This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Moments of non-normal number fields - II

Krishnarjun Krishnamoorthy [email protected], [email protected] Yanqi Lake Beijing Institute of Mathemtical Sciences and Applications (BIMSA), No. 544, Hefangkou Village, Huaibei Town, Huairou District, Beijing.
Abstract.

Suppose KK is a number field and aK(m)a_{K}(m) is the number of integral ideals of norm equal to mm in KK, then for any integer ll, we asymptotically evaluate the sum

mTaKl(m)\sum_{m\leqslant T}a_{K}^{l}(m)

as TT\to\infty. We also consider the moments of the corresponding Dedekind zeta function. We prove lower bounds of expected order of magnitude and slightly improve the known upper bound for the second moment in the non-Galois case.

Key words and phrases:
Moments, Dedekind zeta function, Artin LL functions, Moments
2020 Mathematics Subject Classification:
11F66, 11F30, 11R42, 20C30.

1. Introduction

Suppose that KK is a field extension of degree dd over \mathbb{Q} (that is, a number field). Let aK(m)a_{K}(m) (for mm\in\mathbb{N}) denote the number of integral ideals in KK of norm equal to mm. Let ss be a complex number. The Dedekind zeta function of KK can be expressed as

(1.1) ζK(s):=m=1aK(m)ms.\zeta_{K}(s):=\sum_{m=1}^{\infty}\frac{a_{K}(m)}{m^{s}}.

It can be shown that aK(m)a_{K}(m) is a multiplicative function and satisfies the bound

(1.2) aK(m)ϵmϵa_{K}(m)\ll_{\epsilon}m^{\epsilon}

for any positive ϵ\epsilon. Thus the series (1.1) converges absolutely in the half plane (s)>1\Re(s)>1 where it can be expressed as the following Euler product,

(1.3) ζK(s)=p(1+aK(p)ps+aK(p2)p2s+).\zeta_{K}(s)=\prod_{p}\left(1+\frac{a_{K}(p)}{p^{s}}+\frac{a_{K}(p^{2})}{p^{2s}}+\ldots\right).

Furthermore ζK(s)\zeta_{K}(s) has a meromorphic continuation to the whole complex plane with a simple pole at s=1s=1 and satisfies a functional equation connecting values at ss and 1s1-s (see [Neu99, §5, Chapter VII]).

Dedekind zeta functions are natural generalizations of the Riemann zeta function for number fields. As with the Riemann zeta function, the behavior of ζK(s)\zeta_{K}(s) inside the critical strip 0<(s)<10<\Re(s)<1 is quite mysterious. There are many aspects of the behavior of Dedekind zeta function inside the critical strip that are of interest. In this paper, we focus on understanding the “moments” along the critical line (that is (s)=12\Re(s)=\frac{1}{2})

(1.4) IK(l)(T):=1T|ζK(12+it)|l𝑑t.I_{K}^{(l)}(T):=\int\limits_{1}^{T}\left|\zeta_{K}\left(\frac{1}{2}+it\right)\right|^{l}dt.

Obtaining precise asymptotic for the above integral is a very hard problem, and even the base case of K=K=\mathbb{Q} poses serious difficulties. Thus we turn to the discrete analogue of the above problem, which maybe more accessible. Namely, we ask if we can estimate

(1.5) MK(l)(T):=mTaKl(m)M_{K}^{(l)}(T):=\sum_{m\leqslant T}a_{K}^{l}(m)

for positive integral values of ll. The case when l=1l=1 is classical and maybe deduced as a consequence of the meromorphic continuation of ζK(s)\zeta_{K}(s). This is also analogous to obtaining estimates for averages of the higher order divisor function (often called the Piltz divisor problem). This problem was first considered by Chandrasekharan and Narasimhan for the case when KK was Galois over \mathbb{Q} and when l=2l=2 [CN63] and was generalized by Chandrasekharan and Good for arbitrary ll [CG83] (but still when KK was Galois over \mathbb{Q}). A particular non-Galois case was settled by Fomenko [Fom07] and later improved upon by Lü [L1̈3]. By different methods, this was further generalized in a recent work of the author (along with Kalyan Chakraborty) for many families of non-Galois number fields [CK22]. However, the general problem still remained unsolved. The purpose of this paper is to estimate MKl(T)M_{K}^{l}(T) for any number field KK and any positive integer ll thereby completing the solution to this problem. We also provide a unified treatment which reproduces many of the special cases treated in previous work. The final result conforms to expectations in that the main term is of the order TT times a power of log(T)\log(T).

Before we state the main theorem, we introduce and fix the following notation throughout the paper. Every representation that we consider will be over \mathbb{C}. Let KK be as above and LL be its Galois closure. The degree of KK shall be denoted by dd. Denote the Galois groups Gal(L/)Gal(L/\mathbb{Q}) as GG and its subgroup Gal(L/K)Gal(L/K) as HH. Let 1H1_{H} denote the trivial representation of HH. Denote the corresponding induction to GG as ρH\rho_{H} and its character as χρH\chi_{\rho_{H}}. We now have the following theorem.

Theorem 1.

Suppose that ll is a natural number. There exists an integer 𝗆l\mathsf{m}_{l} such that

(1.6) mTaKl(m)c(l,K)Tlog𝗆l(T)\sum_{m\leqslant T}a_{K}^{l}(m)\sim c(l,K)T\log^{\mathsf{m}_{l}}(T)

for some constant c(l,K)c(l,K) depending on ll and KK, as TT\to\infty. Moreover, we have

𝗆l=(1|G|gGχρHl(g))1.\mathsf{m}_{l}=\left(\frac{1}{|G|}\sum_{g\in G}\chi^{l}_{\rho_{H}}(g)\right)-1.

As we have mentioned before, it is of interest to find asymptotics for IK(l)(T)I_{K}^{(l)}(T) and often such estimates are tied to the estimates for MK(l)(T)M_{K}^{(l)}(T). When KK is Galois over \mathbb{Q} (equivalently HH is the trivial subgroup), lower bounds on IK(l)(T)I_{K}^{(l)}(T) of the expected order of magnitude are known unconditionally [AF12]. Conditionally on the generalized Riemann hypothesis, upper bounds of the correct order of magnitude (except for an ϵ\epsilon) are also known [MTB14]. Below, we provide analogous lower bounds for the non-Galois case. It would be convenient to define

(1.7) βK:=|HG/H|.\beta_{K}:=\left|H\setminus G/H\right|.
Theorem 2.

For any rational k0k\geqslant 0, we have

1T|ζK(12+it)|2kTlogβKk2(T).\int\limits_{1}^{T}\left|\zeta_{K}\left(\frac{1}{2}+it\right)\right|^{2k}\gg T\log^{\beta_{K}k^{2}}(T).

The above lower bound is of the expected order of magnitude. To see this we briefly recall some definitions regarding the Selberg class [Sel92] and direct the reader to [Mur94] for more details. Define the class 𝒮\mathcal{S} to consist of Dirichlet series F(s):=n=1aF(n)nsF(s):=\sum_{n=1}^{\infty}\frac{a_{F}(n)}{n^{s}} which satisfy the following properties:

  1. (1)

    (Region of convergence) The series defining F(s)F(s) converges absolutely for (s)>1\Re(s)>1.

  2. (2)

    (Analytic continuation) F(s)F(s) extends to a meromorphic function so that for some integer m0m\geqslant 0, (s1)mF(s)(s-1)^{m}F(s) is an entire function of finite order.

  3. (3)

    (Functional equation) There are numbers Q>0,αi>0,(ri)0Q>0,\alpha_{i}>0,\Re(r_{i})\geqslant 0 such that

    Φ(s):=Qsi=1dΓ(αis+ri)F(s)\Phi(s):=Q^{s}\prod_{i=1}^{d}\Gamma(\alpha_{i}s+r_{i})F(s)

    satisfies Φ(s)=wΦ(1s¯)¯\Phi(s)=w\overline{\Phi(1-\overline{s})} for some complex number ww with |w|=1|w|=1.

  4. (4)

    (Euler product) F(s)F(s) can be written as the product pFp(s)\prod_{p}F_{p}(s) where Fp(s)=exp(k=1bpk/pks)F_{p}(s)=\exp\left(\sum_{k=1}^{\infty}b_{p^{k}}/p^{ks}\right) where bpk=𝒪(pkθ)b_{p^{k}}=\mathcal{O}(p^{k\theta}) for some θ<1/2\theta<1/2.

  5. (5)

    (Ramanujan hypothesis) aF(n)=𝒪(nϵ)a_{F}(n)=\mathcal{O}(n^{\epsilon}) for any fixed ϵ>0\epsilon>0.

A function F𝒮F\in\mathcal{S} is called primitive if FF cannot be written as a product of any two elements of 𝒮\mathcal{S} except for F=1FF=1\cdot F. Selberg made the following conjectures about the elements in 𝒮\mathcal{S}.

Conjecture (Conjecture A).

For all F𝒮F\in\mathcal{S}, there exists a positive integer nFn_{F} such that

pX|aF(p)|2p=nFloglog(X)+𝒪(1).\sum_{p\leqslant X}\frac{\left|a_{F}(p)\right|^{2}}{p}=n_{F}\log\log(X)+\mathcal{O}(1).
Conjecture (Conjecture B).
  1. (1)

    For any primitive function FF, nF=1n_{F}=1.

  2. (2)

    For two distinct primitive functions F,FF,F^{\prime},

    pTaF(p)aF(p)¯p=𝒪(1).\sum_{p\leqslant T}\frac{a_{F}(p)\overline{a_{F^{\prime}}(p)}}{p}=\mathcal{O}(1).

It is expected that for an irreducible representation ξ\xi of the Galois group GG, the Artin LL function L(s,ξ)L(s,\xi) is a primitive element of the Selberg class. If {ξi}\{\xi_{i}\} is a complete list of irreducible representations of GG, then ζK(s)=iL(s,ξi)ei\zeta_{K}(s)=\prod_{i}L(s,\xi_{i})^{e_{i}} is a decomposition of ζK(s)\zeta_{K}(s) into primitive elements (inside the Selberg class), where we have set ei:=ρH,ξiGe_{i}:=\langle\rho_{H},\xi_{i}\rangle_{G}. In this situation [Hea21, Conjecture 5] along with Lemmas 4 and 7 leads to the conjecture

(1.8) 1T|ζK(12+it)|2k𝑑tc(k,K)TlogβKk2(T),\int\limits_{1}^{T}\left|\zeta_{K}\left(\frac{1}{2}+it\right)\right|^{2k}dt\sim c(k,K)T\log^{\beta_{K}k^{2}}(T),

for k>0k>0 and some constant c(k,K)c(k,K) depending on kk and KK. Finally, if KK is Galois over \mathbb{Q}, HH will be the trivial subgroup and βK=d\beta_{K}=d. The only case of (1.8) known to be true is when KK is a quadratic extension of \mathbb{Q} and when k=1k=1 [Mot70].

Regarding upper bounds for the moments, very little is known unconditionally. From their approximate functional equation, Chandrasekharan and Narasimhan ([CN63, see pg. 61]) were able to deduce that

(1.9) 1T1T|ζK(12+it)|2𝑑t=mcTd2aK2(m)m+𝒪(Td21logdT)=𝒪(Td21logdT)\frac{1}{T}\int\limits_{1}^{T}\left|\zeta_{K}\left(\frac{1}{2}+it\right)\right|^{2}dt=\sum_{m\leqslant cT^{\frac{d}{2}}}\frac{a_{K}^{2}(m)}{m}+\mathcal{O}\left(T^{\frac{d}{2}-1}\log^{d}T\right)=\mathcal{O}\left(T^{\frac{d}{2}-1}\log^{d}T\right)

for some constant c>0c>0 and d>2d>2. As a consequence of Theorem 1, we may improve this as follows.

Theorem 3.

With notation as above, we have

(1.10) 1T1T|ζK(12+it)|2𝑑t=mcTd2aK2(m)m+𝒪(Td21logβKT).\frac{1}{T}\int\limits_{1}^{T}\left|\zeta_{K}\left(\frac{1}{2}+it\right)\right|^{2}dt=\sum_{m\leqslant cT^{\frac{d}{2}}}\frac{a_{K}^{2}(m)}{m}+\mathcal{O}\left(T^{\frac{d}{2}-1}\log^{\beta_{K}}T\right).

In particular,

(1.11) 1T1T|ζK(12+it)|2𝑑t=𝒪(Td21logβKT)\frac{1}{T}\int\limits_{1}^{T}\left|\zeta_{K}\left(\frac{1}{2}+it\right)\right|^{2}dt=\mathcal{O}\left(T^{\frac{d}{2}-1}\log^{\beta_{K}}T\right)

whenever d>2d>2.

The fact that this is indeed an improvement follows from Lemma 6.

2. Preliminaries

For the convenience of the reader, we compile some basic facts which we shall use throughout the proofs.

2.1. Character theory

Given an nn dimensional complex representation ξ\xi of a finite group GG, we denote its character (trace) as χξ\chi_{\xi}. The characters associated to irreducible representations of GG form an orthonormal basis for the class functions on GG with the inner product defined as

(2.1) f1,f2G:=1|G|gGf1(g)f2(g)¯.\langle f_{1},f_{2}\rangle_{G}:=\frac{1}{|G|}\sum_{g\in G}f_{1}(g)\overline{f_{2}(g)}.

Given two representations (ξ1,V1)(\xi_{1},V_{1}) and (ξ2,V2)(\xi_{2},V_{2}) of a group GG, we may consider the representation (ξ1ξ2,V1V2)(\xi_{1}\otimes\xi_{2},V_{1}\otimes V_{2}) defined as (ξ1ξ2)(g)(v1v2)=ξ1(g)(v1)ξ2(g)(v2)(\xi_{1}\otimes\xi_{2})(g)(v_{1}\otimes v_{2})=\xi_{1}(g)(v_{1})\otimes\xi_{2}(g)(v_{2}) for any v1V1v_{1}\in V_{1} and v2V2v_{2}\in V_{2} and extended linearly. This is well-defined and satisfies

(2.2) χξ1ξ2(g)=χξ1(g)χξ2(g)\chi_{\xi_{1}\otimes\xi_{2}}(g)=\chi_{\xi_{1}}(g)\cdot\chi_{\xi_{2}}(g)

for any gGg\in G. In general the tensor product of two irreducible representations is not irreducible. Understanding the decomposition of tensor products of representations into irreducibles is often referred to as the Clebsch-Gordon problem.

2.2. Artin LL functions

We start with a number field LL, which we shall assume is Galois over \mathbb{Q}. Suppose G=Gal(L/)G=Gal(L/\mathbb{Q}). Let vv (associated with the rational prime pp) denote a finite place of \mathbb{Q} and let ww be a place of LL above vv. Let GwG_{w} and IwI_{w} denote the corresponding decomposition and inertia subgroups. We may define an element σw\sigma_{w} of Gw/IwG_{w}/I_{w} called the Frobenius element at ww. Except for finitely many places vv, the inertia subgroup IwI_{w} is trivial and thus in those cases σw\sigma_{w} is an element of the Galois group GG. In any case, as ww runs through the places over vv, the corresponding Frobenius elements (defined modulo inertia) are conjugates of one another. Thus, by abuse of notation, we shall consider the Frobenius at vv (or pp) and denote it by σv\sigma_{v} (or σp\sigma_{p}). This is justified because we shall be primarily interested in functions of σw\sigma_{w} which are invariant under conjugation (such as trace).

Suppose that ξ:GAut(V)\xi:G\to Aut(V) is a representation over a (finite dimensional) complex vector space VV. For every ww, ξ\xi maybe considered as a representation of GwG_{w} on VV and thus yields a representation of Gw/IwG_{w}/I_{w} on the fixed subspace VIwV^{I_{w}}. The Artin LL function attached to the representation ξ\xi is defined by

(2.3) L(s,ξ):=v<1det((Idpsξ(σw))|VIw)=:pLp(ξ,s).L(s,\xi):=\prod_{v<\infty}\frac{1}{\det\left(\left(Id-p^{-s}\xi(\sigma_{w}))\right|V^{I_{w}}\right)}=:\prod_{p}L_{p}(\xi,s).

The product is absolutely convergent for (s)>1\Re(s)>1. We collect some of the important properties of the Artin LL function for future reference.

Proposition 1.

The Artin LL functions defined above has the following properties.

  1. (1)

    L(s,ξ1ξ2)=L(s,ξ1)L(s,ξ2)L(s,\xi_{1}\oplus\xi_{2})=L(s,\xi_{1})L(s,\xi_{2}) for any two representations ξ1\xi_{1} and ξ2\xi_{2}.

  2. (2)

    Suppose KL\mathbb{Q}\subset K\subset L is an intermediate field which is Galois over \mathbb{Q}. Let H=Gal(L/K)H=\mbox{Gal}\left(L/K\right). Then a representation of ξ\xi of G/HG/H may be lifted to a representation ξ~\tilde{\xi} via the canonical projection GG/HG\to G/H. Then L(s,ξ)=L(s,ξ~)L(s,\xi)=L(s,\tilde{\xi}), where the first LL function is considered in the setting of LL over \mathbb{Q} and the second LL function is considered in the setting KK over \mathbb{Q}.

  3. (3)

    Suppose KK is an intermediate field, not necessarily Galois over \mathbb{Q}. Let HH denote the Galois group of LL over KK. For a representation ξ\xi of HH, we have L(s,ξ)=L(s,IndHGξ)L(s,\xi)=L(s,\mbox{Ind}_{H}^{G}\xi), where IndHGξ\mbox{Ind}_{H}^{G}\xi denotes the representation induced from HH to GG.

  4. (4)

    With notation as in the previous statement. Then

    ζK(s)=L(s,IndHG1H)\zeta_{K}(s)=L(s,\mbox{Ind}_{H}^{G}1_{H})

    where 1H1_{H} is the trivial representation of HH.

Remark.

In the sequel, we shall use Proposition 1 repeatedly at various steps without referring back to it every time.

For an irreducible non-trivial representation ξ\xi of GG, the Artin holomorphy conjecture asserts that L(s,ξ)L(s,\xi) continues holomorphically to the whole complex plane. This is unknown at the moment but we have a very general partial result. Brauer’s induction theorem is an important result in representation theory of groups which is particularly consequential in the study Artin LL functions. Suppose that ξ\xi is an irreducible non-trivial representation of GG. Brauer’s theorem establishes the meromorphic continuation of L(s,ξ)L(s,\xi) to the whole complex plane. The following slightly stronger consequence of Brauer’s theorem shall be useful for us in the sequel (see [IK04, Corollary 5.47]).

Lemma 1.

If ξ\xi is a non-trivial irreducible representation of GG, then L(s,ξ)L(s,\xi) has neither zeros nor poles in the region (s)1\Re(s)\geqslant 1.

3. Proof of Theorem 1

In the remainder of this paper, it is convenient, in many instances, to restrict ourselves to unramified primes. As we are omitting only finitely many primes, this does not affect the nature of our results but for the exact constants. The order of growth shall remain the same. The strategy of the proof is essentially that of [CK22] adopted and generalized with modifications for the current needs. Define

(3.1) Dl(s):=m=1aKl(m)ms=p(1+aKl(p)ps+aKl(p2)p2s).D_{l}(s):=\sum_{m=1}^{\infty}\frac{a_{K}^{l}(m)}{m^{s}}=\prod_{p}\left(1+\frac{a_{K}^{l}(p)}{p^{s}}+\frac{a_{K}^{l}(p^{2})}{p^{2s}}\ldots\right).

From (1.2), Dl(s)D_{l}(s) is absolutely convergent for (s)>1\Re(s)>1 where the Euler product expression is valid.

3.1. Meromorphic Continuation of Dl(s)D_{l}(s)

We wish to establish a meromorphic continuation of Dl(s)D_{l}(s) to a larger domain. As ζK(s)=L(s,ρH)\zeta_{K}(s)=L(s,\rho_{H}), for every unramified prime pp, we have

aK(p)=χρH(σp)a_{K}(p)=\chi_{\rho_{H}}(\sigma_{p})

where σp\sigma_{p} is a choice of Frobenius at pp. From (2.2), it follows that

aKl(p)=χρHl(σp),a_{K}^{l}(p)=\chi_{\rho_{H}^{\otimes l}}(\sigma_{p}),

where

ρHl=ρHρHρHl times.\rho_{H}^{\otimes l}=\underbrace{\rho_{H}\otimes\rho_{H}\otimes\ldots\otimes\rho_{H}}_{l\mbox{\tiny{ times}}}.

Therefore from definition,

L(s,ρHl)=E(s)p unramified(1aKl(p)ps+)1,L(s,\rho_{H}^{\otimes l})=E(s)\prod_{p\mbox{\tiny{ unramified}}}\left(1-\frac{a_{K}^{l}(p)}{p^{s}}+\ldots\right)^{-1},

where E(s)E(s) is the product of Euler factors at the ramified primes. Furthermore, for (s)>1\Re(s)>1, we have

L(s,ρHl)\displaystyle L(s,\rho_{H}^{\otimes l}) =p(1aKl(p)ps)1U1(s)\displaystyle=\prod_{p}\left(1-\frac{a_{K}^{l}(p)}{p^{s}}\right)^{-1}U_{1}(s)
(3.2) =p(1+aKl(p)ps+aK2l(p)p2s+)U1(s),\displaystyle=\prod_{p}\left(1+\frac{a_{K}^{l}(p)}{p^{s}}+\frac{a_{K}^{2l}(p)}{p^{2s}}+\ldots\right)U_{1}(s),

where U1(s)U_{1}(s) is holomorphic in the region (s)>12\Re(s)>\frac{1}{2}. Comparing Euler factors with (3.1), we see that

(3.3) Dl(s)=L(s,ρHl)U2(s)D_{l}(s)=L(s,\rho_{H}^{\otimes l})U_{2}(s)

where U2(s)U_{2}(s) is again a function holomorphic in the region (s)>12\Re(s)>\frac{1}{2}. The region of holomorphy of U1(s)U_{1}(s) and U2(s)U_{2}(s) can be deduced using (1.2). In particular, from the meromorphic continuation of L(s,ρHl)L(s,\rho_{H}^{\otimes l}) to the whole complex plane, we may conclude that Dl(s)D_{l}(s) continues meromorphically to the region (s)>12\Re(s)>\frac{1}{2}.

3.2. Completing the proof

Let {ξ}i=1n\{\xi\}_{i=1}^{n} denote the complete set of irreducible representations of GG with ξ1\xi_{1} being the trivial representation. Let

(3.4) χρHl=i=1n𝗆i(l)χξi\chi_{\rho_{H}^{\otimes l}}=\sum_{i=1}^{n}\mathsf{m}_{i}^{(l)}\chi_{\xi_{i}}

denote the decomposition of ρHl\rho_{H}^{\otimes l} into irreducible representations. Translating this into Artin LL functions gives us

L(s,ρHl)=i=1nL𝗆i(l)(s,ξi)=ζ𝗆1(l)(s)i=2nL𝗆i(l)(s,ξi).L(s,\rho_{H}^{\otimes l})=\prod_{i=1}^{n}L^{\mathsf{m}_{i}^{(l)}}(s,\xi_{i})=\zeta^{\mathsf{m}_{1}^{(l)}}(s)\prod_{i=2}^{n}L^{\mathsf{m}_{i}^{(l)}}(s,\xi_{i}).

Thus L(s,ρHl)L(s,\rho_{H}^{\otimes l}) has a pole of order 𝗆1(l)\mathsf{m}_{1}^{(l)} at the point s=1s=1 and is otherwise continuous on the half plane (s)1\Re(s)\geqslant 1 (from Lemma 1). Hence from (3.3), Dl(s)D_{l}(s) is a continuous function in the region (s)1\Re(s)\geqslant 1 but for a pole of order 𝗆1(l)\mathsf{m}_{1}^{(l)} at the point s=1s=1. Now we apply the Delange-Ikehara Tauberian theorem (see [Nar83, Corollary, Pg. 121]) to the Dirichlet series Dl(s)D_{l}(s) and get

MK(l)(T)cTlog𝗆1(l)1(T)M_{K}^{(l)}(T)\sim cT\log^{\mathsf{m}_{1}^{(l)}-1}(T)

for some constant cc. In fact, cc maybe expressed in terms of the leading coefficient in the Laurent series expansion of Dl(s)D_{l}(s) about the point s=1s=1. Finally from definition, we have

(3.5) 𝗆1(l)=ρHl,1GG=1|G|gGχρHl(g).\mathsf{m}_{1}^{(l)}=\langle\rho_{H}^{\otimes l},1_{G}\rangle_{G}=\frac{1}{|G|}\sum_{g\in G}\chi_{\rho_{H}}^{l}(g).

Setting 𝗆l=𝗆1(l)1\mathsf{m}_{l}=\mathsf{m}_{1}^{(l)}-1 completes the proof.

4. Proof of Theorem 2

We first note two applications of the Chebotarev density theorem.

Lemma 2.

For any two representations ρ1,ρ2\rho_{1},\rho_{2} of GG,

pTp unramifiedχρ1(σp)χρ2(σp)¯p=ρ1,ρ2Gloglog(T)+𝒪(1).\underset{p\mbox{ \tiny{unramified}}}{\sum_{p\leqslant T}}\frac{\chi_{\rho_{1}}(\sigma_{p})\overline{\chi_{\rho_{2}}(\sigma_{p})}}{p}=\langle\rho_{1},\rho_{2}\rangle_{G}\log\log(T)+\mathcal{O}(1).
Proof.

We shall restrict ourselves to unramified primes throughout the proof and remove it from notation. From the Chebotarev density theorem, for any conjugacy class CC of GG,

pTσpC1p=|C||G|loglog(T)+𝒪(1).\underset{\sigma_{p}\in C}{\sum_{p\leqslant T}}\frac{1}{p}=\frac{|C|}{|G|}\log\log(T)+\mathcal{O}(1).

Therefore,

pTχρ1(σp)χρ2(σp)¯p\displaystyle\sum_{p\leqslant T}\frac{\chi_{\rho_{1}}(\sigma_{p})\overline{\chi_{\rho_{2}}(\sigma_{p})}}{p} =CpTσpCχρ1(σp)χρ2(σp)¯p\displaystyle=\sum_{C}\underset{\sigma_{p}\in C}{\sum_{p\leqslant T}}\frac{\chi_{\rho_{1}}(\sigma_{p})\overline{\chi_{\rho_{2}}(\sigma_{p})}}{p}
=Cχρ1(gC)χρ2(gC)¯pTσpC1p\displaystyle=\sum_{C}\chi_{\rho_{1}}(g_{C})\overline{\chi_{\rho_{2}}(g_{C})}\underset{\sigma_{p}\in C}{\sum_{p\leqslant T}}\frac{1}{p}
=1|G|Cχρ1(gC)χρ2(gC)¯|C|loglog(T)+𝒪(1)\displaystyle=\frac{1}{|G|}\sum_{C}\chi_{\rho_{1}}(g_{C})\overline{\chi_{\rho_{2}}(g_{C})}|C|\log\log(T)+\mathcal{O}(1)
=(1|G|gGχρ1(g)χρ2(g)¯)loglog(T)+𝒪(1)\displaystyle=\left(\frac{1}{|G|}\sum_{g\in G}\chi_{\rho_{1}}(g)\overline{\chi_{\rho_{2}}(g)}\right)\log\log(T)+\mathcal{O}(1)
=ρ1,ρ2Gloglog(T)+𝒪(1).\displaystyle=\langle\rho_{1},\rho_{2}\rangle_{G}\log\log(T)+\mathcal{O}(1).

Here C\sum_{C} denotes the sum over the various conjugacy classes of GG and gCg_{C} denotes an arbitrary element in each conjugacy class. This completes the proof. ∎

Corollary 1.

With notation as above

nTaKl(p)p=(𝗆l+1)loglog(T)+𝒪(1).\sum_{n\leqslant T}\frac{a_{K}^{l}(p)}{p}=(\mathsf{m}_{l}+1)\log\log(T)+\mathcal{O}(1).
Proof.

We may replace the sum over all the primes with the sum over the unramified primes. We note that, for an unramified prime pp, aKl(p)=χρHl(σp)χ1G(σp)a_{K}^{l}(p)=\chi_{\rho_{H}^{\otimes l}}(\sigma_{p})\cdot\chi_{1_{G}}(\sigma_{p}). Therefore from the previous lemma,

pTaKl(p)p=ρHl,1GGloglog(T)+𝒪(1).\sum_{p\leqslant T}\frac{a_{K}^{l}(p)}{p}=\langle\rho_{H}^{\otimes l},1_{G}\rangle_{G}\log\log(T)+\mathcal{O}(1).

From definition 𝗆1(l)=ρHl,1GG=𝗆l+1\mathsf{m}_{1}^{(l)}=\langle\rho_{H}^{\otimes l},1_{G}\rangle_{G}=\mathsf{m}_{l}+1. This completes the proof. ∎

The proof of the following lemma and corollary follow almost verbatim to the proofs above and hence we omit the details. We use the Chebatorev density theorem in the following form

pTσpC,p unramified1|C||G|Tlog(T).\underset{\sigma_{p}\in C,\ p\mbox{ \tiny{unramified}}}{\sum_{p\leqslant T}}1\sim\frac{|C|}{|G|}\frac{T}{\log(T)}.
Lemma 3.

For any two representations ρ1,ρ2\rho_{1},\rho_{2} of GG,

pTp unramifiedχρ1(σp)χρ2(σp)¯ρ1,ρ2GTlog(T).\underset{p\mbox{ \tiny{unramified}}}{\sum_{p\leqslant T}}\chi_{\rho_{1}}(\sigma_{p})\overline{\chi_{\rho_{2}}(\sigma_{p})}\sim\langle\rho_{1},\rho_{2}\rangle_{G}\frac{T}{\log(T)}.
Corollary 2.

With notation as above

pTaKl(p)(𝗆l+1)Tlog(T)\sum_{p\leqslant T}a_{K}^{l}(p)\sim(\mathsf{m}_{l}+1)\frac{T}{\log(T)}
Lemma 4.

With notation as above

𝗆1(2)=βK.\mathsf{m}_{1}^{(2)}=\beta_{K}.
Proof.

Observe that

ρH2,1GG=1|G|gGχρH2(g)=ρH,ρH¯G.\langle\rho_{H}^{\otimes 2},1_{G}\rangle_{G}=\frac{1}{|G|}\sum_{g\in G}\chi_{\rho_{H}}^{2}(g)=\langle\rho_{H},\overline{\rho_{H}}\rangle_{G}.

From Frobenius reciprocity

ρH,ρH¯G\displaystyle\langle\rho_{H},\overline{\rho_{H}}\rangle_{G} =Ind(1H),Ind(1H)¯G\displaystyle=\langle Ind(1_{H}),\overline{Ind(1_{H})}\rangle_{G}
=1H,Res(Ind(1H)¯)H.\displaystyle=\left\langle 1_{H},Res\left(\overline{Ind(1_{H})}\right)\right\rangle_{H}.

From Mackey’s theorem, we have

Res(Ind(1H)¯)=Res(Ind(1H))¯=gHG/HIndHgH1Hg¯,Res\left(\overline{Ind(1_{H})}\right)=\overline{Res\left(Ind(1_{H})\right)}=\bigoplus_{g\in H\setminus G/H}\overline{Ind_{H_{g}}^{H}1_{H_{g}}},

where Hg:=gHg1HH_{g}:=gHg^{-1}\cap H. Therefore,

1H,Res(Ind(1H)¯)H=gHG/H1H,IndHgH1Hg¯H.\left\langle 1_{H},Res\left(\overline{Ind(1_{H})}\right)\right\rangle_{H}=\sum_{g\in H\setminus G/H}\langle 1_{H},\overline{Ind_{H_{g}}^{H}1_{H_{g}}}\rangle_{H}.

The induced representation of 1Hg1_{H_{g}} contains precisely one copy of 1H1_{H} for every gg and hence ρH2,1GG=|HG/H|=βK\langle\rho_{H}^{\otimes 2},1_{G}\rangle_{G}=\left|H\setminus G/H\right|=\beta_{K} completing the proof. ∎

Proof of Theorem 2.

Our strategy is to apply [AF12, Theorem 2.3] to the Dedekind zeta function ζK(s)\zeta_{K}(s). In their notation, ζK(s)\zeta_{K}(s) satisfies the necessary conditions of their theorem with β=𝗆2+1=𝗆1(2)\beta=\mathsf{m}_{2}+1=\mathsf{m}_{1}^{(2)}. This follows from Lemmas 2 and 3 above. From Lemma 4, 𝗆1(2)=β=βK\mathsf{m}_{1}^{(2)}=\beta=\beta_{K} and we are done. ∎

5. Proof of Theorem 3

The proof follows along the same lines as [CN63, §6], where we use Theorem 1 in place of their [CN63, Theorem 3]. The key improvement of the bound on the off-diagonal terms is given in the following lemma. First we observe from Theorem 1 and the results in the previous section, that

(5.1) mTaK2(m)TlogβK(T).\sum_{m\leqslant T}a_{K}^{2}(m)\ll T\log^{\beta_{K}}(T).
Lemma 5.

Define

S(T):=m<nTaK(m)aK(n)mnlog(nm).S(T):=\sum_{m<n\leqslant T}\frac{a_{K}(m)a_{K}(n)}{\sqrt{mn}\log\left(\frac{n}{m}\right)}.

Then S(T)=𝒪(TlogβK(T))S(T)=\mathcal{O}\left(T\log^{\beta_{K}}(T)\right).

Proof.

In the range n>mn>m, we have log(nm)>1mn\log(\frac{n}{m})>1-\frac{m}{n}. We therefore have

S(T)<m<nTaK(m)aK(n)mn+m<nTaK(m)aK(n)nm.S(T)<\sum_{m<n\leqslant T}\frac{a_{K}(m)a_{K}(n)}{\sqrt{mn}}+\sum_{m<n\leqslant T}\frac{a_{K}(m)a_{K}(n)}{n-m}.

As in [CN63, proof of Lemma 11], the first term is 𝒪(T)\mathcal{O}(T) and the second term is bounded above by mTaK2(m)\sum_{m\leqslant T}a_{K}^{2}(m). The lemma now follows from (5.1). ∎

Now we may follow [CN63, §6] almost verbatim and complete the proof of Theorem 3. The following lemma applied to G=Gal(L/)G=Gal(L/\mathbb{Q}) and H=Gal(L/K)H=Gal(L/K) implies that Theorem 3 is indeed an improvement over (1.9).

Lemma 6.

If GG is a group and HH is a subgroup of GG, then |HG/H|=|G/H||H\setminus G/H|=|G/H| if and only if HH is a normal subgroup of GG.

Proof.

Consider the action of HH on the left on the set of left cosets G/HG/H (the action of hh maps gHgH to (hg)H(hg)H). Then the set of double cosets is in bijection with the orbits under this action (HgHHgH\mapsto orbit of gHgH). It can be shown that the size of the orbit containing HgHHgH under this action is the index [H:HgHg1][H:H\cap gHg^{-1}]. If HH were normal in GG, then each of the above index is 11 and hence |HG/H|=|G/H||H\setminus G/H|=|G/H|. If HH were not normal in GG, then there exists a gGg\in G such that HgHg1H\neq gHg^{-1}. It follows that the size of the orbit containing gHgH has at least two elements and thus the number of orbits (which equals |HG/H||H\setminus G/H|) should be strictly less than the size of the set on which HH is acting (which is |G/H||G/H|). ∎

6. Further remarks

We first have the following lemma.

Lemma 7.

If

ρH=i=1nξiei\rho_{H}=\bigoplus_{i=1}^{n}\xi_{i}^{e_{i}}

is the decomposition of ρH\rho_{H} into irreducible representations, then

𝗆1(2)=iei2.\mathsf{m}_{1}^{(2)}=\sum_{i}e_{i}^{2}.
Proof.

Proceeding as in the previous proof, we have 𝗆1(2)=ρH,ρH¯G\mathsf{m}_{1}^{(2)}=\langle\rho_{H},\overline{\rho_{H}}\rangle_{G}. But

χρH(g)=1|H|rGτH(r1gr).\chi_{\rho_{H}}(g)=\frac{1}{|H|}\sum_{r\in G}\tau_{H}(r^{-1}gr)\in\mathbb{R}.

Here τH\tau_{H} is the characteristic function of HH. Hence ρH=ρH¯\rho_{H}=\overline{\rho_{H}} giving us 𝗆1(2)=ρH,ρHG=i=1nei2\mathsf{m}_{1}^{(2)}=\langle\rho_{H},\rho_{H}\rangle_{G}=\sum_{i=1}^{n}e_{i}^{2}. This completes the proof. ∎

As mentioned above, from the general conjectures regarding the primitivity of Artin LL functions, ζK(s)=i=1nL(s,ξi)ei\zeta_{K}(s)=\prod_{i=1}^{n}L(s,\xi_{i})^{e_{i}} is a decomposition of ζK\zeta_{K} into primitive elements of 𝒮\mathcal{S} (where as before {ξi}\{\xi_{i}\} is a complete set of irreducible representations of GG). If this is the case, Lemma 7 would follow from Selberg’s Conjecture B (see [Mur94, Proposition 2.5(a)] and [Mur94, Corollary 3.2]). In any case the results of this paper would follow if we assume Selberg’s conjectures. As was shown in [Mur94], Selberg’s conjectures imply the Langlands reciprocity conjecture (or the strong Artin conjecture) from where we may show the required asymptotic in Theorem 1 with a power saving error term. In fact it is expected that the right hand side of (1.6) should be of the form TPl(log(T))+𝒪(T1θ)TP_{l}(\log(T))+\mathcal{O}(T^{1-\theta}) for some polynomial PlP_{l} of degree 𝗆l\mathsf{m}_{l} and for some θ>0\theta>0 (as in the main theorem of [CK22]).

Acknowledgement

The author wishes to thank Prof. Sudhir Pujahari for his encouragement.

References

  • [AF12] A. Akbary and B. Fodden, Lower bounds for power moments of LL-functions, Acta Arith. 151 (2012), no. 1, 11–38 (English).
  • [CK22] K. Chakraborty and K. Krishnamoorthy, On moments of non-normal number fields, J. Number Theory 238 (2022), 183–196. MR 4430097
  • [CG83] K. Chandrasekharan and A. Good, On the number of integral ideals in Galois extensions, Monatsh. Math. 95 (1983), no. 2, 99–109. MR 697150
  • [CN63] K. Chandrasekharan and R. Narasimhan, The approximate functional equation for a class of zeta-functions, Math. Ann. 152 (1963), 30–64. MR 153643
  • [Fom07] O. M. Fomenko, Mean values associated with the Dedekind zeta function, Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI) 350 (2007), 187–198, 203. MR 2722976
  • [Hea21] W. Heap, Moments of the dedekind zeta function and other non-primitive l-functions, Mathematical Proceedings of the Cambridge Philosophical Society, vol. 170, Cambridge University Press, 2021, pp. 191–219.
  • [IK04] H. Iwaniec and E. Kowalski, Analytic number theory, Colloq. Publ., Am. Math. Soc., vol. 53, Providence, RI: American Mathematical Society (AMS), 2004 (English).
  • [L1̈3] G. Lü, Mean values connected with the Dedekind zeta-function of a non-normal cubic field, Cent. Eur. J. Math. 11 (2013), no. 2, 274–282. MR 3000644
  • [MTB14] Micah B. Milinovich and Caroline L. Turnage-Butterbaugh, Moments of products of automorphic LL-functions, J. Number Theory 139 (2014), 175–204 (English).
  • [Mot70] Y. Motohashi, A note on the mean value of the Dedekind zeta-function of the quadratic field, Math. Ann. 188 (1970), 123–127. MR 265302
  • [Mur94] M. Ram Murty, Selberg’s conjectures and Artin LL-functions, Bull. Am. Math. Soc., New Ser. 31 (1994), no. 1, 1–14 (English).
  • [Nar83] W. Narkiewicz, Number theory. Transl. from the Polish by S. Kanemitsu, 1983 (English).
  • [Neu99] Jürgen Neukirch, Algebraic number theory, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 322, Springer-Verlag, Berlin, 1999, Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder. MR 1697859
  • [Sel92] Atle Selberg, Old and new conjectures and results about a class of Dirichlet series, Proceedings of the Amalfi conference on analytic number theory, held at Maiori, Amalfi, Italy, from 25 to 29 September, 1989, Salerno: Universitá di Salerno, 1992, pp. 367–385 (English).