the part of Γ that lies of distance ≥ δ from the origin. Then

δlim0ℜ Z

Γδ

Γ(u)

u = 0. (2.35)

The statement is that the real part of the principal value integral PV

Z

Γ

Γ(u)

u (2.36)

is zero at 0. This is due to the fact that the tangent line to Γ at 0 is horizontal.

Proof. Let Γδ be the complementary arc, i.e. the set of points on Γ which are closer to 0 thanδ. With some local branch of log we have to show that

δlim0ℜ Z

Γδ

(log(z−u))

z= 0dµΓ(u) = 0.

Here, with z =x+iγ(x) ∈ Γ, u= t+iγ(t)∈ Γ (with some global t+iγ(t) parametrization of Γ that extends the local parametrizationt+iγ(t) around the origin, see the discussion before (2.4))

ℜ Z

Γδ

(log(z−u))

z= 0dµΓ(u) =

= lim

x0

1 x+iγ(x)

Z

t+iγ(t)Γδ

log|(x+iγ(x))−(t+iγ(t))|

|t+iγ(t)| dµΓ(t+iγ(t)).

(21)

Since the whole integral Z

Γ

log|z−u|dµΓ(u) (2.37)

is constant on Γ (see (2.11)), what we need to show is that the previous expres- sion with Γδreplaced by Γδ tends to 0 asδ→0 (in this case the existence of the limit/derivative follows from what we have just done and from the constancy of (2.37)).

Sincex/(x+iγ(x))→1 asx→0, we need to show that 1

x Z

t+iγ(t)Γδ

log|(x+iγ(x))−(t+iγ(t))|

|t+iγ(t)| dµΓ(t+iγ(t)) =: 1

xI (2.38) is as small in absolute value as we wish for small|x|and small, but fixedδ >0.

Without loss of generality assume x > 0. Let the endpoints of Γδ be −δ1+ iγ(−δ1) andδ2+iγ(δ2),δ1, δ2>0. Thenδ2j+γ(δj)22, and hence, in view ofγ(δj) =O(δj1+α), we have

δj=δ+O(δ1+2α), j= 1,2. (2.39) With some largeN

I = Z δ2

δ1

log|(x+iγ(x))−(t+iγ(t))|

|t+iγ(t)| ωΓ(t+iγ(t))p

1 + (γ(t))2dt

=

Z N x

δ1

+ Z N x

N x

+ Z δ2

N x

=I1+I2+I3. First we deal withI2. It is the sum of

I21= Z N x

N x

log|(x+iγ(x))−(t+iγ(t))|

|x−t| ωΓ(t+iγ(t))p

1 + (γ(t))2dt,

I22=− Z N x

N x

log|t+iγ(t)|

|t| ωΓ(t+iγ(t))p

1 + (γ(t))2dt and

I23= Z N x

N x

log|x−t|

|t| ωΓ(t+iγ(t))p

1 + (γ(t))2dt.

In I21 the log term is log(1 +O((N x)α) = O((N x)α)) because, with some ζ∈[−N x, N x],

|γ(x)−γ(t)|=|x−t||γ(ζ)| ≤ |x−t|C(N x)α,

(22)

soI21=O(N1+αx1+α) =o(x) asx→0. Similarly,I22=o(x). Finally, I23 =

Z N x

N x

log|x−t|

|t|

ωΓ(t+iγ(t))p

1 + (γ(t))2−ωΓ(0) dt

+ ωΓ(0) Z N x

N x

log|x−t|

|t| dt=I231+I232.

The factor after the log term inI231is in absolute value≤C|t|α≤C(N x)αand Z N x

N x

log|x−t|

|t| dt=x

Z N

N

log|1−t|

|t|

dt≤CxlogN,

henceI231=O(Nα(logN)x1+α) =o(x). ForI232, as simple calculation shows, we can write

|I232| ωΓ(0) =

Z N x (N1)x

logx+t t dt≤

Z N x (N1)x

x

tdt≤ x N−1. So|I2| ≤ωΓ(0)x/(N−1) +o(x).

ForI1+I3 we setJ = [−δ1,−N x]∪[N x, δ2] and note that the log term in the integrals in I1 andI3is

ℜlog

1−x+iγ(x) t+iγ(t)

= −ℜx+iγ(x)

t+iγ(t) +Ox t

2

= −xt+γ(x)γ(t)

t2+γ(t)2 +Ox t

2

= −x t +xO

γ(t)2 t3

+O

γ(x)γ(t) t2

+Ox t

2

Here on the rightγ(t)2/t3andγ(t)/t2are integrable, so the contribution to the integral over Γδ of the corresponding terms isxoδ(1) and γ(x)oδ(1) =xoδ(1), respectively, where oδ(1) means a quantity tending to 0 as δ→0. The contri- bution of the termO(x2/t2) is

Z

J

Ox t

2 dt=O

x2 N x

=Ox N

.

Finally, the integral overJ of the term−x/t is equal to

− Z

J

x t

ωΓ(t+iγ(t))p

1 + (γ(t))2−ωΓ(0)

dt+ωΓ(0) Z

J

x

tdt=I4+I5.

(23)

InI4 we have

Γ(t+iγ(t))p

1 + (γ(t))2−ωΓ(0)| ≤C|t|α, so exactly as beforeI4=xoδ(1). Finally,

|I5|=ωΓ(0)x

Z max(δ12) min(δ12)

1

tdt≤Cxδ1+2α

δ =xoδ(1)

where we used (2.39). All in all, we have|I| ≤xoδ(1) +o(x) +O(x/N) which shows that the term in (2.38) is as small as we wish if we select N large and thenδ >0 small (and also xsufficiently small after these selections).

Proposition 2.9 Let

Sn(z) = Y

|ξkj|≥δ

(z−ξkj). (2.40)

Then, for fixed M andz ∈Γ, |z| ≤ M/n, we have |Sn(z)/Sn(0)| = 1 +oδ(1) uniformly in n.

Proof. As always, we setz=x+iγ(x).

1 nlog

Sn(z)

Sn(0) = X

|ξjk|≥δ

1 nlog

1− z

ξjk

(2.41)

is easily seen to converge to Z

Γδ

log1− z u

Γ(u) (2.42)

(recall, that Γδis the part of Γ that lies of distance≥δfrom the origin). Indeed, the same sum on the right of (2.41) with 1/nreplaced byµΓ(Ijk) andξkj replaced byξkj (that was the midpoint of Ijk with respect to arc length) is essentially a Riemannian sum for the integral (2.42), and the sums with ξjk, 1/n and with ξkjΓ(Ijk) are very close because of (2.4) and (2.32) and its analogue for other intervals. The integral in (2.42) is

Z

Γδ

−z u

+O |z| u

2!!

Γ(u), and here

Z

Γδ

O |z| u

2!

Γ(u) =O |z|2

δ

,

(24)

while Z

Γδ

−z u

Γ(u) =−xℜ Z

Γδ

Γ(u)

u +γ(x)ℑ Z

Γδ

Γ(u) u .

Here the second term isO(γ(x)/δ) =o(|x|) =o(|z|) asz→0, and for the first term Proposition 2.8 gives that it isxoδ(1).

Therefore, log

Sn(z)

Sn(0)

=n(1 +o(1))h

|z|oδ(1) +O(|z|2/δ) +o(|z|)i

≤2M oδ(1) for large nand|z| ≤M/n,z∈Γ. This proves the claim.

In the polynomialQnin (2.33) we put all factors (z−ξj0) withM3<|j| ≤ρn, whileSn in (2.40) contained the factors (z−ξkj) with|ξkj| ≥δ. For sufficiently small δ these latter include all ξkj with k > 0 (i.e. which are created for the components Γk,k >0). Furthermore, for smallδ >0 if, with a sufficiently large fixedL we selectρ=ωΓ(0)δ−Lδ1+α, then (2.30) shows thatQn and Sn have no common factors. On the other hand, if we selected ρ = ωΓ(0)δ+Lδ1+α, then (2.30) shows that all the factors (z−ξjk) except for (z−ξj0) with |j| ≤ M3 appear either in Qn or in Sn. We make the former selection, i.e. we set ρ=ωΓ(0)δ−Lδ1+α, and let ˜Sn be the product of all factors (z−ξj0) for which

|j|> ρnbut |ξ0j|< δ (these are the ones with|j|> M3 that appear neither in Qn nor inSn). According to what we have just said, their number is at most 4Lδ1+αn.

Proposition 2.10 We have|S˜n(z)/S˜n(0)|= 1+oδ(1)uniformly innand|z| ≤ M/n,z∈Γ.

Proof. LetH be the set of j’s for which|j|> ρnbut |ξ0j|< δ. Note that all suchξj0’s satisfy|ξj0| ≥δ/2 (see (2.30) and the definition ofρ). Now

log

n(z) S˜n(0) =X

jH

log 1− z

ξj0 =X

jH

O |z|

j0|

!

=O M

n

4Lδ1+αn δ

,

from which the claim follows.

After these preparations we turn to the proof of Theorem 2.1.

From the definition of our polynomials it follows that Pn(z) =Qn(z)Sn(z) ˜Sn(z) Y

M3jM3, j6=0

(z−ξ0j) =:Qn(z)Sn(z) ˜Sn(z)Vn(z),

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

However, this extension does not come directly from the polynomial inverse image method, for there is a huge difference between smooth Jordan arcs and Jordan curves: the interior

Given a set of planar curves (Jordan arcs), each pair of which meets – either crosses or touches – exactly once, we establish an upper bound on the number of touchings5. We show

However, the method of upper and lower solutions for the existence of solution is less developed and hardly few results can be found in the literature dealing with the upper and

∗ AMS Subject Classification 42C05, 30C85, 31A15; Key words: Christoffel functions, asymptotics, families of Jordan curves, harmonic measures.. † Supported by

The the precise asymptotics for the error of best rational approxima- tion of meromorphic functions in integral norm is shown to be a conse- quence of a result of Gonchar

3; the simplified construction of the joining arcs (in the middle) avoids the letter to be thinner in the middle than its upper and lower

The current best bound for incidences between points and general curves in R 2 is the following (better bounds are known for some specific types of curves, such as circles and

W ang , Global bifurcation and exact multiplicity of positive solu- tions for a positone problem with cubic nonlinearity and their applications Trans.. H uang , Classification