• Aucun résultat trouvé

Where are the zeroes of a random p-adic polynomial?

N/A
N/A
Protected

Academic year: 2021

Partager "Where are the zeroes of a random p-adic polynomial?"

Copied!
44
0
0

Texte intégral

(1)

HAL Id: hal-02557280

https://hal.archives-ouvertes.fr/hal-02557280v2

Preprint submitted on 7 Oct 2021

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Where are the zeroes of a random p-adic polynomial?

Xavier Caruso

To cite this version:

Xavier Caruso. Where are the zeroes of a random p-adic polynomial?. 2021. �hal-02557280v2�

(2)

Where are the zeroes of a random p-adic polynomial?

Xavier Caruso October 7, 2021

Abstract

We study the repartition of the roots of a randomp-adic polynomial in an algebraic closure of Qp. We prove that the mean number of roots generating a fixed finite extensionKofQpdepends mostly on the discriminant ofK, an extension containing less roots when it gets more ramified. We prove further that, for any positive integerr, a randomp-adic polynomial of sufficiently large degree has aboutrroots on average in extensions of degree at mostr.

Beyond the mean, we also study higher moments and correlations between the number of roots in two given subsets ofQp(or, more generally, of a finite extension of Qp). In this perspective, we notably establish results highlighting that the roots tend to repel each other and quantify this phenomenon.

Contents

1 Density functions 6

2 Examples and closed formulas 16

3 Some orders of magnitude 24

4 The setup of ´etale algebras 30

Introduction

The distribution of roots of a random real polynomial is a classical subject of research that has been thoroughly studied since the pioneer work of Bloch and Polya [4], Littlewood and Offord [13, 14, 15] and Kac’s famous paper [10], in which an exact formula giving the avegare number of roots of a random polynomial with gaussian coefficients appears for the first time.

Investigating similarly the behaviour of roots ofp-adic random polynomials is a natu- ral question which have recently received some attention. The story starts in 2006 when Evans published the article [8], in which he managed to adapt Kac’s strategy and even- tually compute the average number of zeros in Zp of a random polynomial of degreen with coefficients uniformly1 distributed in Zp. The same year, Buhler, Goldstein, Moews and Rosenberg [5] found formulas for the probability that a randomp-adic polynomial

1By uniform distribution, we mean the distribution coming from the Haar measure on the compact group (Zp,+). It turns out that it is the correctp-adic analogue of the normal distribution.

(3)

has all its roots inQp. After about ten years without further significant contributions, the subject was revived a couple of years ago by Lerario and his collaborators who started to undertake a systematical study of these phenomena. With Kulkarni [11], they notably extend Crofton’s formula to the p-adic setting and derive new estimations on the num- ber of roots of ap-adic polynomial, establishing in particular that a uniformly distributed random polynomial of fixed degree overZp has exactly one root inQp on average, inde- pendently frompand from the degree. On a slightly different note, Ait El Mannsour and Lerario [1] obtain formulas counting the average number of lines in random projectivep- adic varieties. More recently, the case of nonuniform distributions has also been addressed by Shmueli [20], who came up with sharp estimations on the average number of roots.

Most of the aforementioned works are concerned with themean of the random vari- ableZncouting the number of roots of ap-adic polynomial of degreen. Beyond the mean (for which one can rely on Kac’s techniques), obtaining more information about the Zn’s is a fundamental question that has been recently addressed and elegently solved by Bhar- gava, Cremona, Fisher and Gajovi´c [3]. In their paper, they set up a general strategy to compute all probabilities Prob[Zn=r]withnandr running over the integers. In addition, they observed that the formulas they obtained are all rational functions in p which are symmetric under the transformation p p−1. This beautiful and fascinating property remains nowadays quite mysterious.

Apart from the distinction between archimedean and nonarchimedean,Qpdiffers from Rin that its arithmetic is definitely much richer; while the absolute Galois group ofRis somehow boring, that of Qp is large, intricated and encodes much arithmetical subtle information. In other words, the set of finite extensions ofQp has a prominent structure which is part of the strength and the complexity of the p-adic world. Therefore, looking at the roots of a randomp-adic polynomial not only inQp but in an algebraic closureQ¯p ofQp sounds like a very natural and appealing question, which is the one we address in the present paper.

To this end, we fix a finite extension F of Qp together with an algebraic closure F¯ ofF. We endowF¯ with the p-adic norm k · knormalized bykpk = p−[F:Qp] and use the letterq to denote the cardinality of the residue field ofF. Given a positive integer n, a finite extension K of F and a compact open subset U of K, we introduce the random variableZU,ncounting the number of roots inU of a random polynomial of degreenwith coefficients in the ring of integersOF ofF. Our first theorem gives an integral expression of the expected values of theZU,n’s.

Theorem A. There exists a family of functionsρK,n : K R+ (K running over the set of finite extensions ofF included inF¯ andnrunning the set of positive integers) satisfying the following property: for any positive integern, any extension finite extensionK of F sitting insideF¯ and any open subsetU ofE, we have:

E[ZU,n] = X

K⊂K

Z

U∩K

ρK,n(x)dx

where the sum runs over all extensionsK ofF included inK.

The functionsρK,n’s are called thedensity functionsas their values at a given pointx reflect the number of roots one may expect to find in a small neighborhood of x. Our second theorem provides rather precise information about the density functions.

Theorem B. Letnbe a positive integer, letK F¯ be a finite extension ofF and letx K.

Writerfor the degree of the extensionK/F.

(4)

1. (Vanishing)IfF[x]6=K orn < r, thenρK,n(x) = 0.

2. (Continuity)The functionρK,n is continuous onK.

3. (Invariance under isomorphisms) Given a second finite extension Lof F and an iso- mophism ofF-algebrasσ:K L, we haveρK,n(x) =ρL,n σ(x)

. 4. (Transformation under homography)For

a b c d

GL2(OF), we have:

ρK,n

ax+b cx+d

=kcx+dk2r·ρK,n(x).

5. (Monotony) We have ρK,n(x) ρK,n+1(x) and the inequality is strict if and only if F[x] =Kandrn <2r1.

6. (Formulas for extremal degrees)IfF[x] =Kandx∈ OK, then ρK,r(x) =kDKk · 1

# OK/OF[x] ·qr+1qr qr+11 forn2r1, ρK,n(x) =kDKk ·

Z

OF[x]ktkrdt whereDKis the discrimant of the extensionK/F.

A first remarkable consequence of Theorem B is that the functionsρK,n are indepen- dant ofn provided that n 2r1. A similar behaviour was already noticed in [3] for higher moments of the random variablesZF,n. Besides, it is in theory feasible to derive from Theorem B closed formulas forρK,nand its integral overK, at least whenn=r or n2r1. For example, Theorem C below covers the case of quadratic extensions. Before stating it, it is convenient to introduce the notation:

ρn(K) = Z

K

ρK,n(x)dx. (1)

By Theorem A, ρn(K) counts the number of roots of a random polynomial of degreen which fall insideKbut outside all strict subfields ofKcontainingF.

Theorem C. LetK be a quadratic extension ofF. (i) IfK/F is unramified, we have:

ρ2(K) = q2q+ 1 q2+q+ 1,

forn3, ρn(K) = q4+ 1

q4+q3+q2+q+ 1.

(ii) IfK/F is totally ramified, we have:

ρ2(K) =kDKk · q2 q2+q+ 1,

forn3, ρn(K) =kDKk · q2(q2+ 1) q4+q3+q2+q+ 1.

(5)

WhenKis the quadratic unramified extension ofF, we notice thatρn(K)is a rational function in q which is self-reciprocal, i.e. invariant under the transformation q q−1. As recalled previously, this remarkable property also holds for all higher moments of the random variablesZF,n. On the contrary, whenK/F is totally ramified, the functionρd(K) is not self-reciprocal. One can nevertheless recover the expected symmetry by summing up over all totally ramified quadratic extensions of F. Indeed, using Serre’s mass for- mula [18], we end up with:

X

K

ρ2(K) = 2q q2+q+ 1,

forn3, X

K

ρn(K) = 2q(q2+ 1) q4+q3+q2+q+ 1.

(where both sums run over all totally ramified quadratic extensions ofF sitting insideF¯) which are indeed self-reciprocal rational functions inq.

Another panel of interesting corollaries of Theorem B concerns the orders of magnitude of the functions ρK,n’s. Roughly speaking, Theorem B tells us that the size of ρK,n is controlled by thep-adic norm ofDK. It is in fact even more transparent when we integrate over the entire space.

Theorem D. LetK F¯ be a finite extension ofF. Writer for the degree ofK/F andf for its residuel degree. We have the estimations:

ρr(K) kDKk =

11

q

·X

m|f

µ f

m

qm−f + O 1

qf

forn2r1, ρn(K) kDKk =X

m|f

µ f

m

qm−f + O 1

qf

whereµdenotes the Moebius function and the constants hidden in theO()are absolute.

The dominant term in the two sums above is the summand corresponding tom = f and is equal to 1. Hence, ρn(K) is roughly equal to kDKk for n = r or n 2r1.

More precisely, one findsρn(K) =kDKk ·(1 +O(q−1))in both cases. It turns out that this conclusion continues to holds for allnrthanks to the monotony property of Theorem B.

One can also sum up the estimations of Theorem D over all extensions of a fixed degree.

Doing so, we obtain the following theorem.

Theorem E. For any positive integersrandnwithn2r1, we have the estimation:

X

K∈Exr

ρn(K) =X

m|r

ϕr m

qm−r + O

r·log logr qr

whereExrdenotes the set of all extensions ofF of degreerinsideF¯andϕis the Euler’s totient function.

Again, the dominant term in the sum of Theorem E corresponds tom=rand its value is1. Therefore, we conclude that a random polynomial of degreenhas, on average, one root in the ground field F, one more root in the union of extensions of degree 2, one more root in the union of extensions of degree3,etc. until the degree nwhere all roots have been found. Many variations on this theme are possible; for example, one can prove

(6)

that all roots of a random polynomial lie in the maximal unramified extension ofF expect

2

q +O q12

of them. On the other hand, we deduce from Theorem C that the quadratic totally ramified extensions ofF contain 2q+O q12

roots outsideF. We then conclude that there is no more thanO p12

new roots in ramified extensions of degree at least3.

On a different note, it is also quite instructive to study the fluctuations of the density functionsρK,n. Theorem B indicates that they are governed by the size of theOF-algebra generated byx. As a consequence, we deduce that elements which generate a large ex- tensionK but are close for thep-adic distance to a strict subfield ofK have less chance to show up as a root of a random polynomial. In other words, if a rootx of a random polynomial is congruent to an element of a given extensionK modulo a large power of p, it is very likely thatx actually lies inK. In some sense, subfields attract all roots in a neighborhood.

Beyond the mean, it is important to understand higher moments of theZU,n’s to draw a more precise picture of the behaviour of these random variables. We address this question by enlarging a bit our setting: instead of restricting ourselves to finite extensions ofF, we consider more generally products of such extensions,i.e. finite ´etale algebras over F. The nice observation is that Theorem A admits a straightforward generalization to this extended framework. Applying it withE =Kr (for some given finite extensionK ofF) provides information about ther-th moment ofZK,dand, more generally, sheds some light on the distribution ofr-tuples of roots inKr. ForK =F andr = 2, this yoga has already interesting consequences as it permits to compute the covariances between theZU,n’s.

Theorem F. LetU andV be two balls in OF that do not meet. Picku U andv V. We have:

Cov ZU,n, ZV,n

E[ZU,n]·E[ZV,n] = 1 + (q+ 1)2

q2+q+ 1·kuvk − q

q2+q+ 1·kuvk4 for alln3.

Although the above formula might look unattractive at first glance, it is quite in- structive. Indeed, to begin with, it indicates that Cov ZU,n, ZV,n

vanishes if and only ifkuvk= 1. In other words, the random variablesZU,nandZV,n are uncorrelated if and only ifU andV are sufficiently far away. Otherwise,ZU,nandZV,n are correlated and the covariance is always negative (still assuming thatU V =). Moreover, the correlation gets more and more significant whenU andV gets closer. This tends to show that roots repel each other. This conclusion can been understood as a consequence of the general principle that subalgebras attract roots; indeed, noticing that F embeds diagonally into F2, the above principle tells us that if we are given two nearby roots in F of a random polynomial, there is a huge chance that those roots actually coincide, which exactly means that it is unlikely to get nearby distinct roots.

Another amazing benefit of working with ´etale extensions is the existence of mass formulas for the density functions in the spirit of Bhargava’s extension to ´etale algebras of the classical Serre mass formula [2]. Given a finite ´etaleF-algebraE, defineρn(E)by the integral of Eq. (1) and let AutF-alg(E)denote the group ofF-automorphisms ofE.

Theorem G. For any positive integersrandnwithnr, we have:

X

E∈´Etr

ρn(E)

#AutF-alg(E) = 1 (2)

where the summation set´Etr consists of all isomorphism classes of ´etale extensionsEofF of degreer(and the notation#refers to the cardinality).

(7)

Whenr = 1, Eq. (2) reduces to ρn(F) = 1 and so asserts that a random polynomial of degree at least1 has exactly one root in F on average; we then recover Lerario and Kulkani’s result in this case. When r grows, Theorem G roughly says that the above remarkable property continues to hold if we count (weighted) roots in extensions of a fixed degree provided that we pay attention to include all ´etale algebras, and not only fields! Notice however that Theorem D shows that the contribution of actual extensions to the sum in Eq. (2) is about1/r. The most significant part of the mass then comes from nontrivial products of smaller degree extensions.

Organization of the article. The plan of the article follows closely the progression of the introduction. In Section 1, we prove Theorems A and B. In Section 2, we study examples and obtain closed formulas for the density functions in several simple cases. In addition of treating completely the case of quadratic extension (in line with Theorem C), we obtain partial results for extensions of prime degrees and for unramified extensions. Section 3 is devoted to finding estimations of orders of magnitude of the density functions and their integrals; we notably prove Theorems D and E there. Finally, in Section 4, we present the setup of ´etale algebras and extend Theorems A and B to this setting. We then dis- cuss applications to higher moments and mass formulas for density functions, establishing Theorems F and G.

Notations. Throughout the article, we fix a prime numberp, a finite extensionF of Qp and an algebraic closureF¯ofF. We use the letterqto denote the cardinality of the residue field ofF. We writek · kfor thep-adic norm onF, normalized by¯ kpk=p−[F:Qp].

We let n be the space of polynomials of degree at most n with coefficients in OF; we call µn the probability measure on n corresponding to λ⊗n+1F under the canonical identificationn≃ On+1F . In a slight abuse of notations, we continue to writek · kfor the norm onncorresponding to the sup norm onOFn+1(it is the so-calledGauss norm).

Throughout the article, all finite extensions of F are implicitely supposed to be con- tained in F¯. If K is such an extension, we denote by OK its ring of integers and by O×K the group of invertible elements of OK. We letλK be the Haar measure onK nor- malized byλK(OK) = 1. Our normalization choices lead to the transformation formula λK(aH) =kakr·λK(H)whereris the degree of the extensionK/F.

Finally, we use the notation#Ato denote the cardinality of a setA.

1 Density functions

The aim of this section is to define the density functionsρK,n and to prove Theorems A and B. The main ingredient we shall need is a p-adic version of the famous Kac-Rice formula which gives an integral expression for a number of roots of a polynomial. We will establish it in §1.1. In §1.2, we carry out a key computation which will allow us to construct the density functions and prove Theorem B in §1.3. We finally move to the computation of expected number of roots and prove Theorem A is §1.4.

1.1 Thep-adic Kac-Rice formula

Ap-adic version of the Kac-Rice formula already appears in the pioneer work of Evans [8].

Nevertheless, for the purpose of this article, it will be more convenient to use a different formulation from that of Evans (the latter being actually closer to what we usually call the

(8)

“area formula”). For this reason, we prefer taking some time to establish our version of thep-adic Kac-Rice formula and giving a complete proof of it. We refer to [17, Chapter 5]

for the definition of strictly differentiable functions of thep-adic variable.

Theorem 1.1. LetK be a finite extension ofF of degreer. LetU be a compact open subset ofK and letf :U K be a strictly differentiable function. We assume that(f(x), f(x))6= (0,0) for allxU. Then:

#f−1(0) = lim

s→∞ qsr· Z

Ukf(x)kr·1{kf(x)k≤q−s}dx. (3) Proof. Throughout the proof, we denote by Bs the closed ball of K of radius q−s and center 0. If π denotes a uniformizer of K, the set Bs can be alternatively defined by Bs=πsOK. We deduce from the latter equality thatλK(Bs) =kπksr =q−sr.

We consider an element a U such that f(a) = 0. From our assumption, we know that f(a) 6= 0. Therefore, applying [6, Lemma 3.4], we get the existence of a positive integerSa having the following property: for any integers Sa, the functionf induces a bijection from a+f(a)−1Bs to Bs. Up to enlargingSa, we can further assume that kf(x)k=kf(a)kfor allxBSa. We deduce for these two facts that:

Z

a+BSakf(x)kr·1{kf(x)k≤q−s}dx=kf(a)kr·λK f(a)−1Bs

=q−sr (4) for allsSa.

From the previous discussion, we also derive that a is the unique zero of f in BSa. In other words, the set of zeros off is discrete. By compacity, it follows thatf has only finitely many zeros inU. Let us call thema1, . . . , am. SetS= max(Sa1, . . . , Sam)and, for i∈ {1, . . . , m}, writeUi =ai+BSai. Up to enlarging again the Sai’s, we can assume that theUi’s are pairwise disjoint. Summing up the equalities (4), we find:

m

X

i=1

Z

Ui

kf(x)kr·1{kf(x)k≤q−s}dx=m·q−sr (5) provided thatsS. Let V be the complement in U ofU1⊔ · · · ⊔Um. It is compact and the functionf does not vanish on it. Hence, if sis large enough, we havekf(x)k > q−s for allxV. For thoses, we thus get:

Z

V kf(x)kr·1{kf(x)k≤q−s}dx= 0. (6) Combining Eqs. (5) and (6), we find that the equality:

qsr· Z

Ukf(x)kr·1{kf(x)k≤q−s}dx=m

holds true whensis sufficiently large. Passing to the limit, we get the theorem.

We underline that the compacity assumption in Theorem 1.1 cannot be relaxed. For example, taking simplyU =Zp\{0}andf :x7→x, one sees thatfhave no zero inUwhile the right hand side of Eq. (3) converges to1. Roughly speaking, the integral continues to see the missing zero at the origin, which is expected because removing one point from the domain of integration does not alter the value of the integral.

Similarly, the assumption that the zeros offare nondegenerate (i.e.that the derivative does not vanish at these points) is definitely necessary. For example, if we take the function

(9)

f : Zp Qp, x 7→ x2, a simple calculation shows that the right hand side of Eq. (3) converges to p+1p <1. More generally, one can prove that, iff is a polynomial whose roots inU area1, . . . , am and have multiplicityµ1, . . . , µmrespectively, then the right hand side of Eq. (3) converges to:

m

X

i=1

qµiqµi−1 qµi1 .

In other words, a root of multiplicityµdoes not contribute for1but for qµq−qµ−1µ−1 <1.

1.2 A key computation

LetKbe a finite ofF of degreerand letU be an open subset ofK. We aim at computing the expected value of random variableZU,n: ΩnZ∪ {+∞}defined by:

ZU,n(P) = #

xU s.t.f(x) = 0

= lim

s→∞ qsr· Z

UkP(x)kr·1{kP(x)k≤q−s}dx

the second equality coming from Theorem 1.1. For this, roughly speaking, we would like to write down the following calculation:

E[ZU,n] = Z

n

ZU,n(P)dP = Z

n

s→∞lim qsr· Z

UkP(x)kr·1{kP(x)k≤q−s}dx dP

= Z

U

s→∞lim qsr· Z

n

kP(x)kr·1{kP(x)k≤q−s}dP dx

and introduce the density function defined by:

x 7→ lim

s→∞ qsr· Z

n

kP(x)kr·1{kP(x)k≤q−s}dP. (7) However, we have to be a bit more careful because the above limit does not behave quite well everywhere: it takes infinite values on certain subspaces but it turns out that those parts lead to a finite positive contribution when we integrate. The next proposition shows that these issues are somehow localized on strict subfields.

Proposition 1.2. Ifnr and ifx lies inOK and generatesK overF, the limit in Eq.(7) exists and is equal to:

kDKk

# OK/OF[x] · Z

n−r

kQ(x)krdQ whereDK denotes the discriminant of the extensionK/F. Proof. For simplicity, write:

Is=qsr· Z

n

kP(x)kr·1{kP(x)k≤q−s}dP.

LetZ be the minimal monic polynomial of xover F. By our assumptions,Z has degree r and coefficients in OF. This implies that the mapn−r×r−1 n taking(Q, R) to

(10)

QZ+Rpreserves the measure. Performing the corresponding change of variables, we end up with the equality:

Is=qsr· Z

r−1

Z

nr

kQ(x)Z(x) +R(x)kr·1{kR(x)k≤q−s}dQ dR

=qsr· Z

r−1

Z

n−r

kQ(x)Z(x) +R(x)krdQ

1{kR(x)k≤q−s}dR

We consider the evaluation morphism αx : F OF r−1 K taking a polynomial R to R(x). It is F-linear and bijective since the domain of αx is restricted to polynomials of degree strictly less thanr. Its inverseα−1x isF-linear and so, it is continuous. Thus, there exists a positive constantγ such thatkR(x)k ≥γ· kRkfor allRr−1.

Let us assume for a moment that we are given a polynomial R r−1 such that kR(x)k ≤ q−s. By what precedes, we find that kRk ≤ γ−1q−s, from what we further deduce thatkR(x)k ≤γ−1q−s. SinceZ(x)does not vanish, we conclude thatZ(x)must divideR(x)provided thatsis large enough. One can then perform the change of variables Q7→QRZ(x)(x) in the inner integral and get:

Z

n−r

kQ(x)Z(x) +R(x)krdQ= Z

n−r

kQ(x)Z(x)krdQ

=kZ(x)kr· Z

n−r

kQ(x)krdQ.

We are then left with:

Is=qsr· kZ(x)kr· Z

n−r

kQ(x)krdQ · Z

r−1

1{kR(x)k≤q−s}dR. (8) In order to estimate the last factor, we come back to the evaluation morphismαx. Since it is aF-linear isomorphism, it must act on the measures by multiplication by some scalar (namely, its determinant). In other words, there exists a positive constant δ such that

λKx(H)) =δ·µn(H)for all measurable subsetHofr−1. TakingH = Ωr−1, we find δ=λK αx(Ωr−1)

=λK OF[x]

= 1

# OK/OF[x].

As in the proof of Theorem 1.1, we letBs be the closed ball of K of radius q−s centered at0. By definition α−1x (Bs)consists of polynomialsRsuch thatkR(x)k ≤q−s. Moreover, ifsis sufficiently large,α−1x (Bs)sits insider−1, and so:

Z

r−1

1{kR(x)k≤q−s}dR=µn α−1x (Bs)

= # OK/OF[x]

·λK(Bs) = # OK/OF[x]

·q−sr. Plugging this input in Eq. (8), we end up with:

Is=kZ(x)kr·# OK/OF[x]

· Z

n−r

kQ(x)krdQ (9) when s is sufficiently large. The sequence (Is)s≥0 is then eventually constant and con- verges to the limit given by the above formula. In order to conclude the proof of the proposition, it remains to relate the norm of Z(x) with that of the discriminant of K.

(11)

We endowK with the symmetric F-bilinear formb : K×K F, (u, v) 7→ TrK/F(uv).

Let also Bx = (1, x, . . . , xr−1) be the canonical basis of OF[x] over OF and set Bx =

xr−1

Z(x),Zxr−2(x), . . . ,Z1(x)

. BothBx andBx areF-basis of K and it follows from [19, §III.6, Lemma 2] that the matrix ofbin the basisBxandBx is lower-triangular with all diagonal entries equal to1. Its determinant is then1as well. Performing a change of basis, we find that:

detMatBx(b) =±NK/F Z(x)

where, by definition, MatBx(b) is the matrix of bin Bx and NK/F is the norm of K over F. We now consider a basis BofOK overOF. By definition, the discriminant ofK is the determinant of MatB(b). Therefore, if P denotes the transition matrix fromB to Bx, we derive from the change-of-basis formula thatdetMatBx(b) = (detP)2·DK and so:

NK/F Z(x)

=±(detP)2·DK.

On the other hand, noticing that P is also the matrix of αx in the canonical basis, we deduce that:

kdetPk=λK OF[x]

= 1

# OK/OF[x]. Combining the two previous equalities, we end up with:

kZ(x)kr =kNK/F Z(x)

k= 1

# OK/OF[x]2 · kDKk.

Plugging this relation in Eq. (9), we obtain the proposition.

1.3 Construction and properties of the density functions

In this subsection, we construct the density functionsρK,n and establish Theorem B.

Definition 1.3. Letnbe a positive integer andKbe a finite extension ofF of degreer.

Forx∈ OK, we set:

ρK,n(x) = kDKk

# OK/OF[x] · Z

n−r

kQ(x)krdQ ifnr andF[x] =K,

= 0 otherwise.

ForxK,x6∈ OK, we setρK,n(x) =kxk−2r·ρK,n(x−1).

The first part of Definition 1.3 is exactly what we expect after Proposition 1.2. As for the second part, it is motivated by the observation that the transformationP(X) 7→

XnP(X−1)preserves the measures onnand changes a rootxintox−1. In any case, we notice that, sinceKis a discrete valuation field, we havex ∈ OK orx−1 ∈ OK for allx K. Defintion 1.3 then makes sense and leads to a well-defined functionρK,n :K R+, which is called thedensity functiononK of degreen.

The rest of this subsection is devoted to the proof of Theorem B. The vanishing property and the invariance under isomorphisms (Statements 1 and 3 respectively) are clear from the definitions. In what follows, we address the other items of Theorem B one by one (in a slightly different order).

Références

Documents relatifs

The contributions include the concept of position-restricted seeds, particu- larly suitable for short alignments with non-uniform error distribution; a model that captures

Microstructure of a typical cell stack (from left to right: fuel electrode, electrolyte, barrier layer, and oxygen electrode); (a) a typical cross-section, (b) cross-section

Making the method of adaptive precision work within the framework of relaxed arithmetic is possible but tedious; for simplicity, we shall omit these technicalities and

The previous Section 3 shows that the good notion of p-adic regulator comes from the expression of the p-adic finite group R K associated with the class field theory interpretation

Some results about the p-adic Leopoldt transform of [1] and Proposition 1 will allow us to reduce the proof of Theorem 1 to the computation of the rank of a certain matrix whose

Bruno Angles. On the p-adic Leopoldt Transform of a power series.. An element of F p [[T ]] will be called a pseudo-rational function if it is the quotient of two

Ensuite, nous ferons une étude analytique et descriptive des épidémies de fièvre hémorragique dues au virus Ebola en RDC, pays où a été découvert pour la première fois le virus

If there is a periodic orbit of length 4, there must be a starting splitting cycle of length 4. This is possible only in the following case: at the first level, f 1 admit a