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

Local to global principle for expected values

Giacomo Micheli Department of Mathematics
University of South Florida
Tampa, FL 33620, United States of America
[email protected]
Severin Schraven Institute of Mathematics
University of Zurich
Winterthurerstrasse 190
8057 Zurich, Switzerland
[email protected]
 and  Violetta Weger Institute of Mathematics
University of Zurich
Winterthurerstrasse 190
8057 Zurich, Switzerland
[email protected]
Abstract.

This paper constructs a new local to global principle for expected values over free \mathbb{Z}-modules of finite rank. In our strategy we use the same philosophy as Ekedhal’s Sieve for densities, later extended and improved by Poonen and Stoll in their local to global principle for densities. We show that under some additional hypothesis on the system of pp-adic subsets used in the principle, one can use pp-adic measures also when one has to compute expected values (and not only densities). Moreover, we show that our additional hypotheses are sharp, in the sense that explicit counterexamples exist when any of them is missing. In particular, a system of pp-adic subsets that works in the Poonen and Stoll principle is not guaranteed to work when one is interested in expected values instead of densities. Finally, we provide both new applications of the method, and immediate proofs for known results.

Key words and phrases:
Densities, Mean.
2010 Mathematics Subject Classification:

1. Introduction

Let \mathbb{Z} be the set of integers and dd be a positive integer. The problem of computing the “probability”, that a randomly chosen element in d\mathbb{Z}^{d} has a certain property, has a long history dating back to Cesáro [2, 3].

Since no uniform probability distribution exists over \mathbb{Z}, one introduces the notion of density. The density of a set TdT\subset\mathbb{Z}^{d} is defined to be

ρ(T)=limH|T[H,H[d|(2H)d,\rho(T)=\lim_{H\rightarrow\infty}\frac{|T\cap[-H,H[^{d}|}{(2H)^{d}},

if the limit exists. Density results over d\mathbb{Z}^{d} have received a great deal of interest recently [4, 5, 6, 7, 8, 9, 10, 12, 13, 14, 16, 17, 18, 19, 22, 24].

In [20, Lemma 20] Poonen and Stoll show that the computation of densities of many sets SdS\subseteq\mathbb{Z}^{d} defined by local conditions (in the pp-adic sense) can be reduced to measuring the corresponding subsets of the pp-adic integers. This technique is an extension of Ekedahl’s Sieve [23]. An even more general result is [1, Proposition 3.2]. The general philosophy is that, under some reasonable assumptions, one should be able to treat the pp-adic measures of infinitely many sets UppdU_{p}\subseteq\mathbb{Z}_{p}^{d} independently when one looks at the density of the corresponding set p𝒫UpCd\bigcap_{p\in\mathcal{P}}U_{p}^{C}\cap\mathbb{Z}^{d} over the integers.

In [11] the authors computed the expected number of primes, for which an Eisenstein-polynomial satisfies the criterion of Eisenstein. In this paper we prove that this is a special case of a much more general principle that allows to compute expected values from pp-adic measures of nicely chosen systems of pp-adic sets. In fact, in this article we focus on extending the method by Poonen and Stoll to compute expected values of the “random variable” that counts how many times an element is expected to be in one of the UpU_{p}’s. First, we show that the hypothesis of Poonen and Stoll is not sufficient to guarantee a local to global principle for expected values (see Example 12). This led us to add two additional hypotheses on the system (Up)p𝒫(U_{p})_{p\in\mathcal{P}}, that appears in the Poonen and Stoll principle, in order to prove Theorem 13, the main theorem of this article. The additional hypotheses we require are sharp and necessary, as examples 15 and 16 show.

With this extension to the local to global principle one can for example compute for non-coprime mm-tuples of integers how many prime factors they have in common on average, or for rectangular non-unimodular matrices how many primes divide all the basic minors on average, and of course also the result of [11] follows directly.

The paper is organized as follows: in Section 2 we will recall the local to global principle by Poonen and Stoll. In Section 3 we will introduce the definition of expected value of a system of (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} and then state and prove Theorem 13, the main theorem of this paper.

In Section 4 we give some applications of our main theorem that allow very fast computations of expected values over d\mathbb{Z}^{d} (see for example [11] compared with Corollary 18).

2. Preliminaries

Definition 1.

Let dd be a positive integer. The density of a set TdT\subset\mathbb{Z}^{d} is defined to be

ρ(T)=limH|T[H,H[d|(2H)d,\rho(T)=\lim_{H\rightarrow\infty}\frac{|T\cap[-H,H[^{d}|}{(2H)^{d}},

if the limit exists. Then one defines the upper density ρ¯\bar{\rho} and the lower density ρ¯\underline{\rho} equivalently with the lim sup\limsup and the lim inf\liminf respectively.

For convenience, let us restate here the lemma of Poonen and Stoll in [20, Lemma 20]. If SS is a set, then we denote by 2S2^{S} its powerset and by SCS^{C} its complement. Let 𝒫\mathcal{P} be the set of primes and M={}𝒫M_{\mathbb{Q}}=\{\infty\}\cup\mathcal{P} be the set of all places of \mathbb{Q}, where we denote by \infty the unique archimedean place of \mathbb{Q}. Let μ\mu_{\infty} denote the Lebesgue measure on d\mathbb{R}^{d} and μp\mu_{p} the normalized Haar measure on pd\mathbb{Z}_{p}^{d}. For TT a subset of a metric space, let us denote by (T)\partial(T) its boundary, by T¯\overline{T} its closure and by TT^{\circ} its interior. By 0\mathbb{R}_{\geq 0} we denote the non-negative reals. A minor of a matrix AA is called basic, if it is the nonzero determinant of a square submatrix of AA of maximal order.

Theorem 2 ([20, Lemma 20]).

Let dd be a positive integer. Let UdU_{\infty}\subset\mathbb{R}^{d}, such that 0U=U\mathbb{R}_{\geq 0}\cdot U_{\infty}=U_{\infty} and μ((U))=0.\mu_{\infty}(\partial(U_{\infty}))=0. Let s=12dμ(U[1,1]d)s_{\infty}=\frac{1}{2^{d}}\mu_{\infty}(U_{\infty}\cap[-1,1]^{d}). For each prime pp, let UppdU_{p}\subset\mathbb{Z}_{p}^{d}, such that μp((Up))=0\mu_{p}(\partial(U_{p}))=0 and define sp=μp(Up)s_{p}=\mu_{p}(U_{p}). Define the following map

P:d\displaystyle P:\mathbb{Z}^{d} \displaystyle\rightarrow 2M,\displaystyle 2^{M_{\mathbb{Q}}},
a\displaystyle a \displaystyle\mapsto {νMaUν}.\displaystyle\left\{\nu\in M_{\mathbb{Q}}\mid a\in U_{\nu}\right\}.

If the following is satisfied:

(2.1) limMρ¯({adaUpfor some primep>M})=0,\lim_{M\rightarrow\infty}\bar{\rho}\left(\left\{a\in\mathbb{Z}^{d}\mid a\in U_{p}\ \text{for some prime}\ p>M\right\}\right)=0,

then:

  • i)

    νMsν\sum\limits_{\nu\in M_{\mathbb{Q}}}s_{\nu} converges.

  • ii)

    For 𝒮2M,\mathcal{S}\subset 2^{M_{\mathbb{Q}}}, ρ(P1(𝒮))\rho(P^{-1}(\mathcal{S})) exists, and defines a measure on 2M2^{M_{\mathbb{Q}}}.

  • iii)

    For each finite set S2MS\in 2^{M_{\mathbb{Q}}}, we have that

    ρ(P1({S}))=νSsννS(1sν),\rho(P^{-1}(\{S\}))=\prod_{\nu\in S}s_{\nu}\prod_{\nu\not\in S}(1-s_{\nu}),

    and if 𝒮\mathcal{S} consists of infinite subsets of 2M2^{M_{\mathbb{Q}}}, then ρ(P1(𝒮))=0.\rho(P^{-1}(\mathcal{S}))=0.

To show that (2.1) is satisfied, one can often apply the following useful lemma, that can be deduced from the result in [23].

Lemma 3 ([21, Lemma 2]).

Let dd and MM be positive integers. Let f,g[x1,,xd]f,g\in\mathbb{Z}[x_{1},\ldots,x_{d}] be relatively prime. Define

SM(f,g)={adf(a)g(a)0modpfor some primep>M},S_{M}(f,g)=\left\{a\in\mathbb{Z}^{d}\mid f(a)\equiv g(a)\equiv 0\mod p\ \text{for some prime}\ p>M\right\},

then

limMρ¯(SM(f,g))=0.\lim_{M\rightarrow\infty}\bar{\rho}(S_{M}(f,g))=0.

3. The local to global principle for expected values

Observe, that in Theorem 2 one could always choose the finite set SS to be the empty set, which for our purpose will be convenient.

Corollary 4.

For all νM\nu\in M_{\mathbb{Q}} let UνU_{\nu} be chosen as in Theorem 2, corresponding to a finite set S2MS\in 2^{M_{\mathbb{Q}}}. Let us define

Uν={UνCνS,UννS,U^{\prime}_{\nu}=\begin{cases}U_{\nu}^{C}&\nu\in S,\\ U_{\nu}&\nu\not\in S,\end{cases}

and hence

sν={1sννS,sννS,s^{\prime}_{\nu}=\begin{cases}1-s_{\nu}&\nu\in S,\\ s_{\nu}&\nu\not\in S,\end{cases}

and define

P:d\displaystyle P^{\prime}:\mathbb{Z}^{d} \displaystyle\rightarrow 2M,\displaystyle 2^{M_{\mathbb{Q}}},
a\displaystyle a \displaystyle\mapsto {νMaUν}.\displaystyle\left\{\nu\in M_{\mathbb{Q}}\mid a\in U^{\prime}_{\nu}\right\}.

Then we get

  • i)

    νMsν\sum\limits_{\nu\in M_{\mathbb{Q}}}s^{\prime}_{\nu} converges.

  • ii)

    For 𝒮2M,\mathcal{S}\subset 2^{M_{\mathbb{Q}}}, ρ(P1(𝒮))\rho(P^{\prime-1}(\mathcal{S})) exists and defines a measure on 2M2^{M_{\mathbb{Q}}}.

  • iii)

    ρ(P1({}))=νM(1sν)=ρ(P1({S}))\rho(P^{\prime-1}(\{\emptyset\}))=\prod\limits_{\nu\in M_{\mathbb{Q}}}(1-s^{\prime}_{\nu})=\rho(P^{-1}(\{S\})), where PP is the map as in Theorem 2.

The proof is straightforward and thus we omit the details.

For a fixed νM\nu\in M_{\mathbb{Q}} and UνU_{\nu} as in Theorem 2, the density of UνdU_{\nu}\cap\mathbb{Z}^{d} can be computed as follows.

Corollary 5.

Let νM\nu\in M_{\mathbb{Q}} and UνU_{\nu} be chosen as in Theorem 2, then

ρ(Uνd)=μν(Uν)=sν.\rho(U_{\nu}\cap\mathbb{Z}^{d})=\mu_{\nu}(U_{\nu})=s_{\nu}.
Proof.

We set

Uν={Uνν=ν,νν,U^{\prime}_{\nu^{\prime}}=\begin{cases}U_{\nu}&\nu^{\prime}=\nu,\\ \emptyset&\nu^{\prime}\neq\nu,\end{cases}

and let

P:d\displaystyle P^{\prime}:\mathbb{Z}^{d} \displaystyle\rightarrow 2M,\displaystyle 2^{M_{\mathbb{Q}}},
a\displaystyle a \displaystyle\mapsto {νMaUν}.\displaystyle\left\{\nu\in M_{\mathbb{Q}}\mid a\in U^{\prime}_{\nu}\right\}.

Then by Theorem 2 we have ρ(Uνd)=ρ(P1({ν}))=sν\rho(U_{\nu}\cap\mathbb{Z}^{d})=\rho(P^{\prime-1}(\{\nu\}))=s_{\nu}. ∎

We observe, that the elements AdA\in\mathbb{Z}^{d}, which are in UνU_{\nu} for infinitely many νM\nu\in M_{\mathbb{Q}} have density zero.

Lemma 6.

Let (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} be as in Theorem 2. Then we have that

ρ({AdAUνfor infinitely manyνM})=0.\rho(\{A\in\mathbb{Z}^{d}\mid A\in U_{\nu}\ \text{for infinitely many}\ \nu\in M_{\mathbb{Q}}\})=0.
Proof.

Since UνU_{\nu} were chosen as in Theorem 2, condition (2.1) holds, i.e.,

limMρ¯({AdAUpfor some primep>M})=0.\lim\limits_{M\rightarrow\infty}\bar{\rho}\left(\left\{A\in\mathbb{Z}^{d}\mid A\in U_{p}\ \text{for some prime}\ p>M\right\}\right)=0.

Let us call

CM\displaystyle C_{M} ={AdAUpfor some primep>M},\displaystyle=\left\{A\in\mathbb{Z}^{d}\mid A\in U_{p}\ \text{for some prime}\ p>M\right\},
I\displaystyle I ={AdAUνfor infinitely manyνM}.\displaystyle=\left\{A\in\mathbb{Z}^{d}\mid A\in U_{\nu}\ \text{for infinitely many}\ \nu\in M_{\mathbb{Q}}\right\}.

Clearly ICMI\subset C_{M} for all MM\in\mathbb{N}, hence

ρ¯(I)limMρ¯(CM)=0\overline{\rho}(I)\leq\lim\limits_{M\to\infty}\overline{\rho}(C_{M})=0

and thus ρ(I)=0\rho(I)=0. ∎

Over d\mathbb{Z}^{d} one can give a definition of mean or expected value (see for example [11]), as we will now explain. Observe, that an event with “probability” zero should not have any influence on the expected value, this legitimates that over d\mathbb{Z}^{d} we will exclude the elements AdA\in\mathbb{Z}^{d}, which are in infinitely many UνU_{\nu}, namely AIA\in I. Let us define [H,H[Id=([H,H[dd)I[-H,H[^{d}_{I}=([-H,H[^{d}\cap\mathbb{Z}^{d})\setminus I.

Definition 7.

Let HH and dd be positive integers and assume that (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} satisfy the assumptions of Theorem 2, then we define the expected value of the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} to be

μ=limHA[H,H[Id{νMAUν}(2H)d,\mu=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}}\mid\{\nu\in M_{\mathbb{Q}}\mid A\in U_{\nu}\}\mid}{(2H)^{d}}},

if it exists.

This limit essentially gives the expected value of the number of places ν\nu, such that a “random” element in d\mathbb{Z}^{d} is in UνU_{\nu}.

Remark 8.

The reader should notice that the mean of the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} should be thought as the “expected value” of the function a|P(a)|a\mapsto|P(a)|, where PP is the map of Theorem 2.

.

Definition 9.

For a set TT, for which we can compute its density via the local to global principle as in Theorem 2, we say that a system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} corresponds to TT, if TC=P1({})T^{C}=P^{-1}(\{\emptyset\}).

Observe that we can restrict Definition 7 to subsets of [H,H[Id[-H,H[^{d}_{I}, i.e., we define the expected value of the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} restricted to T[H,H[IdT\subset[-H,H[^{d}_{I} to be

μT=limHA[H,H[IdT{νMAUν}[H,H[IdT,\mu_{T}=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}\cap T}\mid\{\nu\in M_{\mathbb{Q}}\mid A\in U_{\nu}\}\mid}{\mid[-H,H[^{d}_{I}\cap T\mid}},

if it exists. Note, that this is analogous to the conditional expected value.

Remark 10.

If (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} corresponds to TT, then the expected value of the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} restricted to TT should be thought as the “expected value” of the function a|P(a)|a\mapsto|P(a)|, where PP is the map of Theorem 2, when restricted to the elements aa of d\mathbb{Z}^{d} such that |P(a)|1|P(a)|\geq 1.

One can easily pass from μ\mu to μT\mu_{T} and viceversa:

Lemma 11.

If the density of TT exists and is nonzero and TT is such that TCP1({})T^{C}\subseteq P^{-1}(\{\emptyset\}), then μ\mu exists iff μT\mu_{T} exists. Furthermore, in that case it holds that μ=μTρ(T)\mu=\mu_{T}\rho(T).

Proof.

We observe that

μT\displaystyle\mu_{T} =limHA[H,H[IdT{νMAUν}[H,H[IdT\displaystyle=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}\cap T}\mid\{\nu\in M_{\mathbb{Q}}\mid A\in U_{\nu}\}\mid}{\mid[-H,H[^{d}_{I}\cap T\mid}}
=limHA[H,H[IdT{νMAUν}(2H)d(2H)d[H,H[IdT\displaystyle=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}\cap T}\mid\{\nu\in M_{\mathbb{Q}}\mid A\in U_{\nu}\}\mid}{(2H)^{d}}}\displaystyle{\frac{(2H)^{d}}{\mid[-H,H[^{d}_{I}\cap T\mid}}
=limHA[H,H[IdT{νMAUν}(2H)d1ρ(T).\displaystyle=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}\cap T}\mid\{\nu\in M_{\mathbb{Q}}\mid A\in U_{\nu}\}\mid}{(2H)^{d}}}\displaystyle{\frac{1}{\rho(T)}}.

Let us define

τ(A,ν)={1AUν,0else.\tau(A,\nu)=\begin{cases}1&A\in U_{\nu},\\ 0&\text{else.}\end{cases}

Note that one can write [H,H[IdT[-H,H[^{d}_{I}\cap T as [H,H[Id([H,H[IdTC)[-H,H[^{d}_{I}\setminus([-H,H[^{d}_{I}\cap T^{C}), doing so we observe that we can ignore TT:

μTρ(T)\displaystyle\mu_{T}\rho(T) =limHA[H,H[IdTνMτ(A,ν)(2H)d\displaystyle=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}\cap T}\sum\limits_{\nu\in M_{\mathbb{Q}}}\tau(A,\nu)}{(2H)^{d}}}
=limHA[H,H[IdνMτ(A,ν)A[H,H[IdTcνMτ(A,ν)(2H)d.\displaystyle=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}}\sum\limits_{\nu\in M_{\mathbb{Q}}}\tau(A,\nu)-\sum\limits_{A\in[-H,H[^{d}_{I}\cap T^{c}}\sum\limits_{\nu\in M_{\mathbb{Q}}}\tau(A,\nu)}{(2H)^{d}}}.

Since TCP1({})T^{C}\subseteq P^{-1}(\{\emptyset\}), it holds that τ(A,ν)=0\tau(A,\nu)=0 for all ATCA\in T^{C} and hence we are left with μ\mu. ∎

In the applications of this paper one usually chooses TC=P1({})T^{C}=P^{-1}(\{\emptyset\}), and computes the expected value restricted to TT. This is a natural choice, since by the definition of TCT^{C} it holds that none of its elements lie in any of the UνU_{\nu}, thus we are only considering the subset TT, where nonzero values are added to the expected value.

One would now expect that, if the pp-adic measures of the UpU_{p}’s of Theorem 2 were essentially behaving like probabilities, one would have that the mean of the system (Up)p𝒫(U_{p})_{p\in\mathcal{P}} (as as defined in Definition 7) would be equal to p𝒫sp\sum_{p\in\mathcal{P}}s_{p}, since the density of UpdU_{p}\cap\mathbb{Z}^{d} is equal to sps_{p} (for example, this always happens when one has Up=U_{p}=\emptyset for all but finitely many UpU_{p}’s). Note that this would also be the result if we could simply move the limit inside the series. This is not the case: in fact, Condition (2.1) of Theorem 2, is not enough to ensure the existence of the mean as in the natural Definition 7. The next example shows a case where Condition (2.1) is verified, but the mean does not exist.

Example 12.

We set U=U_{\infty}=\emptyset and for all jj\in\mathbb{N} with 2nj<2n+12^{n}\leq j<2^{n+1} we define Upj={2n}U_{p_{j}}=\{2^{n}\}, where pjp_{j} denotes the jjth prime number. As p\mathbb{Z}_{p} with the pp-adic metric is a metric space, we get that UpjU_{p_{j}} is a closed set and thus Borel-measurable, of measure zero. Furthermore, every ball in the pp-adic metric contains infinitely many elements, which implies (Upj)=U¯pjUpj=Upj=Upj\partial(U_{p_{j}})=\overline{U}_{p_{j}}\setminus U_{p_{j}}^{\circ}=U_{p_{j}}\setminus\emptyset=U_{p_{j}}. As the pp-adic Haar measure is invariant under translation and normalized, we get that all finite sets are null sets. In particular, we get μpj((Upj))=0\mu_{p_{j}}(\partial(U_{p_{j}}))=0. Furthermore, we have

ρ¯(p𝒫:p>MUp)ρ¯(νMUν)=ρ¯({2nn})=0.\displaystyle\overline{\rho}\left(\bigcup_{p\in\mathcal{P}:p>M}U_{p}\right)\leq\overline{\rho}\left(\bigcup_{\nu\in M_{\mathbb{Q}}}U_{\nu}\right)=\overline{\rho}(\{2^{n}\ \mid\ n\in\mathbb{N}\})=0.

Hence, Condition (2.1) is satisfied, even without taking the limit in MM. Let nn\in\mathbb{N}, then for 2nH<2n+12^{n}\leq H<2^{n+1} we have

{νM[H,H[Uν}={p1,p2,,p2n+11}.\displaystyle\{\nu\in M_{\mathbb{Q}}\ \mid\ [-H,H[\cap U_{\nu}\neq\emptyset\}=\{p_{1},p_{2},\dots,p_{2^{n+1}-1}\}.

Thus, we have for all nn\in\mathbb{N}

A[2n+1,2n+1[I|{νM:AUν}|22n+1=2n+1122n+1n12\displaystyle\sum_{A\in[-2^{n+1},2^{n+1}[_{I}}\frac{|\{\nu\in M_{\mathbb{Q}}\ :\ A\in U_{\nu}\}|}{2\cdot 2^{n+1}}=\frac{2^{n+1}-1}{2\cdot 2^{n+1}}\stackrel{{\scriptstyle n\rightarrow\infty}}{{\longrightarrow}}\frac{1}{2}

and

A[(2n+1+1),2n+1+1[I|{νM:AUν}|2(2n+1+1)=2n+212(2n+1+1)n1.\displaystyle\sum_{A\in[-(2^{n+1}+1),2^{n+1}+1[_{I}}\frac{|\{\nu\in M_{\mathbb{Q}}\ :\ A\in U_{\nu}\}|}{2(2^{n+1}+1)}=\frac{2^{n+2}-1}{2(2^{n+1}+1)}\stackrel{{\scriptstyle n\rightarrow\infty}}{{\longrightarrow}}1.

Hence, the expected value of the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} does not exist, even though it satisfies all conditions of Theorem 2.

Now we state the main theorem, which is a local to global principle for expected values, extending the results in [20, Lemma 20].

Theorem 13.

Let HH and dd be positive integers. Let UdU_{\infty}\subset\mathbb{R}^{d}, such that 0U=U\mathbb{R}_{\geq 0}\cdot U_{\infty}=U_{\infty} and μ((U))=0.\mu_{\infty}(\partial(U_{\infty}))=0. Let s=12dμ(U[1,1]d)s_{\infty}=\frac{1}{2^{d}}\mu_{\infty}(U_{\infty}\cap[-1,1]^{d}). For each prime pp, let UppdU_{p}\subset\mathbb{Z}_{p}^{d}, such that μp((Up))=0\mu_{p}(\partial(U_{p}))=0 and define sp=μp(Up)s_{p}=\mu_{p}(U_{p}). Define the following map

P:d\displaystyle P:\mathbb{Z}^{d} \displaystyle\rightarrow 2M,\displaystyle 2^{M_{\mathbb{Q}}},
a\displaystyle a \displaystyle\mapsto {νMaUν}.\displaystyle\left\{\nu\in M_{\mathbb{Q}}\mid a\in U_{\nu}\right\}.

If (2.1) is satisfied and for some α[0,)\alpha\in[0,\infty) there exists an absolute constant cc\in\mathbb{Z}, such that for all H1H\geq 1 and for all A[H,H[IdA\in[-H,H[^{d}_{I} one has that

(3.1) |{p𝒫p>Hα,AUp[H,H[Id}|<c\left|\left\{p\in\mathcal{P}\mid p>H^{\alpha},A\in U_{p}\cap[-H,H[_{I}^{d}\right\}\right|<c

and and that there exists a sequence (vp)p𝒫(v_{p})_{p\in\mathcal{P}}, such that for all p<Hαp<H^{\alpha} one has that

(3.2) Up[H,H[Id\displaystyle\mid U_{p}\cap[-H,H[^{d}_{I}\mid vp(2H)d,\displaystyle\leq v_{p}(2H)^{d},
(3.3) p𝒫vp\displaystyle\sum_{p\in\mathcal{P}}v_{p} converges,\displaystyle\text{converges},

then it follows that the mean of the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}}, exists and is given by:

μ\displaystyle\mu =νMsν.\displaystyle=\sum\limits_{\nu\in M_{\mathbb{Q}}}s_{\nu}.
Remark 14.

In a nuthshell, the additional conditions give control on the number of UpU_{p} for which an element in [H,H[Id[-H,H[_{I}^{d} can live in: Condition 3.1 gives control in the large pp regime (and small HH), and Condition 3.2 (with 3.3) gives control in the small pp regime (and large HH). None of the conditions can be removed, as we will show later with counterexamples in each case.

Proof.

For H,M>0H,M>0 we split

A[H,H[Id{νMAUν}(2H)d=s1(H,M)+s2(H,M)+s3(H,M),\displaystyle\sum\limits_{A\in[-H,H[^{d}_{I}}\frac{\mid\{\nu\in M_{\mathbb{Q}}\mid A\in U_{\nu}\}\mid}{(2H)^{d}}=s_{1}(H,M)+s_{2}(H,M)+s_{3}(H,M),

where

s1(H,M)\displaystyle s_{1}(H,M) =A[H,H[Id{p𝒫MHα<p,AUp}(2H)d,\displaystyle=\sum\limits_{A\in[-H,H[^{d}_{I}}\frac{\mid\{p\in\mathcal{P}\mid M\leq H^{\alpha}<p,A\in U_{p}\}\mid}{(2H)^{d}},
s2(H,M)\displaystyle s_{2}(H,M) =A[H,H[Id{p𝒫M<p<Hα,AUp}(2H)d,\displaystyle=\sum\limits_{A\in[-H,H[^{d}_{I}}\frac{\mid\{p\in\mathcal{P}\mid M<p<H^{\alpha},A\in U_{p}\}\mid}{(2H)^{d}},
s3(H,M)\displaystyle s_{3}(H,M) =A[H,H[Id{ν𝒫ν= or νM,AUν}(2H)d.\displaystyle=\sum\limits_{A\in[-H,H[^{d}_{I}}\frac{\mid\{\nu\in\mathcal{P}\mid\nu=\infty\text{ or }\nu\leq M,A\in U_{\nu}\}\mid}{(2H)^{d}}.

We are going to show that for j{1,2}j\in\{1,2\} we have

lim supMlim supH|sj(H,M)|=0andlimMlimHs3(H,M)=ννMsν,\limsup_{M\to\infty}\limsup_{H\to\infty}|s_{j}(H,M)|=0\quad\text{and}\quad\lim_{M\rightarrow\infty}\lim_{H\rightarrow\infty}s_{3}(H,M)=\sum_{\nu_{\nu\in M_{\mathbb{Q}}}}s_{\nu},

which readily implies that

μ=limHA[H,H[Id{νMAUν}(2H)d\displaystyle\mu=\lim\limits_{H\to\infty}\displaystyle{\frac{\sum\limits_{A\in[-H,H[^{d}_{I}}\mid\{\nu\in M_{\mathbb{Q}}\mid A\in U_{\nu}\}\mid}{(2H)^{d}}}

exists and that

μ=νMsν.\displaystyle\mu=\sum_{\nu\in M_{\mathbb{Q}}}s_{\nu}.

First we consider the case α0\alpha\neq 0. Let us define for H>0H>0 and AdA\in\mathbb{Z}^{d}

(3.4) {p𝒫p>Hα,AUp[H,H[Id}=A,H.\mid\left\{p\in\mathcal{P}\mid p>H^{\alpha},A\in U_{p}\cap[-H,H[_{I}^{d}\right\}\mid=\ell_{A,H}.

Notice that thanks to condition (3.1) there exists a constant c>0c>0 independent of AA and HH such that

A,H<c.\displaystyle\ell_{A,H}<c.

Therefore, we get

0\displaystyle 0 lim supMlim supH|s1(H,M)|\displaystyle\leq\limsup_{M\to\infty}\limsup_{H\to\infty}|s_{1}(H,M)|
lim supMlim supHA[H,H[IdM<p𝒫UpA,H(2H)d\displaystyle\leq\limsup_{M\to\infty}\limsup_{H\to\infty}\sum\limits_{A\in[-H,H[^{d}_{I}\cap\bigcup_{M<p\in\mathcal{P}}U_{p}}\frac{\ell_{A,H}}{(2H)^{d}}
lim supMlim supHA[H,H[IdM<p𝒫Upc(2H)d\displaystyle\leq\limsup_{M\to\infty}\limsup_{H\to\infty}\sum\limits_{A\in[-H,H[^{d}_{I}\cap\bigcup_{M<p\in\mathcal{P}}U_{p}}\frac{c}{(2H)^{d}}
=clim supMlim supH|[H,H[IdM<p𝒫Up|(2H)d\displaystyle=c\limsup_{M\to\infty}\limsup_{H\to\infty}\ \displaystyle{\frac{|[-H,H[_{I}^{d}\cap\bigcup_{M<p\in\mathcal{P}}U_{p}|}{(2H)^{d}}}
=clim supMρ¯(M<p𝒫Up)=0,\displaystyle=c\limsup_{M\to\infty}\ \overline{\rho}\left(\bigcup\limits_{M<p\in\mathcal{P}}U_{p}\right)=0,

where the last equality follows from Condition (2.1). Using (3.2) and (3.3) we get

0lim supMlim supH|s2(H,M)|\displaystyle 0\leq\limsup_{M\to\infty}\limsup_{H\to\infty}|s_{2}(H,M)| =lim supMlim supHp𝒫,M<p<Hα|Up[H,H[Id|(2H)d\displaystyle=\limsup_{M\to\infty}\limsup_{H\to\infty}\sum_{p\in\mathcal{P},\ M<p<H^{\alpha}}\frac{|U_{p}\cap[-H,H[_{I}^{d}|}{(2H)^{d}}
lim supMlim supHp𝒫,M<p<Hαvp=0.\displaystyle\leq\limsup_{M\to\infty}\limsup_{H\to\infty}\sum_{p\in\mathcal{P},\ M<p<H^{\alpha}}v_{p}=0.

For α=0\alpha=0 on the other hand, we have for M>1M>1 that s1(H,M)=0=s2(H,M)s_{1}(H,M)=0=s_{2}(H,M). Using Corollary 5 we get

limMlimHs3(H,M)\displaystyle\lim_{M\rightarrow\infty}\lim_{H\rightarrow\infty}s_{3}(H,M) =limMlimHνM,νM or ν=[H,H[IdUν(2H)d\displaystyle=\lim_{M\to\infty}\lim_{H\to\infty}\sum\limits_{\nu\in M_{\mathbb{Q}},\ \nu\leq M\text{ or }\nu=\infty}\displaystyle{\frac{\mid[-H,H[^{d}_{I}\cap U_{\nu}\mid}{(2H)^{d}}}
=limMνM,νM orν=ρ(Uνd)\displaystyle=\lim_{M\to\infty}\sum\limits_{\nu\in M_{\mathbb{Q}},\ \nu\leq M\ \text{ or}\ \nu=\infty}\rho(U_{\nu}\cap\mathbb{Z}^{d})
=limMνM,νMorν=sν\displaystyle=\lim_{M\to\infty}\sum\limits_{\nu\in M_{\mathbb{Q}},\ \nu\leq M\ \text{or}\ \nu=\infty}s_{\nu}
=νMsν.\displaystyle=\sum_{\nu\in M_{\mathbb{Q}}}s_{\nu}.

A natural question is whether some of the conditions in Theorem 13 are redundant. This is not the case as the next two examples show.

Example 15.

In this example, we construct (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} such that Conditions (3.1) and (2.1) are verified but the conclusion of Theorem 13 does not hold. As in Example 12, we denote by pjp_{j} the jjth prime. We choose

U=,Upj={pj,pj+1,,p2j}.U_{\infty}=\emptyset,\quad U_{p_{j}}=\{p_{j},p_{j+1},\dots,p_{2^{j}}\}.

The same argument as in Example 12 shows that μpj((Upj))=0\mu_{p_{j}}(\partial(U_{p_{j}}))=0. By the prime number theorem we have that 𝒫\mathcal{P} has density zero and thus our choice satisfies Condition (2.1) even without taking the limit in MM.

Note that for pj>H>0p_{j}>H>0 we have Upj[H,H[=U_{p_{j}}\cap[-H,H[=\emptyset and thus Condition (3.1) is satisfied with c=1=mc=1=m. Now we will show that the conclusion of the theorem does not hold. First we show that μ=\mu=\infty, if the limit exists. Note that for this example [H,H[I=[H,H[[-H,H[_{I}=[-H,H[ for all H>0H>0. Let LL\in\mathbb{N}, then we compute

A[p2L,p2L[|{νM|AUν}|2p2L=j=12L|Upj[p2L,p2L[|2p2L\displaystyle\sum_{A\in[-p_{2^{L}},p_{2^{L}}[}\frac{|\{\nu\in M_{\mathbb{Q}}\ |\ A\in U_{\nu}\}|}{2p_{2^{L}}}=\sum_{j=1}^{2^{L}}\frac{|U_{p_{j}}\cap[-p_{2^{L}},p_{2^{L}}[|}{2p_{2^{L}}}
=j=1L1|{pj,,p2j}[p2L,p2L[|2p2L+j=L2L1|{pj,,p2L1}|2p2L\displaystyle=\sum_{j=1}^{L-1}\frac{|\{p_{j},\dots,p_{2^{j}}\}\cap[-p_{2^{L}},p_{2^{L}}[|}{2p_{2^{L}}}+\sum_{j=L}^{2^{L}-1}\frac{|\{p_{j},\dots,p_{2^{L}-1}\}|}{2p_{2^{L}}}
=j=1L12jj+12p2L+j=L2L12Lj2p2L\displaystyle=\sum_{j=1}^{L-1}\frac{2^{j}-j+1}{2p_{2^{L}}}+\sum_{j=L}^{2^{L}-1}\frac{2^{L}-j}{2p_{2^{L}}}
=12p2L[(2L2)L(L1)2+(L1)+2L(2LL)(2L1)2L2+L(L1)2].\displaystyle=\frac{1}{2p_{2^{L}}}\left[\left(2^{L}-2\right)-\frac{L(L-1)}{2}+(L-1)+2^{L}(2^{L}-L)-\frac{(2^{L}-1)2^{L}}{2}+\frac{L(L-1)}{2}\right].

Hence, for LL sufficiently large we get

(3.5) A[p2L,p2L[|{νM|AUν}|2p2L15p2L22L.\sum_{A\in[-p_{2^{L}},p_{2^{L}}[}\frac{|\{\nu\in M_{\mathbb{Q}}\ |\ A\in U_{\nu}\}|}{2p_{2^{L}}}\geq\frac{1}{5p_{2^{L}}}2^{2L}.

Thus, using the prime number theorem, we obtain

lim supHA[H,H[I|{νM|AUν}|(2H)\displaystyle\limsup_{H\rightarrow\infty}\sum_{A\in[-H,H[_{I}}\frac{|\{\nu\in M_{\mathbb{Q}}\ |\ A\in U_{\nu}\}|}{(2H)} lim supLA[p2L,p2L[|{νM|AUν}|2p2L\displaystyle\geq\limsup_{L\rightarrow\infty}\sum_{A\in[-p_{2^{L}},p_{2^{L}}[}\frac{|\{\nu\in M_{\mathbb{Q}}\ |\ A\in U_{\nu}\}|}{2p_{2^{L}}}
lim supL15p2L22L\displaystyle\geq\limsup_{L\rightarrow\infty}\frac{1}{5p_{2^{L}}}2^{2L}
=lim supL2L5ln(2)L=.\displaystyle=\limsup_{L\rightarrow\infty}\frac{2^{L}}{5\ln(2)L}=\infty.

As noted above, all finite sets are null sets for the pp-adic Haar measure. Thus, we have sp=0s_{p}=0 for all p𝒫p\in\mathcal{P} and s=12μ()=0s_{\infty}=\frac{1}{2}\mu_{\infty}(\emptyset)=0 and hence

νMsν=0.\sum_{\nu\in M_{\mathbb{Q}}}s_{\nu}=0.

This means the conclusion of the theorem does not hold, if we only assume Conditions (2.1) and (3.1).

Example 16.

Next we construct an example that satisfies Conditions (2.1), (3.2), (3.3), and we show that the conclusion of the theorem does not hold. We set U=U_{\infty}=\emptyset and for p𝒫{p2nn}p\in\mathcal{P}\setminus\{p_{2^{n}}\ \mid\ n\in\mathbb{N}\} we define Up=U_{p}=\emptyset. Inductively, we define the remaining Up2nU_{p_{2^{n}}}. We start with Up1={1}U_{p_{1}}=\{1\}. If Up2n={m2}U_{p_{2^{n}}}=\{m^{2}\} for mm\in\mathbb{N}, then we define

Up2n+1={{m2},if |{jjn,Up2j={m2}}|<m3,{(m+1)2},else.\displaystyle U_{p_{2^{n+1}}}=\begin{cases}\{m^{2}\},&\text{if }|\{j\in\mathbb{N}\ \mid\ j\leq n,U_{p_{2^{j}}}=\{m^{2}\}\}|<m^{3},\\ \{(m+1)^{2}\},&\text{else}.\end{cases}

The same argument as for Example 12 applies here and gives μpj((Upj))=0\mu_{p_{j}}(\partial(U_{p_{j}}))=0. We compute

ρ¯(νMUν)=ρ¯({n2:n})\displaystyle\overline{\rho}\left(\bigcup_{\nu\in M_{\mathbb{Q}}}U_{\nu}\right)=\overline{\rho}\left(\{n^{2}\ :\ n\in\mathbb{N}\}\right) =lim supH|{n:n2<H}|2H\displaystyle=\limsup_{H\rightarrow\infty}\frac{|\{n\in\mathbb{N}\ :\ n^{2}<H\}|}{2H}
lim supHH2H=0.\displaystyle\leq\limsup_{H\rightarrow\infty}\frac{\sqrt{H}}{2H}=0.

Hence, Condition (2.1) is satisfied. We have for H>pnH>p_{n}

|[H,H[Upn|={1if n=2k for some k,0otherwise.\displaystyle|[-H,H[\cap U_{p_{n}}|=\begin{cases}1&\text{if }n=2^{k}\text{ for some }k\in\mathbb{N},\\ 0&\text{otherwise}.\end{cases}

Thus, we may pick

vpn={1pnif n=2k for some k,0otherwise.\displaystyle v_{p_{n}}=\begin{cases}\frac{1}{p_{n}}&\text{if }n=2^{k}\ \text{ for some }k\in\mathbb{N},\\ 0&\text{otherwise}.\end{cases}

By the prime number theorem, there exists some constant D>0D>0, such that

pn𝒫vpn=k11p2k=k12kln(2k)p2k12kln(2k)k1D2kln(2k)<.\displaystyle\sum_{p_{n}\in\mathcal{P}}v_{p_{n}}=\sum_{k\geq 1}\frac{1}{p_{2^{k}}}=\sum_{k\geq 1}\frac{2^{k}\ln(2^{k})}{p_{2^{k}}}\frac{1}{2^{k}\ln(2^{k})}\leq\sum_{k\geq 1}\frac{D}{2^{k}\ln(2^{k})}<\infty.

Thus, conditions (3.2) and (3.3) are satisfied.
By construction, we have

|{νM|AUν}|={m3A=m2 for some m>0,0else.\displaystyle|\{\nu\in M_{\mathbb{Q}}\ |\ A\in U_{\nu}\}|=\begin{cases}m^{3}&A=m^{2}\text{ for some }m\in\mathbb{N}_{>0},\\ 0&\text{else}.\end{cases}

Therefore, we get

A[H,H[I|{νM|AUν}|(2H)\displaystyle\sum_{A\in[-H,H[_{I}}\frac{|\{\nu\in M_{\mathbb{Q}}\ |\ A\in U_{\nu}\}|}{(2H)} m=1H1m32H=12HH24(H1)2\displaystyle\geq\sum_{m=1}^{\lfloor\sqrt{H}\rfloor-1}\frac{m^{3}}{2H}=\frac{1}{2H}\frac{\lfloor\sqrt{H}\rfloor^{2}}{4}(\lfloor\sqrt{H}\rfloor-1)^{2}
132(H1)2.\displaystyle\geq\frac{1}{32}(\lfloor\sqrt{H}\rfloor-1)^{2}.

Hence,

lim supHA[H,H[I|{νM|AUν}|(2H)=.\displaystyle\limsup_{H\rightarrow\infty}\sum_{A\in[-H,H[_{I}}\frac{|\{\nu\in M_{\mathbb{Q}}\ |\ A\in U_{\nu}\}|}{(2H)}=\infty.

On the other hand, by the same argument as in the previous example we obtain

νMsν=0.\displaystyle\sum_{\nu\in M_{\mathbb{Q}}}s_{\nu}=0.

4. Applications

We can apply Theorem 13 to sets, whose densities were computed via the local to global principle of Theorem 2 and fulfill Conditions (3.1), (3.2) and (3.3).

For example we can compute the expected number of common prime divisors of all basic minors of a rectangular non-unimodular matrix. The rigorous statement reads as follows.

Corollary 17.

Let n<mn<m be positive integers, and let us denote by RR the set of rectangular unimodular matrices in n×m\mathbb{Z}^{n\times m}. Then the corresponding system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} is given as in [18], i.e., U=U_{\infty}=\emptyset and for p𝒫p\in\mathcal{P} denote by UpU_{p} the set of all matrices in pn×m\mathbb{Z}_{p}^{n\times m} whose nn-minors are all divisible by pp.

Then the expected value of the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} exists and is given by

(4.1) μ\displaystyle\mu =νMsν=p𝒫(1i=0n1(11pmi)).\displaystyle=\sum_{\nu\in M_{\mathbb{Q}}}s_{\nu}=\sum\limits_{p\in\mathcal{P}}\left(1-\prod\limits_{i=0}^{n-1}\left(1-\frac{1}{p^{m-i}}\right)\right).

And the average number of primes that divide all nn-minors of a rectangular non-unimodular matrix is given by

(4.2) μRC=μ1i=0n11ζ(mi),\mu_{R^{C}}=\displaystyle{\frac{\mu}{1-\prod\limits_{i=0}^{n-1}\frac{1}{\zeta(m-i)}}},

where ζ\zeta denotes the Riemann zeta function.

Proof.

Recall from [18], that all conditions of Theorem 2 are satisfied for the corresponding system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}}, that for p𝒫p\in\mathcal{P} we have that

sp=(1i=0n1(11pmi))s_{p}=\left(1-\prod\limits_{i=0}^{n-1}\left(1-\frac{1}{p^{m-i}}\right)\right)

and thus

ρ(R)=ρ(P1({}))=i=0n11ζ(mi).\rho(R)=\rho(P^{-1}(\{\emptyset\}))=\prod\limits_{i=0}^{n-1}\frac{1}{\zeta(m-i)}.

Thanks to Lemma 11 we are left with proving that the additional assumptions on the system (Uν)νM(U_{\nu})_{\nu\in M_{\mathbb{Q}}} of Theorem 13 are satisfied.

For this, let us choose α=1\alpha=1 and denote by fif_{i} the function associating to An×mA\in\mathbb{Z}^{n\times m} some fixed nn-minor. Then we have for all A[H,H[n×mA\in[-H,H[^{n\times m} the inequality fi(A)(2H)nf_{i}(A)\leq(2H)^{n} for all ii. We exclude that fi(A)f_{i}(A) vanishes for all ii, since then we land in II.

Recall, that

A,H=|{p𝒫p>H,AUp[H,H[Inm}|.\ell_{A,H}=\left|\left\{p\in\mathcal{P}\mid p>H,\ A\in U_{p}\cap[-H,H[^{nm}_{I}\right\}\right|.

Thus, we have that

HA,Hp𝒫p>H,AUp[H,H[Inmp.H^{\ell_{A,H}}\leq\prod\limits_{\begin{subarray}{c}p\in\mathcal{P}\\ p>H,\ A\in U_{p}\cap[-H,H[^{nm}_{I}\end{subarray}}p.

Further, observe that

p𝒫p>H,AUp[H,H[Inmp=gcd((fi(A))i)(2H)n.\prod\limits_{\begin{subarray}{c}p\in\mathcal{P}\\ p>H,\ A\in U_{p}\cap[-H,H[^{nm}_{I}\end{subarray}}p=\gcd((f_{i}(A))_{i})\leq(2H)^{n}.

Hence, we get that HA,H(2H)nH^{\ell_{A,H}}\leq(2H)^{n}, and thus Condition (3.1) is satisfied.

To verify Condition (3.2) we want to show that there exists a sequence (vp)p𝒫(v_{p})_{p\in\mathcal{P}} such that for all p<Hp<H we have that Up[H,H[Inmvp(2H)nm\mid U_{p}\cap[-H,H[^{nm}_{I}\mid\leq v_{p}(2H)^{nm}. The set of non-full rank matrices over 𝔽p\mathbb{F}_{p} has size

pnmi=0n1(pmpi)2npm(n1)+n1=2np(m+1)(n1).p^{nm}-\prod\limits_{i=0}^{n-1}(p^{m}-p^{i})\leq 2^{n}p^{m(n-1)+n-1}=2^{n}p^{(m+1)(n-1)}.

We can fix one non-full rank n×mn\times m matrix over 𝔽p\mathbb{F}_{p}, for which we have less than or equal to 2np(m+1)(n1)2^{n}p^{(m+1)(n-1)} choices. For this fixed matrix there are less than or equal to (2Hp)nm(\lceil\frac{2H}{p}\rceil)^{nm} lifts to n×m[H,H[nm\mathbb{Z}^{n\times m}\cap[-H,H[^{nm}. Hence, we have for p<Hp<H

Up[H,H[Inm\displaystyle\mid U_{p}\cap[-H,H[^{nm}_{I}\mid 2n(2Hp)nmp(m+1)(n1)\displaystyle\leq 2^{n}\left(\left\lceil\frac{2H}{p}\right\rceil\right)^{nm}p^{(m+1)(n-1)}
2n(2Hp+1)nmp(m+1)(n1)\displaystyle\leq 2^{n}\left(\frac{2H}{p}+1\right)^{nm}p^{(m+1)(n-1)}
2n(3H)nmpnmm+n1nm\displaystyle\leq 2^{n}(3H)^{nm}p^{nm-m+n-1-nm}
6nmHnm1p2.\displaystyle\leq 6^{nm}H^{nm}\frac{1}{p^{2}}.

Thus (vp)p𝒫(v_{p})_{p\in\mathcal{P}} can be chosen to be (6nmp2)p𝒫(\frac{6^{nm}}{p^{2}})_{p\in\mathcal{P}}, which also satisfies Condition (3.3). Hence, (4.1) follows and Lemma 11 implies (4.2). ∎

By choosing n=1n=1 we get the mean of numbers of primes dividing non-coprime mm-tuples of integers and of course choosing n=1n=1 and m=2m=2 will give the mean of numbers of primes dividing non-coprime pairs of integers.

The results of [11] regarding the average amount of primes for which a non-monic Eisenstein-polynomial satisifies the criterion of Eisenstein follow immediately as corollary using Theorem 13.

Corollary 18.

Let d2d\geq 2 be an integer. The expected number of primes pp for which an Eisenstein polynomial of degree dd is pp-Eisenstein, is given by

(1p𝒫(1(p1)2pd+2))1p𝒫(p1)2pd+2.\left(1-\prod_{p\in\mathcal{P}}\left(1-\frac{(p-1)^{2}}{p^{d+2}}\right)\right)^{-1}\sum_{p\in\mathcal{P}}\frac{(p-1)^{2}}{p^{d+2}}.
Proof.

In Theorem 13 simply use the system Up=(ppp2p)×ppd1×(ppp)U_{p}=(p\mathbb{Z}_{p}\setminus p^{2}\mathbb{Z}_{p})\times p\mathbb{Z}_{p}^{d-1}\times(\mathbb{Z}_{p}\setminus p\mathbb{Z}_{p}) and choose α=1\alpha=1. Let E(H)E(H) be the set of Eisenstein polynomials of height at most HH. Condition (3.1) is trivially verified, as no polynomial can be Eisenstein with respect to a prime larger than its height. Condition (3.3) is verified thanks to the rough estimate |E(H)|2H/pdH|E(H)|\leq\lceil 2H/p\rceil^{d}\cdot H, and this is enough for our purposes because p𝒫1/pd\sum_{p\in\mathcal{P}}1/p^{d} converges for d2d\geq 2.

Moreover, using the system E¯p\overline{E}_{p}’s given in [15] we can obtain the expected number of primes for which a given polynomial f(x)f(x) is Eisenstein for some shift f(x+i)f(x+i). For the sake of completeness and to show how easy it is to apply Theorem 13 when the system of UpU_{p}’s is given, let us now compute this expected value. Let d3d\geq 3 be a positive integer and let E¯p\overline{E}_{p} be the set of polynomials ff of degree dd in p[x]\mathbb{Z}_{p}[x] such that there exists ii for which f(x+i)f(x+i) is pp-Eisenstein in p[x]\mathbb{Z}_{p}[x]. Cleary, the expected value of the system (Up)p𝒫=(E¯p)p𝒫(U_{p})_{p\in\mathcal{P}}=(\overline{E}_{p})_{p\in\mathcal{P}} (where every E¯p\overline{E}_{p} is seen as a subset of pd+1\mathbb{Z}_{p}^{d+1}) is exactly the number we are interested in. Let us verify that the system satisfies the three conditions

  • Condition (2.1) has already been verified in [15], as it is needed to compute the density of the set of shifted Eisenstein polynomials.

  • Condition (3.1) is immediate with the choice α=1\alpha=1. In fact, the primes pp, for which a polynomial is pp-Eisenstein, divide the discriminant of the polynomial. Notice that f(x+i)f(x+i) has the same discriminant as f(x)f(x), which is bounded by CH2d2CH^{2d-2} (as the discriminant is homogeneous of degree 2d22d-2), for some absolute constant CC. Thus, one obtains that the product of the primes pp larger than HH for which f(x+i)f(x+i) is pp-Eisenstein for some ii is bounded by an absolute constant.

  • Condition (3.3) is also easy to verify. First, observe that if f(x+i)f(x+i) is pp-Eisenstein for some ii\in\mathbb{N}, then ii can be chosen less than pp (see for example [15, Lemma 6]). So that the set of shifted pp-Eisenstein polynomials SS in [H,H[d+1[-H,H[^{d+1} is absolutely bounded by |E(H)|p|E(H)|\cdot p, where E(H)E(H) is the set of pp-Eisenstein polynomials of degree d+1d+1 and height at most HH. Finally, one can very roughly give the estimate |E(H)|2H/pdH|E(H)|\leq\lceil 2H/p\rceil^{d}\cdot H, which is anyway enough for our purposes as p𝒫1/pd1\sum_{p\in\mathcal{P}}1/p^{d-1} converges for d3d\geq 3.

Using the computation of the pp-adic measure of E¯p\overline{E}_{p} from [15] and Lemma 11 (to obtain the restricted mean), one obtains the following result:

Corollary 19.

For d3d\geq 3, the mean of the system (E¯p)p𝒫(\overline{E}_{p})_{p\in\mathcal{P}} (i.e., the expected number of primes for which a polynomial f(x)f(x) is Eisenstein after some shift) is

p𝒫(p1)2pd+1.\sum\limits_{{p\in\mathcal{P}}}\frac{(p-1)^{2}}{p^{d+1}}.

The expected value restricted to the set of shifted Eisenstein polynomials is

(1p𝒫(1(p1)2pd+1))1p𝒫(p1)2pd+1.\left(1-\prod\limits_{{p\in\mathcal{P}}}\left(1-\frac{(p-1)^{2}}{p^{d+1}}\right)\right)^{-1}\sum\limits_{{p\in\mathcal{P}}}\frac{(p-1)^{2}}{p^{d+1}}.

Acknowledgments

The second author is thankful to the Swiss National Science Foundation grant number 20020_172623. The third author is partially supported by Swiss National Science Foundation grant number 188430.

References

  • [1] Martin Bright, Tim Daniel Browning, and Daniel Loughran. Failures of weak approximation in families. Compositio Mathematica, 152(7):1435–1475, 2016.
  • [2] Ernesto Cesaro. Question 75 (solution). Mathesis, 3:224–225, 1883.
  • [3] Ernesto Cesaro. Probabilité de certains faits arithméthiques. Mathesis, 4:150–151, 1884.
  • [4] Edoardo Dotti and Giacomo Micheli. Eisenstein polynomials over function fields. Applicable Algebra in Engineering, Communication and Computing, 27(2):159–168, 2016.
  • [5] Andrea Ferraguti. The set of stable primes for polynomial sequences with large Galois group. Proceedings of the American Mathematical Society, 2018.
  • [6] Andrea Ferraguti and Giacomo Micheli. On the Mertens–Césaro theorem for number fields. Bulletin of the Australian Mathematical Society, 93(02):199–210, 2016.
  • [7] Xiangqian Guo and Guangyu Yang. The probability of rectangular unimodular matrices over 𝔽q[x]\mathbb{F}_{q}[x]. Linear algebra and its applications, 438(6):2675–2682, 2013.
  • [8] Randell Heyman and Igor E. Shparlinski. On the number of Eisenstein polynomials of bounded height. Applicable Algebra in Engineering, Communication and Computing, 24(2):149–156, 2013.
  • [9] Randell Heyman and Igor E. Shparlinski. On shifted Eisenstein polynomials. Periodica Mathematica Hungarica, 69(2):170–181, 2014.
  • [10] Julia Lieb. Uniform probability and natural density of mutually left coprime polynomial matrices over finite fields. Linear Algebra and its Applications, 539:134–159, 2018.
  • [11] Shilin Ma, Kevin McGown, Devon Rhodes, and Mathias Wanner. On the number of primes for which a polynomial is Eisenstein. INTEGERS, 18:2, 2018.
  • [12] Gérard Maze. Natural density distribution of Hermite normal forms of integer matrices. Journal of Number Theory, 131(12):2398–2408, 2011.
  • [13] Franz Mertens. Ueber einige Asymptotische Gesetze der Zahlentheorie. Journal für die reine und angewandte Mathematik, 77:289–338, 1874.
  • [14] Giacomo Micheli. A local to global principle for densities over function fields. arXiv preprint arXiv:1701.01178, 2017.
  • [15] Giacomo Micheli and Reto Schnyder. The density of shifted and affine Eisenstein polynomials. Proceedings of the American Mathematical Society, 144(11):4651–4661, 2016.
  • [16] Giacomo Micheli and Reto Schnyder. The density of unimodular matrices over integrally closed subrings of function fields. In Contemporary Developments in Finite Fields and Applications, pages 244–253. World Scientific, 2016.
  • [17] Giacomo Micheli and Reto Schnyder. On the density of coprime mm-tuples over holomorphy rings. International Journal of Number Theory, 12(03):833–839, 2016.
  • [18] Giacomo Micheli and Violetta Weger. On rectangular unimodular matrices over the algebraic integers. SIAM Journal on Discrete Mathematics, 33(1):425–437, 2019.
  • [19] James E. Nymann. On the probability that kk positive integers are relatively prime. Journal of Number Theory, 4(5):469–473, 1972.
  • [20] Bjorn Poonen and Michael Stoll. The Cassels-Tate pairing on polarized Abelian varieties. Annals of Mathematics, 150(3):1109–1149, 1999.
  • [21] Bjorn Poonen and Michael Stoll. A local-global principle for densities. In Scott D. Ahlgren, George E. Andrews, and K. Ono, editors, Topics in Number Theory, volume 467 of Mathematics and Its Applications, pages 241–244. Springer US, 1999.
  • [22] Brian D. Sittinger. The probability that random algebraic integers are relatively rr-prime. Journal of Number Theory, 130(1):164–171, 2010.
  • [23] Ekedahl Torsten. An infinite version of the Chinese remainder theorem. Commentarii mathematici Universitatis Sancti Pauli= Rikkyo Daigaku sugaku zasshi, 40(1):53–59, 1991.
  • [24] Yinghui Wang and Richard P Stanley. The Smith normal form distribution of a random integer matrix. SIAM Journal on Discrete Mathematics, 31(3):2247–2268, 2017.