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

Generalizations of Alladi’s formula for arithmetical semigroups

Lian Duan Department of Mathematics, Colorado State University, Fort Collins, CO 80523, USA [email protected] Ning Ma Department of Mathematics, State University of New York at Buffalo, Buffalo, NY 14260, USA [email protected]  and  Shaoyun Yi Department of Mathematics, University of South Carolina, Columbia, SC 29208, USA [email protected]
Abstract.

In this article, we prove that a general version of Alladi’s formula with Dirichlet convolution holds for arithmetical semigroups satisfying Axiom AA or Axiom A#A^{\#}. As applications, we apply our main results to certain semigroups coming from algebraic number theory, arithmetical geometry and graph theory.

Key words and phrases:
Arithmetical semigroups, Alladi’s formula, Equidistribution, Möbius function, Arithmetical geometry, Number theory, Graph theory
2020 Mathematics Subject Classification:
Primary 11N80, 11R45, 20M13; Secondary 11G25, 11M41, 11R44, 11R58.

1. Introduction and main results

Let μ:1{0,±1}\mu:\mathbb{Z}_{\geq 1}\to\{0,\pm 1\} be the Möbius function defined by μ(n)=(1)k\mu(n)=(-1)^{k} if nn is the product of kk distinct primes, and zero otherwise. Let pmin(n)p_{\min}(n) be the smallest prime factor of nn with pmin(1)=1p_{\min}(1)=1. In 1977, Alladi [1] discovered a relationship between the Möbius function μ(n)\mu(n) and the density of primes in arithmetic progressions by showing

(1) n2pmin(n)(modk)μ(n)n=1φ(k)-\sum_{\begin{smallmatrix}n\geq 2\\ p_{\min}(n)\equiv\ell\,(\operatorname{mod}k)\end{smallmatrix}}\frac{\mu(n)}{n}=\frac{1}{\varphi(k)}

for any positive integer kk and integer \ell with (,k)=1(\ell,k)=1, where φ\varphi is Euler’s totient function. Recently, Wang [32] proved that if an arithmetic function a:1a:\mathbb{Z}_{\geq 1}\to\mathbb{C} satisfies a(1)=1a(1)=1 and n=2|a(n)|nloglogn<\sum_{n=2}^{\infty}\frac{|a(n)|}{n}\log\log n<\infty, then

(2) n2pmin(n)Sμa(n)n=δ(S),-\sum_{\begin{smallmatrix}n\geq 2\\ p_{\min}(n)\in S\end{smallmatrix}}\frac{\mu*a(n)}{n}=\delta(S),

where μa\mu*a is the Dirichlet convolution of μ\mu and aa and δ(S)\delta(S) is the natural density for a set SS of primes. In particular, if one takes aa to be such that a(1)=1a(1)=1 and a(n)=0a(n)=0 for all n2n\in\mathbb{Z}_{\geq 2}, then (2) recovers the result of (1) by the prime number theorem in arithmetic progressions in which we have δ(S)=1/φ(k)\delta(S)=1/\varphi(k). Notice that 1\mathbb{Z}_{\geq 1} is a multiplicative semigroup generated by rational prime integers. Indeed, the semigroup structure satisfies the Axiom AA (see Sect. 1.1). This motivates us to consider the following question:

Question 1.1.

To what extend one can generalize the Alladi’s formula and still get an analogue of (2)?

In this work, we discover proper formulations of Wang’s result over all arithmetical semigroups satisfying Axiom AA or Axiom A#A^{\#}. More precisely, due to the fact that the “smallest prime factor” of an element in a general arithmetical semigroup 𝒢\mathcal{G} is not well-defined, we focus on those arithmetic functions aa only supported on the distinguished set 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) defined in (5) and (11). However, introducing this distinguished set causes a new difficulty which cannot be handled by Wang’s method in [32]. To deal with this difficulty, we use another method to avoid analyzing distinguished sets. By this trick, we manage to show a variation of Alladi’s formula with convolution still hold for every such arithmetical semigroup; see Theorem 1.3 and Theorem 1.6 for details. Moreover, we apply our main results specifically to the arithmetical semigroups of interests, i.e., from algebraic number theory, algebraic geometry and graph theory; see Sect. 4 for explicit details. We hope our results and methods will find applications in future researches.

Recall [15, Chap. 1, § 1] that an arithmetical semigroup is a pair (𝒢,)(\mathcal{G},\|\cdot\|) (or simply 𝒢\mathcal{G} if there is no confusion about \|\cdot\|). Here 𝒢\mathcal{G} is a commutative semigroup with the identity element e𝒢e_{\mathcal{G}}, which is freely generated by a finite or countable subset 𝒫\mathcal{P} of primes of 𝒢\mathcal{G}. Explicitly,

𝒢\colonequals{finite productPiai:Pi𝒫 and ai0 for all i}.\mathcal{G}\colonequals\left\{\prod_{\text{finite product}}P_{i}^{a_{i}}\colon P_{i}\in\mathcal{P}\text{ and }a_{i}\geq 0\text{ for all }i\right\}.

In addition, there exists a discrete real-valued norm mapping \|\cdot\| on 𝒢\mathcal{G} such that

  1. (1)(1)

    e𝒢=1\|e_{\mathcal{G}}\|=1, and P>1\|P\|>1 for every P𝒫P\in\mathcal{P}.

  2. (2)(2)

    gh=gh\|gh\|=\|g\|\cdot\|h\| for all g,h𝒢g,h\in\mathcal{G}.

  3. (3)(3)

    For each real number x>0x>0, the total number 𝒩(x)\mathcal{N}(x) of elements g𝒢g\in\mathcal{G} of norm gx\left\|g\right\|\leq x is finite.

Note that the conditions (1)-(3) are equivalent to conditions (1) and (2) together with

  1. (3)(3)^{\prime}

    For each real number x>0x>0, the total number π(x)\pi(x) of elements P𝒫P\in\mathcal{P} of norm Px\left\|P\right\|\leq x is finite.

Sometimes, instead of considering the norm mapping as above, it will be more convenient to consider a degree mapping \partial on 𝒢\mathcal{G} (see [17, Sect. 1.1]), which is defined as (g)\colonequalslogcg\partial(g)\colonequals\log_{c}\left\|g\right\| for some fixed constant c>1c>1. So the conditions (1)(3)(1)-(3) above read as

  1. (1#1^{\#})

    (e𝒢)=0\partial(e_{\mathcal{G}})=0 and (P)>0\partial(P)>0 for every P𝒫P\in\mathcal{P}.

  2. (2#2^{\#})

    (gh)=(g)+(h)\partial(gh)=\partial(g)+\partial(h) for all g,h𝒢g,h\in\mathcal{G}.

  3. (3#3^{\#})

    For each real number x>0x>0, the total number 𝒩#(x)\mathcal{N}^{\#}(x) of elements g𝒢g\in\mathcal{G} with (g)x\partial(g)\leq x is finite.

Among arithmetical semigroups, we are particularly interested in two types for which some additional axiom is satisfied. In the following subsections, each type together with the corresponding main result will be introduced.

1.1. Axiom AA type arithmetical semigroups

Definition 1.2.

[15, Chap. 4, § 1] An arithmetical semigroup 𝒢\mathcal{G} satisfies Axiom AA if

(3) 𝒩(x)=c𝒢x+O(xη)\mathcal{N}(x)=c_{\mathcal{G}}x+O(x^{\eta})

with suitable constants c𝒢>0c_{\mathcal{G}}>0 and 0<η<10<\eta<1111In [15, Chap. 4, § 1] an arithmetical semigroup (𝒢,)(\mathcal{G},\|\cdot\|) satisfies Axiom AA if 𝒩(x)=c𝒢xδ+O(xηδ),\mathcal{N}(x)=c_{\mathcal{G}}x^{\delta}+O(x^{\eta\delta}), with suitable constants c𝒢>0c_{\mathcal{G}}>0 and 0<η<10<\eta<1. In our work, for technical convenience we always normalize 𝒢\mathcal{G} by taking the new norm g\colonequals|g|δ\left\|g\right\|\colonequals|g|^{\delta} such that (3) holds..

For any subset S𝒫S\subseteq\mathcal{P}, we say that SS has a natural density δ(S)\delta(S) , if the following limit exists:

(4) δ(S)\colonequalslimxπ𝒢,S(x)π𝒢(x),\delta(S)\colonequals\lim_{x\to\infty}\frac{\pi_{\mathcal{G},S}(x)}{\pi_{\mathcal{G}}(x)},

where π𝒢,S(x)\colonequals#{PS:Px}\pi_{\mathcal{G},S}(x)\colonequals\#\left\{P\in S:\|P\|\leq x\right\} and π𝒢(x)\colonequalsπ𝒢,𝒫(x)\pi_{\mathcal{G}}(x)\colonequals\pi_{\mathcal{G},\mathcal{P}}(x). For g,h𝒢g,h\in\mathcal{G}, we say h|gh|g, if there exist r𝒢r\in\mathcal{G} such that hr=ghr=g. Let N(g)\colonequalsmin{P:P|g}\mathrm{N}_{-}(g)\colonequals\min\{\|P\|\colon P|g\} be the minimum norm of all prime factors of g𝒢g\in\mathcal{G}. And we say that gg is distinguishable if ge𝒢g\neq e_{\mathcal{G}} and there is a unique prime factor, say Pmin(g)P_{\min}(g), of gg attaining the minimum norm N(g)\mathrm{N}_{-}(g). Furthermore, we define222Note that our definition is a little different from [5, 27, 18] in the sense that they do not include the identity.

(5) 𝔇(𝒢,S)\colonequals{g𝒢:g is distinguishable and Pmin(g)S}{e𝒢}.\mathfrak{D}(\mathcal{G},S)\colonequals\left\{g\in\mathcal{G}:g\text{ is distinguishable and }P_{\min}(g)\in S\right\}\cup\{e_{\mathcal{G}}\}.

We will also need the following natural generalization of the Möbius function to elements g𝒢:g\in\mathcal{G}\colon

μ𝒢(g)={1 if g=e𝒢,(1)k if k>0 and g=P1Pk,0 if P2|g for some prime P.\mu_{\mathcal{G}}(g)=\begin{cases}1&\text{ if }g=e_{\mathcal{G}},\\ (-1)^{k}&\text{ if }k>0\text{ and }g=P_{1}\cdots P_{k},\\ 0&\text{ if }P^{2}|g\text{ for some prime }P.\end{cases}

Throughout, we will write μ(g)\colonequalsμ𝒢(g)\mu(g)\colonequals\mu_{\mathcal{G}}(g) for simplicity when the context is clear. An arithmetic function a:𝒢a:\mathcal{G}\to\mathbb{C} is said to be supported on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) if a(g)=0a(g)=0 unless g𝔇(𝒢,S)g\in\mathfrak{D}(\mathcal{G},S). And the Dirichlet convolution of μ\mu and aa is defined by μa(g)=h|gμ(h)a(g/h)\mu*a(g)=\sum_{h|g}\mu(h)a(g/h). Moreover, we say that aa is the identity of the convolution ring of arithmetic functions over elements of 𝒢\mathcal{G} if a(g)=0a(g)=0 for all ge𝒢g\neq e_{\mathcal{G}} and a(e𝒢)=1a(e_{\mathcal{G}})=1. With all above notations, we can state our first main result of this work.

Theorem 1.3.

Given an Axiom AA type arithmetical semigroup 𝒢\mathcal{G}, assume that S𝒫S\subseteq\mathcal{P} has a natural density δ(S)\delta(S). Suppose that a:𝒢a\colon\mathcal{G}\to\mathbb{C} is an arithmetic function supported on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) with a(e𝒢)=1a(e_{\mathcal{G}})=1, and

limx2gxg𝔇(𝒢,S)|a(g)|gloglogg<.\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|g\right\|\leq x\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{|a(g)|}{\left\|g\right\|}\log\log\left\|g\right\|<\infty.

Let μa\mu*a be the Dirichlet convolution of μ\mu and aa. Then

(6) limx2gxg𝔇(𝒢,S)μa(g)g=δ(S).-\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|g\right\|\leq x\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)}{\left\|g\right\|}=\delta(S).

In particular, if aa is the identity of the convolution ring of arithmetic functions over elements of 𝒢\mathcal{G}, then

(7) limx2gxg𝔇(𝒢,S)μ(g)g=δ(S).-\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|g\right\|\leq x\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu(g)}{\left\|g\right\|}=\delta(S).
Corollary 1.4.

With 𝒢,S\mathcal{G},S and 𝒫\mathcal{P} as in Theorem 1.3, suppose that a:𝒢a\colon\mathcal{G}\to\mathbb{C} is an arithmetic function supported on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) with a(e𝒢)=1a(e_{\mathcal{G}})=1 and |a(g)|gα|a(g)|\ll\|g\|^{-\alpha} for some α>0\alpha>0, then

(8) limx2gxg𝔇(𝒢,S)μa(g)φ(g)=δ(S),-\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|g\right\|\leq x\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)}{\varphi(g)}=\delta(S),

where φ(g)\varphi(g) is Euler’s totient function defined by φ(g)=gP|g(11P).\varphi(g)=\|g\|\prod_{P|g}\left(1-\frac{1}{\|P\|}\right).

1.2. Axiom A#A^{\#} type arithmetical semigroups

In this case (unless otherwise stated), it shall be assumed that the degree map \partial has image in 0\mathbb{Z}_{\geq 0}. Furthermore, we call 𝒢\mathcal{G} together with \partial an additive arithmetical semigroup. In particular, we write

𝒢\colonequals{finite sumaiPi:Pi𝒫 and ai0 for all i}.\mathcal{G}\colonequals\left\{\sum_{\text{finite sum}}a_{i}P_{i}\colon P_{i}\in\mathcal{P}\text{ and }a_{i}\geq 0\text{ for all }i\right\}.

In the following let G#(n)G^{\#}(n) be the total number of elements of degree nn in 𝒢\mathcal{G}.

Definition 1.5.

[17, Sect. 1.1] An arithmetical semigroup 𝒢\mathcal{G} satisfies Axiom A# if

(9) G#(n)=c𝒢qn+O(qηn)G^{\#}(n)=c_{\mathcal{G}}q^{n}+O(q^{\eta n})

with suitable constants c𝒢>0c_{\mathcal{G}}>0, q>1q>1 and 0η<10\leq\eta<1.

In addition, we assume that Z𝒢(q1)0Z_{\mathcal{G}}(-q^{-1})\neq 0.333This is a technical condition to simplify our arguments. In fact (see e.g. [17]) many consequences of Axiom A#A^{\#} are unrelated to the value of Z𝒢(q1)Z_{\mathcal{G}}(-q^{-1}), and so the simplifying additional condition would only sometimes become relevant. For the semigroups which do not satisfy this condition, one can refer to [14]. Here the function Z𝒢(z)\colonequalsP(1z(P))1Z_{\mathcal{G}}(z)\colonequals\prod_{P}(1-z^{\partial(P)})^{-1} is the so-called zeta function of 𝒢\mathcal{G}.

Similarly but differently, for 𝒢\mathcal{G} satisfying Axiom A#A^{\#} as above, we define the natural density δ(S)\delta(S) for a subset S𝒫S\subseteq\mathcal{P} by the following limit if it exists:

(10) δ(S)\colonequalslimnπ𝒢,S#(n)π𝒢#(n),\delta(S)\colonequals\lim_{n\to\infty}\frac{\pi^{\#}_{\mathcal{G},S}(n)}{\pi^{\#}_{\mathcal{G}}(n)},

where π𝒢,S#(n)\colonequals#{PS:(P)=n}\pi^{\#}_{\mathcal{G},S}(n)\colonequals\#\left\{P\in S:\partial(P)=n\right\} and π𝒢#(n)\colonequalsπ𝒢,𝒫#(n)\pi^{\#}_{\mathcal{G}}(n)\colonequals\pi^{\#}_{\mathcal{G},\mathcal{P}}(n). For g,h𝒢g,h\in\mathcal{G}, we say ghg\geq h, or equivalently, h|gh|g, if there exists r𝒢r\in\mathcal{G} such that h+r=gh+r=g. Let d(g)\colonequalsmin{(P):P|g}d_{-}(g)\colonequals\min\left\{\partial(P):P|g\right\} be the minimum degree of all prime factors of g𝒢g\in\mathcal{G}. We say that gg is distinguishable if ge𝒢g\neq e_{\mathcal{G}} and there is a unique prime factor, say Pmin(g)P_{\min}(g), of gg attaining the minimum degree d(g)d_{-}(g). Furthermore, we define

(11) 𝔇#(𝒢,S)\colonequals{g𝒢:g is distinguishable and Pmin(g)S}{e𝒢}.\mathfrak{D}^{\#}(\mathcal{G},S)\colonequals\left\{g\in\mathcal{G}:g\text{ is distinguishable and }P_{\min}(g)\in S\right\}\cup\{e_{\mathcal{G}}\}.

We will write 𝔇(𝒢,S)\colonequals𝔇#(𝒢,S)\mathfrak{D}(\mathcal{G},S)\colonequals\mathfrak{D}^{\#}(\mathcal{G},S) for simplicity when the context is clear. Moreover, the general Möbius function μ𝒢\mu_{\mathcal{G}}, arithmetic functions supported on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) and the Dirichlet convolution of two arithmetic functions are defined in the same way as those in the previous subsection (perhaps except that one needs to translate the group law from multiplication to addition). In the following theorem, we fix the norm map :𝒢\|\cdot\|:\mathcal{G}\to\mathbb{C} to be g=q(g)\|g\|=q^{\partial(g)} with the base qq as in Definition 1.5.

Theorem 1.6.

Given an Axiom A#A^{\#} type arithmetical semigroup 𝒢\mathcal{G}, assume that S𝒫S\subseteq\mathcal{P} has a natural density δ(S)\delta(S). Suppose that a:𝒢a\colon\mathcal{G}\to\mathbb{C} is an arithmetic function supported on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) with a(e𝒢)=1a(e_{\mathcal{G}})=1, and

limn1(g)ng𝔇(𝒢,S)|a(g)|gloglogg<.\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{|a(g)|}{\|g\|}\log\log\|g\|<\infty.

Let μa\mu*a be the Dirichlet convolution of μ\mu and aa. Then

(12) limn1(g)ng𝔇(𝒢,S)μa(g)g=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)}{\|g\|}=\delta(S).

In particular, if aa is the identity of the convolution ring of arithmetic functions over elements of 𝒢\mathcal{G}, then

(13) limn1(g)ng𝔇(𝒢,S)μ(g)g=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu(g)}{\|g\|}=\delta(S).
Corollary 1.7.

With 𝒢,S\mathcal{G},S and 𝒫\mathcal{P} as in Theorem 1.6, suppose that a:𝒢a\colon\mathcal{G}\to\mathbb{C} is an arithmetic function supported on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) with a(e𝒢)=1a(e_{\mathcal{G}})=1 and |a(g)|gα|a(g)|\ll\|g\|^{-\alpha} for some α>0\alpha>0, then

(14) limn1(g)ng𝔇(𝒢,S)μa(g)φ(g)=δ(S),-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)}{\varphi(g)}=\delta(S),

where φ(g)\varphi(g) is Euler’s totient function defined by φ(g)=gP|g(11P).\varphi(g)=\|g\|\prod_{P|g}\left(1-\frac{1}{\|P\|}\right).

1.3. Prior work

Alladi’s original result (1) was generalized by Dawsey [5] to the setting of Chebotarev densities for finite Galois extensions of \mathbb{Q}, and further by Sweeting and Woo [27] to number fields. Later, Kural-McDonald-Sah [18] generalized all these results to natural densities of sets of prime ideals of a number field KK. A recent work [9] of Wang with the first and third authors of this article showed the analogue of Kural-McDonald-Sah’s result over global function fields. Wang [30, 31, 32] showed the analogues of these results over \mathbb{Q} for some arithmetic functions other than μ\mu. It is worth to notice that in the partition setting (subject to Axiom CC [15, Chap. 8, § 1]), a series work of Schneider [24, 25], Ono-Wagner-Schneider [23, 22] and Dawsey-Just-Schneider [6] have built the framework of the corresponding generalizations. The interested readers are recommended to their papers.

1.4. A quick summary of some applications

There are many examples of arithmetical semigroups satisfying Axiom AA or Axiom A#A^{\#}; see [15, 17, 29]. As for a reference of other types, see [15, Chap. 8, § 1]. In this subsection, we list the applications of Theorem 1.3 and Theorem 1.6 to some examples of interest in number theory, arithmetical geometry and graph theory. To save the space, only the essential information will be included in this subsection, and more detailed background of each example will be provided in Sect. 4. In particular, we would like to fix the following notations.

  1. (1)

    For a number field KK, we will denote the semigroup of integral ideals by K\mathcal{I}_{K}.

  2. (2)

    For a dd-dimensional smooth projective variety XX over a finite field 𝔽q\mathbb{F}_{q}, where q=prq=p^{r} is a power of prime integer pp, we will denote the semigroup of effective 0-cycles of XX by 𝒜X\mathcal{A}_{X}.

  3. (3)

    For a finite connected undirected graph GG without degree-11 vertices, we will denote the semigroup generated by classes of primitive paths by 𝒜G\mathcal{A}_{G}.

In Table 1, we summarize the necessary information for each of above examples. Specifically, for each of the above examples, we list the type, the prime elements and the general elements of the associated semigroup. We also specify the degree map (if it makes sense) and the norm map those are involved in the corresponding result. As a support that each semigroup in this table satisfies either Axiom AA or Axiom A#A^{\#}, we list its corresponding function 𝒩\mathcal{N} or G#G^{\#}, which is equivalent to the associated prime number theorem (PNT).

Semigroup K\mathcal{I}_{K} 𝒜X\mathcal{A}_{X} 𝒜G\mathcal{A}_{G}
Type Axiom AA Axiom A#A^{\#} Axiom A#A^{\#}
Prime elements prime ideals 𝔭\mathfrak{p} prime 0-cycles PP over 𝔽q\mathbb{F}_{q} equivalent classes [P][P] of primitive paths
General elements integral ideals 𝔞\mathfrak{a} of 𝒪K\mathcal{O}_{K} effective 0-cycles AA of XX over 𝔽q\mathbb{F}_{q} finite formal sum [C]=ai[Pi],ai0[C]=\sum a_{i}[P_{i}],a_{i}\geq 0
Degree map \partial (P)=#{geometric points above P}\partial(P)=\#\{\text{geometric points above }P\} ([P])=length of P\partial([P])=\text{length of }P
Norm map \|\cdot\| 𝔭=#(𝒪K/𝔭)\|\mathfrak{p}\|=\#(\mathcal{O}_{K}/\mathfrak{p}) P=(qd)(P)\|P\|=(q^{d})^{\partial(P)} [P]=1/RG([P])\|[P]\|=1/R_{G}^{\partial([P])}
𝒩\mathcal{N} or G#G^{\#} 𝒩(x)=cKx+O(x11/[K:])\mathcal{N}(x)=c_{K}x+O(x^{1-1/[K:\mathbb{Q}]}) G#(n)=cX(qd)ΔGn+O(q(d1/2)n)G^{\#}(n)=c_{X}(q^{d})^{\Delta_{G}n}+O(q^{(d-1/2)n}) G#(n)=cG/RGΔGn+O(1/RGΔGηn)G^{\#}(n)=c_{G}/R_{G}^{\Delta_{G}n}+O(1/R_{G}^{\Delta_{G}\eta n})
PNT πK(x)=xlogx+O(xexp(cKlogx))\pi_{K}(x)=\dfrac{x}{\log x}+O(x\,\mathrm{exp}(-c_{K}\sqrt{\log x})) πX#(n)=(qd)nn+O(q(d1/2)nn)\pi^{\#}_{X}(n)=\dfrac{(q^{d})^{n}}{n}+O\left(\dfrac{q^{(d-1/2)n}}{n}\right) πG#(n)=1nRGΔGn+O(1nRGΔGηn)\pi^{\#}_{G}(n)=\dfrac{1}{nR^{\Delta_{G}n}_{G}}+O\left(\dfrac{1}{nR_{G}^{\Delta_{G}\eta n}}\right)
Main result Corollary 1.8 Corollary 1.10 Corollary 1.14
Reference section Sect. 4.1 Sect. 4.2 Sect. 4.3
Table 1. Examples of semigroups

It follows from Table 1 and Theorem 1.3 that we have the following result for the semigroup K\mathcal{I}_{K} of integral ideals of a number field KK.

Corollary 1.8.

For a number field KK, assume that S𝒫S\subseteq\mathcal{P} is a subset of prime ideals with natural density δ(S)\delta(S). For any arithmetic function a:Ka:\mathcal{I}_{K}\to\mathbb{C} supported on 𝔇(K,S)\mathfrak{D}(\mathcal{I}_{K},S) with a(𝒪K)=1a(\mathcal{O}_{K})=1, and

limx2𝔞x𝔞𝔇(K,S)|a(𝔞)|𝔞loglog𝔞<.\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|\mathfrak{a}\right\|\leq x\\ \mathfrak{a}\in\mathfrak{D}(\mathcal{I}_{K},S)\end{smallmatrix}}\frac{|a(\mathfrak{a})|}{\left\|\mathfrak{a}\right\|}\log\log\left\|\mathfrak{a}\right\|<\infty.

Then

(15) limx2𝔞x𝔞𝔇(K,S)μa(𝔞)𝔞=δ(S).-\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|\mathfrak{a}\right\|\leq x\\ \mathfrak{a}\in\mathfrak{D}(\mathcal{I}_{K},S)\end{smallmatrix}}\frac{\mu*a(\mathfrak{a})}{\left\|\mathfrak{a}\right\|}=\delta(S).

In particular, we have

(16) limx2𝔞x𝔞𝔇(K,S)μ(𝔞)𝔞=δ(S).-\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|\mathfrak{a}\right\|\leq x\\ \mathfrak{a}\in\mathfrak{D}(\mathcal{I}_{K},S)\end{smallmatrix}}\frac{\mu(\mathfrak{a})}{\left\|\mathfrak{a}\right\|}=\delta(S).
Remark 1.9.

Note that (16) recovers the main theorem in [18]. In this sense, Corollary 1.8 is a generalization of the result in [18] to some arithmetic functions other than μ\mu. On the other hand, if we let K=K=\mathbb{Q}, then (15) in this case is exactly the main theorem of [32]. And so in this sense, Corollary 1.8 generalizes the result in [32] to all number fields.

It follows from Table 1 and Theorem 1.6 that we have the following result for the semigroup 𝒜X\mathcal{A}_{X} of effective 0-cycles of a dd-dimensional smooth projective variety XX over a finite field 𝔽q\mathbb{F}_{q}.

Corollary 1.10.

For a smooth projective dd-dimensional variety XX defined over a finite field 𝔽q\mathbb{F}_{q}, assume that S𝒫S\subseteq\mathcal{P} is a subset of prime 0-cycles with natural density δ(S)\delta(S). For any arithmetic function a:𝒜Xa:\mathcal{A}_{X}\to\mathbb{C} supported on 𝔇(𝒜X,S)\mathfrak{D}(\mathcal{A}_{X},S) with a(0)=1a(0)=1, and

limn1(A)nA𝔇(𝒜X,S)|a(A)|AloglogA<.\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(A)\leq n\\ A\in\mathfrak{D}(\mathcal{A}_{X},S)\end{smallmatrix}}\frac{|a(A)|}{\|A\|}\log\log\|A\|<\infty.

Then

(17) limn1(A)nA𝔇(𝒜X,S)μa(A)A=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(A)\leq n\\ A\in\mathfrak{D}(\mathcal{A}_{X},S)\end{smallmatrix}}\frac{\mu*a(A)}{\|A\|}=\delta(S).

In particular, we have

(18) limn1(A)nA𝔇(𝒜X,S)μ(A)A=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(A)\leq n\\ A\in\mathfrak{D}(\mathcal{A}_{X},S)\end{smallmatrix}}\frac{\mu(A)}{\|A\|}=\delta(S).
Remark 1.11.

When XX has the dimension one, we have d=1d=1 in the above setups. Then one can check that in this case the semigroup 𝒜X\mathcal{A}_{X} coincides with the semigroup of effective divisors of XX. It follows that (18) recovers the main theorem of [9]. Hence one can think Corollary 1.10 as a generalization of the result in [9] to some arithmetic functions other than μ\mu for all higher dimensional varieties.

Remark 1.12.

We note that the “smooth and projective” condition in Corollary 1.10 is not essential. In fact, if XX is a geometrically connected scheme, then 0-cycles of XX are still well defined and so analogous results in Table 1 can be deduced from its Hasse-Weil zeta function by a similar argument as in Sect. 4.2; see [3] for more details about the zeta function of XX and the prime number theorem in general cases.

Remark 1.13.

Based on Corollary 1.10, we can also deduce some other consequences regarding to the distribution of hyperplanes of projective spaces; see Sect. 4.2 for more details.

Similarly, it follows from Table 1 and Theorem 1.6 that we have the following result for the semigroup 𝒜G\mathcal{A}_{G} generated by classes of primitive paths of a finite connected undirected graph GG without degree-11 vertices.

Corollary 1.14.

For a finite connected undirected graph GG without degree-11 vertices, assume that S𝒫S\subseteq\mathcal{P} is a subset of classes of primitive paths with natural density δ(S)\delta(S). For any arithmetic function a:𝒜Ga:\mathcal{A}_{G}\to\mathbb{C} supported on 𝔇(𝒜G,S)\mathfrak{D}(\mathcal{A}_{G},S) with a([0])=1a([0])=1, and

limn1([C])n[C]𝔇(𝒜G,S)|a([C])|[C]loglog[C]<.\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial([C])\leq n\\ [C]\in\mathfrak{D}(\mathcal{A}_{G},S)\end{smallmatrix}}\frac{|a([C])|}{\|[C]\|}\log\log\|[C]\|<\infty.

Then

(19) limn1([C])n[C]𝔇(𝒜G,S)μa([C])[C]=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial([C])\leq n\\ [C]\in\mathfrak{D}(\mathcal{A}_{G},S)\end{smallmatrix}}\frac{\mu*a([C])}{\|[C]\|}=\delta(S).

In particular, we have

(20) limn1([C])n[C]𝔇(𝒜G,S)μ([C])[C]=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial([C])\leq n\\ [C]\in\mathfrak{D}(\mathcal{A}_{G},S)\end{smallmatrix}}\frac{\mu([C])}{\|[C]\|}=\delta(S).

1.5. Organization of this paper

We will prove Theorem 1.6 for Axiom A#A^{\#} type arithmetical semigroup and Theorem 1.3 for Axiom AA type arithmetical semigroup in Sect. 2 and Sect. 3, respectively. Explicitly, to prove Theorem 1.6, we first give an analogous equidistribution result of the largest (in terms of degree ()\partial(\cdot)) prime factors of elements with the same degree in 𝒢\mathcal{G} (Lemma 2.1) and an analogue of Alladi’s duality identity to 𝒢\mathcal{G} (Lemma 2.2). Next, we prove the intermediate theorem (Theorem 2.3) and then Theorem 1.6. Due to the fact that the proofs of the results for Axiom AA type case use a similar strategy as that for Axiom A#A^{\#} type case, we will sketch its proofs and only emphasis on key steps in Sect. 3. Finally, in Sect. 4, we explain the contents in Table 1 in details, and also provide some more examples.

2. Proofs of main results for Axiom A#A^{\#} type arithmetical semigroups

In this section, we discuss the situation of Axiom A#A^{\#} type arithmetical semigroups. More precisely, there is a key result in this section, Theorem 2.3, and then Theorem 1.6 follows as an immediate consequence. Throughout this section, we fix 𝒢\mathcal{G} to be an Axiom A#A^{\#} type arithmetical semigroup, and adopt the definitions and notations as in Sect. 1.2.

For any element g𝒢g\in\mathcal{G}, we take d+(g)\colonequalsmax{(P):P|g}d^{+}(g)\colonequals\max\left\{\partial(P):P|g\right\} to be the largest degree of all prime factors of gg, and d+(e𝒢)=0d^{+}(e_{\mathcal{G}})=0. Let

(21) QS#(g)\colonequals#{PS:(P)=d+(g),P|g}Q_{S}^{\#}(g)\colonequals\#\left\{P\in S:\partial(P)=d^{+}(g),P|g\right\}

be the number of prime factors of gg in SS attaining the maximal degree d+(g)d^{+}(g). We will write QS(g)\colonequalsQS#(g)Q_{S}(g)\colonequals Q_{S}^{\#}(g) for simplicity when the context is clear. Then, with a similar process as in the proof of [9, Theorem 4.4] one can have the analogous asymptotic estimate for QS(g)Q_{S}(g) as follows.

Lemma 2.1.

If S𝒫S\subseteq\mathcal{P} has a natural density δ(S)\delta(S), then

(22) (g)=nQS(g)=c𝒢δ(S)qn+o(qn).\sum_{\partial(g)=n}Q_{S}(g)=c_{\mathcal{G}}\delta(S)q^{n}+o(q^{{n}}).

We will also require an analogue of Alladi’s duality property [1, Lemma 1] to 𝒢\mathcal{G} for our purpose.

Lemma 2.2.

Suppose that f:f:\mathbb{N}\to\mathbb{C} is an arithmetic function with f(0)=0f(0)=0. Then for any g𝒢g\in\mathcal{G} we have

(23) h|gμ(h)1𝔇(𝒢,S)(h)f(d(h))=QS(g)f(d+(g)).\sum_{h|g}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))=-Q_{S}(g)f(d^{+}(g)).

Here 1𝔇(𝒢,S)1_{\mathfrak{D}(\mathcal{G},S)} is the indicator function on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) defined as in (11).

Proof.

This directly generalizes [9, Lemma 3.1] and the desired result follows from a similar argument. ∎

Moreover, we will also need the following intermediate theorem.

Theorem 2.3.

Let ff be any bounded arithmetic function with f(0)=0f(0)=0. Let aa be an arithmetic function as in Theorem 1.6. The followings are equivalent:

(24) (g)=nQS(g)f(d+(g))c𝒢δ(S)qn.\sum_{\partial(g)=n}Q_{S}(g)f(d^{+}(g))\sim c_{\mathcal{G}}\delta(S)q^{n}.
(25) limn1(g)ng𝔇(𝒢,S)μ(g)f(d(g))g=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu(g)f(d_{-}(g))}{\|g\|}=\delta(S).
(26) limn1(g)ng𝔇(𝒢,S)μa(g)f(d(g))g=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)f(d_{-}(g))}{\|g\|}=\delta(S).

Before going into the details of proof of Theorem 2.3, we would like to give the proof of Theorem 1.6 based on this theorem.

Proof of Theorem 1.6.

The desired result (12) follows immediately from Lemma 2.1 and Theorem 2.3 by taking f(n)=1f(n)=1 for all n1n\in\mathbb{Z}_{\geq 1}. ∎

We will separate the proof of Theorem 2.3 into two parts: “(24)(25)\eqref{Eqn: equidis of largest prime factors}\Leftrightarrow\eqref{Eqn: analogue of main result in global function feild}” and “(25)(26)\eqref{Eqn: analogue of main result in global function feild}\Leftrightarrow\eqref{Eqn: to write into partial sum}”.

2.1. Proof of “(24)(25)\eqref{Eqn: equidis of largest prime factors}\Leftrightarrow\eqref{Eqn: analogue of main result in global function feild}

First, we prove the following three lemmas which estimate several partial sums involving the general Möbius function μ\mu.

Lemma 2.4.

Let 𝒢\mathcal{G} be an Axiom A#A^{\#} type arithmetical semigroup. Then for some 0η<10\leq\eta<1 we have

(27) (g)=nμ(g)=O(qηn).\sum_{\partial(g)=n}\mu(g)=O(q^{\eta n}).
Proof.

It is clear that for Re(s)>1\operatorname{Re}(s)>1 we have

1ζ𝒢(s)=g𝒢μ(g)gs=(g)0μ(g)q(g)s.\frac{1}{\zeta_{\mathcal{G}}(s)}=\sum_{g\in\mathcal{G}}\frac{\mu(g)}{\|g\|^{s}}=\sum_{\partial(g)\geq 0}\frac{\mu(g)}{q^{\partial(g)s}}.

By taking T=qsT=q^{-s}, we can rewrite this series as

Z𝒢(T)=n=0Cμ(n)Tn,Z_{\mathcal{G}}(T)=\sum_{n=0}^{\infty}C_{\mu}(n)T^{n},

where Cμ(n)=(g)=nμ(g)C_{\mu}(n)=\sum_{\partial(g)=n}\mu(g). Consider the following integral

(28) 12πi|T|=qη1Z𝒢(T)dTT(n+1)=Cμ(n),\frac{1}{2\pi i}\int_{|T|=q^{-\eta}}\frac{1}{Z_{\mathcal{G}}(T)}\frac{dT}{T^{(n+1)}}=C_{\mu}(n),

where we can choose 0η<10\leq\eta<1 such that Z𝒢(T)Z_{\mathcal{G}}(T) has no zeros in the closed disk |T|qη|T|\leq q^{-\eta}. Then using a similar argument as in the proof of [14, Theorem 1], we have that

(29) Cμ(n)=(g)=nμ(g)=O(qηn).C_{\mu}(n)=\sum_{\partial(g)=n}\mu(g)=O(q^{\eta n}).

Lemma 2.5.

For any m,n0m,n\in\mathbb{Z}_{\geq 0}, define

(30) C(n,m)\colonequals(g)=nd(g)>mμ(g).C(n,m)\colonequals\sum_{\begin{smallmatrix}\partial(g)=n\\ d_{-}(g)>m\end{smallmatrix}}\mu(g).

Then there exists some 0η<10\leq\eta<1 such that

(31) C(n,m)qηnexp(qm).C(n,m)\ll q^{\eta n}\mathrm{exp}(q^{m}).

In particular, we have

(32) M(n,m)\colonequals(g)nd(g)>mμ(g)qηnexp(qm).M(n,m)\colonequals\sum_{\begin{smallmatrix}\partial(g)\leq n\\ d_{-}(g)>m\end{smallmatrix}}\mu(g)\ll q^{\eta n}\mathrm{exp}(q^{m}).
Proof.

As in the proof of Lemma 2.4, if we consider the function

(33) 1ζ𝒢(s)(P)m(1P)s)1=g𝒢d(g)>mμ(g)gs,\frac{1}{\zeta_{\mathcal{G}}(s)}\prod_{\partial(P)\leq m}(1-\|P\|)^{-s})^{-1}=\sum_{\begin{smallmatrix}g\in\mathcal{G}\\ d_{-}(g)>m\end{smallmatrix}}\frac{\mu(g)}{\|g\|^{s}},

then one can see that

(34) C(n,m)qηn(P)m(1Pη)1C(n,m)\ll q^{\eta n}\prod_{\partial(P)\leq m}\left(1-{\|P\|^{-\eta}}\right)^{-1}

uniformly for n,m1n,m\geq 1. By the abstract prime number theorem (e.g. see [16, Chap. 8], [4]), for the product over (P)m\partial(P)\leq m we have that

(P)mlog(1Pη)(P)mPη(P)m1kmqk/kqm.-\sum_{\partial(P)\leq m}\log(1-\|P\|^{-\eta})\ll\sum_{\partial(P)\leq m}\|P\|^{-\eta}\ll\sum_{\partial(P)\leq m}1\ll\sum_{k\leq m}q^{k}/k\ll q^{m}.

Thus, the desired result (31) follows. As for the estimate (32), we have

M(n,m)=nC(,m)exp(qm)nqη=qηnexp(qm)nqη(n)qηnexp(qm).M(n,m)=\sum_{\ell\leq n}C(\ell,m)\ll\mathrm{exp}(q^{m})\sum_{\ell\leq n}q^{\eta\ell}=q^{\eta n}\mathrm{exp}(q^{m})\sum_{\ell\leq n}q^{\eta(\ell-n)}\ll q^{\eta n}\mathrm{exp}(q^{m}).

Lemma 2.6.

For any bounded function ff, we have

(35) (g)nμ(g)1𝔇(𝒢,S)(g)f(d(g))=Of,𝒢(qnloglogn).\sum_{\partial(g)\leq n}\mu(g)1_{\mathfrak{D}(\mathcal{G},S)}(g)f(d_{-}(g))=O_{f,\mathcal{G}}\left(\frac{q^{n}}{\log\log n}\right).
Proof.

First, we break up the sum based on the degree of the minimal prime factor Pmin(g)P_{\min}(g) of g𝔇(𝒢,S)g\in\mathfrak{D}(\mathcal{G},S).

(g)nμ(g)1𝔇(𝒢,S)(g)f(d(g))\displaystyle\sum_{\partial(g)\leq n}\mu(g)1_{\mathfrak{D}(\mathcal{G},S)}(g)f(d_{-}(g)) =(P)nPSf((P))(g)nPmin(g)=Pμ(g)\displaystyle=\sum_{\begin{smallmatrix}\partial(P)\leq n\\ P\in S\end{smallmatrix}}f(\partial(P))\sum_{\begin{smallmatrix}\partial(g)\leq n\\ P_{\min}(g)=P\end{smallmatrix}}\mu(g)
=(P)mPSf((P))(g)nPmin(g)=Pμ(g)\displaystyle=\sum_{\begin{smallmatrix}\partial(P)\leq m\\ P\in S\end{smallmatrix}}f(\partial(P))\sum_{\begin{smallmatrix}\partial(g)\leq n\\ P_{\min}(g)=P\end{smallmatrix}}\mu(g)
+m<(P)nPSf((P))(g)nPmin(g)=Pμ(g)\displaystyle\qquad+\sum_{\begin{smallmatrix}m<\partial(P)\leq n\\ P\in S\end{smallmatrix}}f(\partial(P))\sum_{\begin{smallmatrix}\partial(g)\leq n\\ P_{\min}(g)=P\end{smallmatrix}}\mu(g)
(36) \colonequalsS1+S2,\displaystyle\colonequals S_{1}+S_{2},

where mm is to be chosen later. For the inside sum, observe that

(37) (g)nPmin(g)=Pμ(g)=(g)(P)nd(g)>(P)μ(g).\sum_{\begin{smallmatrix}\partial(g)\leq n\\ P_{\min}(g)=P\end{smallmatrix}}\mu(g)=-\sum_{\begin{smallmatrix}\partial(g)-\partial(P)\leq n\\ d_{-}(g)>\partial(P)\end{smallmatrix}}\mu(g).

Thus, by (37), Lemma 2.5 and the abstract prime number theorem for 𝒢\mathcal{G}, for S1S_{1} we have

(38) S1\displaystyle S_{1} =(P)mPSf((P))M(n(P),(P))\displaystyle=-\sum_{\begin{smallmatrix}\partial(P)\leq m\\ P\in S\end{smallmatrix}}f(\partial(P))M(n-\partial(P),\partial(P))
(P)m|M(n(P),(P))|\displaystyle\ll\sum_{\partial(P)\leq m}|M(n-\partial(P),\partial(P))|
(P)mqη(n(P))exp(q(P))\displaystyle\ll\sum_{\partial(P)\leq m}q^{\eta(n-\partial(P))}\exp\left(q^{\partial(P)}\right)
qηn1kmqkkqηkexp(qk)\displaystyle\ll q^{\eta n}\sum_{1\leq k\leq m}\frac{q^{k}}{k}\cdot q^{-\eta k}\exp\left(q^{k}\right)
(39) qηn+(1η)mexp(qm)/m.\displaystyle\ll q^{\eta n+(1-\eta)m}\exp\left(q^{m}\right)/m.

For S2S_{2}, we have

(40) S2m<(P)n(g)nPmin(g)=P1Φ(n,m),S_{2}\ll\sum_{m<\partial(P)\leq n}\sum_{\begin{smallmatrix}\partial(g)\leq n\\ P_{\min}(g)=P\end{smallmatrix}}1\leq\Phi(n,m),

where Φ(n,m)\colonequals(g)nd(g)>m1\Phi(n,m)\colonequals\sum_{\begin{smallmatrix}\partial(g)\leq n\\ d_{-}(g)>m\end{smallmatrix}}1 for m,n1m,n\geq 1. Moreover, by the sieve of Eratosthenes, we have

(41) Φ(n,m)=c𝒢qn(P)m(1P1)+O(n 2c𝒢qm).\Phi(n,m)=c_{\mathcal{G}}q^{n}\prod_{\partial(P)\leq m}\left(1-\|P\|^{-1}\right)+O\left(n\,2^{c_{\mathcal{G}}q^{m}}\right).

Again by the abstract prime number theorem for 𝒢\mathcal{G}, with a standard computation we obtain that

(42) Φ(n,m)qnm+n 2c𝒢qm.\Phi(n,m)\ll\frac{q^{n}}{m}+n\,2^{c_{\mathcal{G}}q^{m}}.

Taking m=[loglogn]m=[\log\log n] and combining (2.1), (38), (40), and (42) together, the desired estimate (35) follows. ∎

Now we are ready to prove “(24)(25)\eqref{Eqn: equidis of largest prime factors}\Leftrightarrow\eqref{Eqn: analogue of main result in global function feild}”.

Proof of “(24)(25)\eqref{Eqn: equidis of largest prime factors}\Leftrightarrow\eqref{Eqn: analogue of main result in global function feild}”.

By Lemma 2.2, we have

δ(S)c𝒢qn\displaystyle-\delta(S)c_{\mathcal{G}}q^{n} (g)=nQS(g)f(d+(g))\displaystyle\sim-\sum_{\partial(g)=n}Q_{S}(g)f(d^{+}(g))
=(g)=nghμ(h)1𝔇(𝒢,S)(h)f(d(h))\displaystyle=\sum_{\partial(g)=n}\sum_{g\geq h}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))
=(h)nμ(h)1𝔇(𝒢,S)(h)f(d(h))(r)=n(h)1\displaystyle=\sum_{\partial(h)\leq n}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))\sum_{\partial(r)=n-\partial(h)}1
=(h)nμ(h)1𝔇(𝒢,S)(h)f(d(h))(c𝒢qnh+O(qη(n(h))))\displaystyle=\sum_{\partial(h)\leq n}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))\left(c_{\mathcal{G}}\frac{q^{n}}{\|h\|}+O(q^{\eta(n-\partial(h))})\right)
=(h)nμ(h)1𝔇(𝒢,S)(h)f(d(h))hc𝒢qn+(h)nμ(h)1𝔇(𝒢,S)(h)f(d(h))O(qη(n(h)))\displaystyle=\sum_{\partial(h)\leq n}\frac{\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))}{\|h\|}c_{\mathcal{G}}q^{n}+\sum_{\partial(h)\leq n}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))O(q^{\eta(n-\partial(h))})
(43) \colonequalsS3+S4.\displaystyle\colonequals S_{3}+S_{4}.

To prove “(24)(25)\eqref{Eqn: equidis of largest prime factors}\Leftrightarrow\eqref{Eqn: analogue of main result in global function feild}”, it suffices to show that S4=o(qn)S_{4}=o(q^{n}). In fact, for S4S_{4} we have

S4\displaystyle S_{4} =(h)nμ(h)1𝔇(𝒢,S)(h)f(d(h))O(qη(n(h)))\displaystyle=\sum_{\partial(h)\leq n}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))O(q^{\eta(n-\partial(h))})
=(h)nN0μ(h)1𝔇(𝒢,S)(h)f(d(h))O(qη(n(h)))+N0<(h)nμ(h)1𝔇(𝒢,S)(h)f(d(h))O(qη(n(h)))\displaystyle=\sum_{\partial(h)\leq n-N_{0}}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))O(q^{\eta(n-\partial(h))})+\sum_{N_{0}<\partial(h)\leq n}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(d_{-}(h))O(q^{\eta(n-\partial(h))})
\colonequalsS5+S6.\displaystyle\colonequals S_{5}+S_{6}.

It follows from Lemma 2.6 and the straightforward calculations by taking N0=(η+ε)n<nN_{0}=(\eta+\varepsilon)n<n for arbitrary ε>0\varepsilon>0 (for example, we can take ε=(1η)/2>0\varepsilon=(1-\eta)/2>0) that S5=o(qn)S_{5}=o(q^{n}) and S6=o(qn)S_{6}=o(q^{n}) as desired. ∎

2.2. Proof of “(25)(26)\eqref{Eqn: analogue of main result in global function feild}\Leftrightarrow\eqref{Eqn: to write into partial sum}

Our proof relies on the estimate of R(n,m)R(n,m) defined by

(44) R(n,m)\colonequals0(g)nd(g)>mμ(g)g,R(n,m)\colonequals\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ d_{-}(g)>m\end{smallmatrix}}\frac{\mu(g)}{\|g\|},

where m,nm,n are a pair of non-negative integers. We set Pmin(e𝒢)=P_{\min}(e_{\mathcal{G}})=\infty for convenience. First, we give an elementary bound for R(n,m)R(n,m) in the following lemma.

Lemma 2.7.

For any n,m0n,m\geq 0, we have that

(45) R(n,m)=O(1).R(n,m)=O(1).
Proof.

The approach is similar to that in [28]. For m0m\in\mathbb{Z}_{\geq 0}, define 𝒫m\colonequals{P(P)>m}\mathcal{P}_{m}\colonequals\{P\mid\partial(P)>m\} and 𝒫mc\colonequals{P(P)m}\mathcal{P}^{c}_{m}\colonequals\{P\mid\partial(P)\leq m\}. Let 𝒫m\langle\mathcal{P}_{m}\rangle and 𝒫mc\langle\mathcal{P}^{c}_{m}\rangle be monoids generated by 𝒫m\mathcal{P}_{m} and 𝒫mc\mathcal{P}^{c}_{m}, respectively. It is easy to check that

(46) 1𝒫mc(g)=r|gr𝒫mμ(r).1_{\langle\mathcal{P}^{c}_{m}\rangle}(g)=\sum_{\begin{subarray}{c}r|g\\ r\in\langle\mathcal{P}_{m}\rangle\end{subarray}}\mu(r).

Next, we consider the elements gg of degree nn and by (46) we have that

(g)=ng𝒫mc1=\displaystyle\sum_{\begin{subarray}{c}\partial(g)=n\\ g\in\langle\mathcal{P}^{c}_{m}\rangle\end{subarray}}1= (g)=nr|gr𝒫mμ(r)\displaystyle\sum_{\partial(g)=n}\sum_{\begin{subarray}{c}r|g\\ r\in\langle\mathcal{P}_{m}\rangle\end{subarray}}\mu(r)
=\displaystyle= (r)nr𝒫mμ(r)(h)=n(r)1\displaystyle\sum_{\begin{subarray}{c}\partial(r)\leq n\\ r\in\langle\mathcal{P}_{m}\rangle\end{subarray}}\mu(r)\sum_{\partial(h)=n-\partial(r)}1
=\displaystyle= (r)nr𝒫mμ(r)(c𝒢qn(r)+c0qη(n(r)))\displaystyle\sum_{\begin{subarray}{c}\partial(r)\leq n\\ r\in\langle\mathcal{P}_{m}\rangle\end{subarray}}\mu(r)\left(c_{\mathcal{G}}q^{n-\partial(r)}+c_{0}q^{\eta(n-\partial(r))}\right)
=\displaystyle= c𝒢qn(r)nr𝒫mμ(r)r+c0qηn(r)nr𝒫mμ(r)rη,\displaystyle c_{\mathcal{G}}q^{n}\sum_{\begin{subarray}{c}\partial(r)\leq n\\ r\in\langle\mathcal{P}_{m}\rangle\end{subarray}}\frac{\mu(r)}{\|r\|}+c_{0}q^{\eta n}\sum_{\begin{subarray}{c}\partial(r)\leq n\\ r\in\langle\mathcal{P}_{m}\rangle\end{subarray}}\frac{\mu(r)}{\|r\|^{\eta}},

where c0c_{0} is some constant number. It follows that

|(r)nr𝒫mμ(r)r|\displaystyle\Big{|}\sum_{\begin{subarray}{c}\partial(r)\leq n\\ r\in\langle\mathcal{P}_{m}\rangle\end{subarray}}\frac{\mu(r)}{\|r\|}\Big{|}\ll 1qn(g)=n1+q(η1)n(r)n1rη\displaystyle\frac{1}{q^{n}}\sum_{\partial(g)=n}1+q^{(\eta-1)n}\sum_{\partial(r)\leq n}\frac{1}{\|r\|^{\eta}}
=\displaystyle= c𝒢+O(q(η1)n)+q(η1)n(r)n1rη.\displaystyle c_{\mathcal{G}}+O(q^{(\eta-1)n})+q^{(\eta-1)n}\sum_{\partial(r)\leq n}\frac{1}{\|r\|^{\eta}}.
\displaystyle\ll q(η1)nk=0n(r)=k1rη\displaystyle q^{(\eta-1)n}\sum_{k=0}^{n}\sum_{\partial(r)=k}\frac{1}{\|r\|^{\eta}}
=\displaystyle= q(η1)nk=0n1qηk(c𝒢qk+O(qηk))\displaystyle q^{(\eta-1)n}\sum_{k=0}^{n}\frac{1}{q^{\eta k}}\left(c_{\mathcal{G}}q^{k}+O(q^{\eta k})\right)
\displaystyle\ll 1.\displaystyle 1.

Thus, this proves the lemma. ∎

In addition to the elementary bound for R(n,m)R(n,m) as above, we will also need a refined estimate of R(n,m)R(n,m) when 0mloglogn0\leq m\leq\log\log n for our purpose. More precisely, we will show the following lemma.

Lemma 2.8.

For 0mloglogn0\leq m\leq\log\log n and some constant 0η<10\leq\eta<1 only depending on 𝒢\mathcal{G}, we have that

(47) R(n,m)q(η1+ε)n,R(n,m)\ll q^{(\eta-1+\varepsilon)n},

where ε\varepsilon is an arbitrary positive number.

Proof.

We will separate two cases to discuss, i.e., m=0m=0 and m0m\neq 0. If m=0m=0, then we have

R(n,0)=(g)nμ(g)g.R(n,0)=\sum_{\begin{smallmatrix}\partial(g)\leq n\end{smallmatrix}}\frac{\mu(g)}{\|g\|}.

Assuming that n1n\geq 1, by the principle of inclusion-exclusion and a similar computation as in the proof of Lemma 2.7, we can get

0=(g)=nr|gμ(r)=c𝒢qn(r)nμ(r)r+c0qηn(r)nμ(r)rη\displaystyle 0=\sum_{\partial(g)=n}\sum_{r|g}\mu(r)=c_{\mathcal{G}}q^{n}\sum_{\partial(r)\leq n}\frac{\mu(r)}{\|r\|}+c_{0}q^{\eta n}\sum_{\partial(r)\leq n}\frac{\mu(r)}{\|r\|^{\eta}}

for some constant c0c_{0}. It follows that

(48) (r)nμ(r)r=c0c𝒢q(η1)nk=0n(r)=kμ(r)rη.\sum_{\partial(r)\leq n}\frac{\mu(r)}{\|r\|}=-\frac{c_{0}}{c_{\mathcal{G}}}q^{(\eta-1)n}\sum_{k=0}^{n}\sum_{\partial(r)=k}\frac{\mu(r)}{\|r\|^{\eta}}.

Consider k1k\geq 1 and by Lemma 2.4, then the inside summation

(49) (r)=kμ(r)rη=1qηk(r)=kμ(r)1.\sum_{\partial(r)=k}\frac{\mu(r)}{\|r\|^{\eta}}=\frac{1}{q^{\eta k}}\sum_{\partial(r)=k}\mu(r)\ll 1.

It is clear that (49) also holds for k=0k=0. Thus, we have

k=0n(r)=kμ(r)rηn.\displaystyle\sum_{k=0}^{n}\sum_{\partial(r)=k}\frac{\mu(r)}{\|r\|^{\eta}}\ll n.

Therefore, by (48) and (49) we obtain that

(50) (r)nμ(r)r=O(q(η1+ε)n),\sum_{\partial(r)\leq n}\frac{\mu(r)}{\|r\|}=O(q^{(\eta-1+\varepsilon)n}),

where ε\varepsilon is an arbitrary positive number. This proves the desired (47) for the special case m=0m=0.

Now suppose that 1mloglogn1\leq m\leq\log\log n, by Lemma 2.5 we have that

(51) M(n,m)qηn.M(n,m)\ll q^{\eta n}.

It follows from [17, Proposition (1.2.1)] that ζ𝒢(s)\zeta_{\mathcal{G}}(s) has a simple pole at s=1s=1. And so we have that

(52) (g)0μ(g)g=lims11ζ𝒢(s)=0.\sum_{\partial(g)\geq 0}\frac{\mu(g)}{\|g\|}=\lim_{s\to 1}\frac{1}{\zeta_{\mathcal{G}}(s)}=0.

Let 𝒫mc={P(P)m}\mathcal{P}^{c}_{m}=\{P\mid\partial(P)\leq m\} as in the proof of Lemma 2.7. It is clear that 𝒫mc\mathcal{P}^{c}_{m} is a finite set. It implies that

(53) (g)0d(g)>mμ(g)g=((g)0μ(g)g)(P𝒫mc(11P)1)=0.\sum_{\begin{subarray}{c}\partial(g)\geq 0\\ d_{-}(g)>m\end{subarray}}\frac{\mu(g)}{\|g\|}=\left(\sum_{\partial(g)\geq 0}\frac{\mu(g)}{\|g\|}\right)\cdot\left(\prod_{P\in\mathcal{P}^{c}_{m}}\left(1-\frac{1}{\|P\|}\right)^{-1}\right)=0.

Finally, we show the desired result by using (51), (53) and summation by parts. In particular, let yy be fixed. To apply summation by parts, we put

(54) My(x)=0(g)xd(g)>yμ(g).M_{y}(x)=\sum_{\begin{subarray}{c}0\leq\partial(g)\leq x\\ d_{-}(g)>y\end{subarray}}\mu(g).

It follows that

R(x,y)=\displaystyle R(x,y)= (g)>xd(g)>yμ(g)g=xdMy(t)qt\displaystyle-\sum_{\begin{smallmatrix}\partial(g)>x\\ d_{-}(g)>y\end{smallmatrix}}\frac{\mu(g)}{\|g\|}=-\int_{x}^{\infty}\frac{dM_{y}(t)}{q^{t}}
=\displaystyle= My(x)qxlnqxMy(t)qt𝑑t\displaystyle\frac{M_{y}(x)}{q^{x}}-\ln q\int_{x}^{\infty}\dfrac{M_{y}(t)}{q^{t}}\,dt
\displaystyle\ll q(η1)x+lnqxexp(ct)𝑑t\displaystyle q^{(\eta-1)x}+\ln q\int_{x}^{\infty}\mathrm{exp}(-c\sqrt{t})\,dt
\displaystyle\ll q(η1+ε)x,\displaystyle q^{(\eta-1+\varepsilon)x},

where ε\varepsilon is an arbitrary positive number. Hence, for 1mloglogn1\leq m\leq\log\log n we also obtain R(n,m)q(η1+ε)nR(n,m)\ll q^{(\eta-1+\varepsilon)n} as desired. In conclusion, we complete the proof of the lemma. ∎

Now we are ready to prove “(25)(26)\eqref{Eqn: analogue of main result in global function feild}\Leftrightarrow\eqref{Eqn: to write into partial sum}”.

Proof of “(25)(26)\eqref{Eqn: analogue of main result in global function feild}\Leftrightarrow\eqref{Eqn: to write into partial sum}”.

Set f()=0f(\infty)=0 for convenience. First, we break up the partial sum of (26) into two sums:

0(g)ng𝔇(𝒢,S)μa(g)gf(d(g))=\displaystyle\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)}{\|g\|}f(d_{-}(g))= 0(g)ng𝔇(𝒢,S)f(d(g))gr|gμ(r)a(gr)\displaystyle\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{f(d_{-}(g))}{\|g\|}\sum_{r|g}\mu(r)a(g-r)
=\displaystyle= 0(g)ng𝔇(𝒢,S)f(d(g))gμ(g)+0(g)ng𝔇(𝒢,S)f(d(g))gr|grgμ(r)a(gr)\displaystyle\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{f(d_{-}(g))}{\|g\|}\mu(g)+\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{f(d_{-}(g))}{\|g\|}\sum_{\begin{subarray}{c}r|g\\ r\neq g\end{subarray}}\mu(r)a(g-r)
\colonequals\displaystyle\colonequals S7+S8.\displaystyle S_{7}+S_{8}.

To show “(25)(26)\eqref{Eqn: analogue of main result in global function feild}\Leftrightarrow\eqref{Eqn: to write into partial sum}”, it suffices to show S8=o(1)S_{8}=o(1). Let g=g+Pmin(g)g=g^{\prime}+P_{\min}(g). Then we have

S8=\displaystyle S_{8}= 0(g)ng𝔇(𝒢,S)f(d(g))gr|grgμ(r)a(gr)\displaystyle\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{f(d_{-}(g))}{\|g\|}\sum_{\begin{subarray}{c}r|g\\ r\neq g\end{subarray}}\mu(r)a(g-r)
=\displaystyle= 0(Pmin(g))nPmin(g)Sf((Pmin(g)))Pmin(g)0(g)n(Pmin(g))d(g)>(Pmin(g))1gr|gPmin(g)|rrgμ(r)a(g+Pmin(g)r)\displaystyle\sum_{\begin{subarray}{c}0\leq\partial(P_{\min}(g))\leq n\\ P_{\min}(g)\in S\end{subarray}}\frac{f(\partial(P_{\min}(g)))}{\|P_{\min}(g)\|}\sum_{\begin{smallmatrix}0\leq\partial(g^{\prime})\leq n-\partial(P_{\min}(g))\\ d_{-}(g^{\prime})>\partial(P_{\min}(g))\end{smallmatrix}}\frac{1}{\|g^{\prime}\|}\sum_{\begin{subarray}{c}r|g\\ P_{\min}(g)|r\\ r\neq g\end{subarray}}\mu(r)a(g^{\prime}+P_{\min}(g)-r)
+0(Pmin(g))nPmin(g)Sf((Pmin(g)))Pmin(g)0(g)n(Pmin(g))d(g)>(Pmin(g))1gr|gμ(r)a(g+Pmin(g)r)\displaystyle+\sum_{\begin{subarray}{c}0\leq\partial(P_{\min}(g))\leq n\\ P_{\min}(g)\in S\end{subarray}}\frac{f(\partial(P_{\min}(g)))}{\|P_{\min}(g)\|}\sum_{\begin{smallmatrix}0\leq\partial(g^{\prime})\leq n-\partial(P_{\min}(g))\\ d_{-}(g^{\prime})>\partial(P_{\min}(g))\end{smallmatrix}}\frac{1}{\|g^{\prime}\|}\sum_{\begin{subarray}{c}r|g^{\prime}\end{subarray}}\mu(r)a(g^{\prime}+P_{\min}(g)-r)
\colonequals\displaystyle\colonequals S9+S10.\displaystyle S_{9}+S_{10}.

Next, we will show that both S9S_{9} and S10S_{10} are error terms of size o(1)o(1).

  1. (1)

    For S9S_{9}, let r=Pmin(g)+rr=P_{\min}(g)+r^{\prime} and g=h+rg^{\prime}=h^{\prime}+r^{\prime} and then exchange the order of the first and the second summations. Since the function aa only supports on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S), we have that

    S9=\displaystyle S_{9}= 0(Pmin(g))nPmin(g)Sf((Pmin(g)))Pmin(g)1(h)n(Pmin(g))d(h)>(Pmin(g))h𝔇(𝒢,S)a(h)h1(r)n(Pmin(g))(h)d(r)>(Pmin(g))μ(r)r\displaystyle-\sum_{\begin{subarray}{c}0\leq\partial(P_{\min}(g))\leq n\\ P_{\min}(g)\in S\end{subarray}}\frac{f(\partial(P_{\min}(g)))}{\|P_{\min}(g)\|}\sum_{\begin{subarray}{c}1\leq\partial(h^{\prime})\leq n-\partial(P_{\min}(g))\\ d_{-}(h^{\prime})>\partial(P_{\min}(g))\\ h^{\prime}\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{a(h^{\prime})}{\|h^{\prime}\|}\sum_{\begin{subarray}{c}1\leq\partial(r^{\prime})\leq n-\partial(P_{\min}(g))-\partial(h^{\prime})\\ d_{-}(r^{\prime})>\partial(P_{\min}(g))\end{subarray}}\frac{\mu(r^{\prime})}{\|r^{\prime}\|}
    =\displaystyle= 1(h)nh𝔇(𝒢,S)a(h)h1(Pmin(g))n(h)(Pmin(g))<d(h)Pmin(g)Sf((Pmin(g)))Pmin(g)1(r)n(Pmin(g))(h)d(r)>(Pmin(g))μ(r)r\displaystyle-\sum_{\begin{subarray}{c}1\leq\partial(h^{\prime})\leq n\\ h^{\prime}\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{a(h^{\prime})}{\|h^{\prime}\|}\sum_{\begin{subarray}{c}1\leq\partial(P_{\min}(g))\leq n-\partial(h^{\prime})\\ \partial(P_{\min}(g))<d_{-}(h^{\prime})\\ P_{\min}(g)\in S\end{subarray}}\frac{f(\partial(P_{\min}(g)))}{\|P_{\min}(g)\|}\sum_{\begin{subarray}{c}1\leq\partial(r^{\prime})\leq n-\partial(P_{\min}(g))-\partial(h^{\prime})\\ d_{-}(r^{\prime})>\partial(P_{\min}(g))\end{subarray}}\frac{\mu(r^{\prime})}{\|r^{\prime}\|}
    \displaystyle\ll 1(h)[loglogn]h𝔇(𝒢,S)|a(h)|h1(Pmin(g))n(h)(Pmin(g))<d(h)Pmin(g)S1Pmin(g)|R(n(Pmin(g))(h),(Pmin(g)))|\displaystyle\sum_{\begin{subarray}{c}1\leq\partial(h^{\prime})\leq[\log\log n]\\ h^{\prime}\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{|a(h^{\prime})|}{\|h^{\prime}\|}\sum_{\begin{subarray}{c}1\leq\partial(P_{\min}(g))\leq n-\partial(h^{\prime})\\ \partial(P_{\min}(g))<d_{-}(h^{\prime})\\ P_{\min}(g)\in S\end{subarray}}\frac{1}{\|P_{\min}(g)\|}|R(n-\partial(P_{\min}(g))-\partial(h^{\prime}),\partial(P_{\min}(g)))|
    +\displaystyle+ [loglogn]+1(h)nh𝔇(𝒢,S)|a(h)|h1(Pmin(g))n(h)(Pmin(g))<d(h)Pmin(g)S1Pmin(g)\displaystyle\sum_{\begin{subarray}{c}[\log\log n]+1\leq\partial(h^{\prime})\leq n\\ h^{\prime}\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{|a(h^{\prime})|}{\|h^{\prime}\|}\sum_{\begin{subarray}{c}1\leq\partial(P_{\min}(g))\leq n-\partial(h^{\prime})\\ \partial(P_{\min}(g))<d_{-}(h^{\prime})\\ P_{\min}(g)\in S\end{subarray}}\frac{1}{\|P_{\min}(g)\|}
    \displaystyle\ll qc(n2loglogn)1(h)[loglogn]h𝔇(𝒢,S)|a(h)|hloglogh+[loglogn]+1(h)nh𝔇(𝒢,S)|a(h)|hloglogh\displaystyle q^{-c(n-2\log\log n)}\sum_{\begin{subarray}{c}1\leq\partial(h^{\prime})\leq[\log\log n]\\ h^{\prime}\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{|a(h^{\prime})|}{\|h^{\prime}\|}\log\log\|h^{\prime}\|+\sum_{\begin{subarray}{c}[\log\log n]+1\leq\partial(h^{\prime})\leq n\\ h^{\prime}\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{|a(h^{\prime})|}{\|h^{\prime}\|}\log\log\|h^{\prime}\|
    =\displaystyle= O(qc(n2loglogn))+o(1)=o(1).\displaystyle O(q^{-c(n-2\log\log n)})+o(1)=o(1).

    Here the constant c=(η1+ε)>0c=-(\eta-1+\varepsilon)>0 by Lemma 2.8 and the last inequality follows from the assumption of aa as in Theorem 1.6.

  2. (2)

    Similarly for S10S_{10}, let g=h+rg^{\prime}=h+r and then we have that

    S10=\displaystyle S_{10}= 0(Pmin(g))nPmin(g)Sf((Pmin(g)))Pmin(g)1(h)n(Pmin(g))d(h)>(Pmin(g))orh=0a(h+Pmin(g))h1(r)n(Pmin(g))(h)d(r)>(Pmin(g))μ(r)r\displaystyle\sum_{\begin{subarray}{c}0\leq\partial(P_{\min}(g))\leq n\\ P_{\min}(g)\in S\end{subarray}}\frac{f(\partial(P_{\min}(g)))}{\|P_{\min}(g)\|}\sum_{\begin{subarray}{c}1\leq\partial(h)\leq n-\partial(P_{\min}(g))\\ d_{-}(h)>\partial(P_{\min}(g))\\ \text{or}\,h=0\end{subarray}}\frac{a(h+P_{\min}(g))}{\|h\|}\sum_{\begin{subarray}{c}1\leq\partial(r)\leq n-\partial(P_{\min}(g))-\partial(h)\\ d_{-}(r)>\partial(P_{\min}(g))\end{subarray}}\frac{\mu(r)}{\|r\|}
    =\displaystyle= 0(h)n0(Pmin(g))n(h)(Pmin(g))<d(h)Pmin(g)Sf((Pmin(g)))a(h+Pmin(g))Pmin(g)h1(r)n(Pmin(g))(h)d(r)>(Pmin(g))μ(r)r\displaystyle\sum_{0\leq\partial(h)\leq n}\sum_{\begin{subarray}{c}0\leq\partial(P_{\min}(g))\leq n-\partial(h)\\ \partial(P_{\min}(g))<d_{-}(h)\\ P_{\min}(g)\in S\end{subarray}}\frac{f(\partial(P_{\min}(g)))a(h+P_{\min}(g))}{\|P_{\min}(g)\|\|h\|}\sum_{\begin{subarray}{c}1\leq\partial(r)\leq n-\partial(P_{\min}(g))-\partial(h)\\ d_{-}(r)>\partial(P_{\min}(g))\\ \end{subarray}}\frac{\mu(r)}{\|r\|}
    \displaystyle\ll 0(h)[loglogn]1(Pmin(g))n(h)(Pmin(g))<d(h)Pmin(g)S|a(h+Pmin(g))|Pmin(g)h|R(n(Pmin(g))(h),(Pmin(g)))|\displaystyle\sum_{0\leq\partial(h)\leq[\log\log n]}\sum_{\begin{subarray}{c}1\leq\partial(P_{\min}(g))\leq n-\partial(h)\\ \partial(P_{\min}(g))<d_{-}(h)\\ P_{\min}(g)\in S\end{subarray}}\frac{|a(h+P_{\min}(g))|}{\|P_{\min}(g)\|\|h\|}|R(n-\partial(P_{\min}(g))-\partial(h),\partial(P_{\min}(g)))|
    +\displaystyle+ [loglogn]+2(g)ng𝔇(𝒢,S)|a(g)|g\displaystyle\sum_{\begin{subarray}{c}[\log\log n]+2\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{|a(g)|}{\|g\|}
    \displaystyle\ll qc(n2loglogn)1(g)ng𝔇(𝒢,S)|a(g)|g+[loglogn]+2(g)ng𝔇(𝒢,S)|a(g)|g\displaystyle q^{-c(n-2\log\log n)}\sum_{\begin{subarray}{c}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{|a(g)|}{\|g\|}+\sum_{\begin{subarray}{c}[\log\log n]+2\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{|a(g)|}{\|g\|}
    =\displaystyle= O(qc(n2loglogn))+o(1)=o(1).\displaystyle O(q^{-c(n-2\log\log n)})+o(1)=o(1).

Again, the constant c=(η1+ε)>0c=-(\eta-1+\varepsilon)>0 by Lemma 2.8 and the last inequality follows from the assumption of aa as in Theorem 1.6. Thus, we conclude that

(55) 0(g)ng𝔇(𝒢,S)μa(g)gf(d(g))=0(g)ng𝔇(𝒢,S)μ(g)gf(d(g))+o(1).\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)}{\|g\|}f(d_{-}(g))=\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu(g)}{\|g\|}f(d_{-}(g))+o(1).

2.3. Proof of Corollary 1.7

In this section, we prove Corollary 1.7 in details. For the convenience of the readers, we cite it here.

Corollary 2.9.

Suppose that a:𝒢a\colon\mathcal{G}\to\mathbb{C} is an arithmetic function supported on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) with a(e𝒢)=1a(e_{\mathcal{G}})=1 and |a(g)|gα|a(g)|\ll\|g\|^{-\alpha} for some α>0\alpha>0. If S𝒫S\subseteq\mathcal{P} has a natural density δ(S)\delta(S), then

(56) limn1(g)ng𝔇(𝒢,S)μa(g)φ(g)=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{\mu*a(g)}{\varphi(g)}=\delta(S).

Here, φ(g)\varphi(g) is Euler’s totient function defined by φ(g)=gP|g(11P).\varphi(g)=\|g\|\prod_{P|g}\left(1-\frac{1}{\|P\|}\right).

Proof.

The proof is similar to the approach in [32]. In particular, we will apply Theorem 1.6 to show (56). Define

(57) b(g)\colonequalsh|gμa(h)hφ(h).b(g)\colonequals\sum_{h|g}\mu*a(h)\frac{\|h\|}{\varphi(h)}.

Then by the Möbius inversion formula, we have

(58) μa(g)φ(g)=μb(g)g.\frac{\mu*a(g)}{\varphi(g)}=\frac{\mu*b(g)}{\|g\|}.

Clearly, b(e𝒢)=1b(e_{\mathcal{G}})=1. By Theorem 1.6, to prove (14), it suffices to show that

(59) 0(g)ng𝔇(𝒢,S)|b(g)|gloglogg<.\sum_{\begin{smallmatrix}0\leq\partial(g)\leq n\\ g\in\mathfrak{D}(\mathcal{G},S)\end{smallmatrix}}\frac{|b(g)|}{\|g\|}\log\log\|g\|<\infty.

By the definition of μa(r)\mu*a(r) and go through a similar argument for elements in 𝒢\mathcal{G} as in [32], we have

(60) b(g)=h|g,(h,gh)=0hφ(h)a(h)P|gh11P.b(g)=\sum_{h|g,(h,g-h)=0}\frac{\|h\|}{\varphi(h)}a(h)\prod_{\begin{smallmatrix}P|g-h\end{smallmatrix}}\frac{1}{1-\|P\|}.

Furthermore, we have that

(61) |b(g)|h|g,(h,gh)=0|a(h)|hφ(h)P|gh1P1.|b(g)|\leq\sum_{h|g,(h,g-h)=0}|a(h)|\frac{\|h\|}{\varphi(h)}\prod_{\begin{smallmatrix}P|g-h\end{smallmatrix}}\frac{1}{\|P\|-1}.

On the other hand, by the standard elementary technique, for every 0<α10<\alpha\leq 1 one can show that

(62) g(1α)φ(g)0as (g).\frac{\|g\|^{(1-\alpha)}}{\varphi(g)}\to 0\quad\mbox{as $\partial(g)\to\infty$}.

In particular, we have

(63) gφ(g)gα2.\frac{\|g\|}{\varphi(g)}\ll\|g\|^{\frac{\alpha}{2}}.

Thus, by (61) and the assumption that |a(g)|gα|a(g)|\ll\|g\|^{-\alpha} with α>0\alpha>0, we get the following estimate for b(g)b(g):

(64) |b(g)|r|grα2P|gr1P1.|b(g)|\leq\sum_{r|g}\|r\|^{-\frac{\alpha}{2}}\prod_{\begin{smallmatrix}P|g-r\end{smallmatrix}}\frac{1}{\|P\|-1}.

Put

(65) c(g)=r|grα2P|gr1P1.c(g)=\sum_{r|g}\|r\|^{-\frac{\alpha}{2}}\prod_{\begin{smallmatrix}P|g-r\end{smallmatrix}}\frac{1}{\|P\|-1}.

Then c(g)c(g) is the Dirichlet convolution of c1(g)=gα2c_{1}(g)=\|g\|^{-\frac{\alpha}{2}} and c2(g)=P|g1P1c_{2}(g)=\prod_{P|g}\frac{1}{\|P\|-1} . It is easy to see that

(66) (g)=0c1(g)gs=ζ𝒢(s+α2),\sum_{\partial(g)=0}^{\infty}\frac{c_{1}(g)}{\|g\|^{s}}=\zeta_{\mathcal{G}}(s+\frac{\alpha}{2}),

which is absolutely convergent on Re(s)>1α/2\mathrm{Re}(s)>1-\alpha/2 by the analytic property of ζ𝒢(s)\zeta_{\mathcal{G}}(s). Moreover, one can show that

(67) (g)=0c2(g)gs\sum_{\partial(g)=0}^{\infty}\frac{c_{2}(g)}{\|g\|^{s}}

is absolutely convergent on σ=Re(s)>0\sigma=\mathrm{Re}(s)>0. Observing that c2(g)c_{2}(g) is multiplicative, we have

(68) (g)N|c2(g)|gσ(P)N(1+n1|c2(nP)|Pnσ).\sum_{\partial(g)\leq N}\frac{|c_{2}(g)|}{\|g\|^{\sigma}}\leq\prod_{\partial(P)\leq N}\left(1+\sum_{n\geq 1}\frac{|c_{2}(nP)|}{\|P\|^{n\sigma}}\right).

On the other hand, we have

(P)0n1|c2(nP)|Pnσ=\displaystyle\sum_{\partial(P)\geq 0}\sum_{n\geq 1}\frac{|c_{2}(nP)|}{\|P\|^{n\sigma}}= P1(P1)(Pσ1)\displaystyle\sum_{P}\frac{1}{(\|P\|-1)(\|P\|^{\sigma}-1)}
=\displaystyle= m(P)=m1(qm1)(qmσ1)\displaystyle\sum_{m}\sum_{\partial(P)=m}\frac{1}{(q^{m}-1)(q^{m\sigma}-1)}
\displaystyle\ll m1qmσ<\displaystyle\sum_{m}\frac{1}{q^{m\sigma}}<\infty

for σ=Re(s)>0\sigma=\mathrm{Re}(s)>0. By the well known theorem of infinity product we conclude that (67) is absolutely convergent for Re(s)>0\mathrm{Re}(s)>0. It follows that on Re(s)>σ0\mathrm{Re}(s)>\sigma_{0}, where σ0=max{1α/2,0}<1\sigma_{0}=\max\{1-\alpha/2,0\}<1, we have

(69) (g)=0c(g)gs=ζ𝒢(s+α2)(g)=0c2(g)gs.\sum_{\partial(g)=0}^{\infty}\frac{c(g)}{\|g\|^{s}}=\zeta_{\mathcal{G}}(s+\frac{\alpha}{2})\sum_{\partial(g)=0}^{\infty}\frac{c_{2}(g)}{\|g\|^{s}}.

Therefore, the derivative of (g)=0c(g)gs\sum_{\partial(g)=0}^{\infty}c(g)\|g\|^{-s} is convergent at s=1s=1, which implies that

(70) (g)=0c(g)glogg<.\sum_{\partial(g)=0}^{\infty}\frac{c(g)}{\|g\|}\log\|g\|<\infty.

It follows immediately from (64) and (65) that

(71) (g)=0|b(g)|glogg<.\sum_{\partial(g)=0}^{\infty}\frac{|b(g)|}{\|g\|}\log\|g\|<\infty.

Hence, by (71) we obtain (59) and so the desired result (14) follows.

3. Proofs of main results for Axiom AA type arithmetical semigroups

In this section, we discuss the situation of Axiom AA type arithmetical semigroups. Similarly as in Sect. 2, there is also a key result, Theorem 3.7, and then Theorem 1.3 follows as an immediate consequence. Throughout this section, we fix 𝒢\mathcal{G} to be an Axiom AA type arithmetical semigroup, and adopt the definitions and notations in Sect. 1.1.

For any element g𝒢g\in\mathcal{G}, we take N+(g)\colonequalsmax{P:P|g}\mathrm{N}^{+}(g)\colonequals\max\left\{\left\|P\right\|:P|g\right\} to be the largest norm of all prime factors of gg, and N+(e𝒢)=1\mathrm{N}^{+}(e_{\mathcal{G}})=1. We also define the number QS(g)Q_{S}(g) in the same way for Axiom A#A^{\#} type arithmetical semigroups. More precisely, we let

(72) QS(g)\colonequals#{PS:P=N+(g),P|g}Q_{S}(g)\colonequals\#\left\{P\in S:\left\|P\right\|=\mathrm{N}^{+}(g),P|g\right\}

be the number of prime factors of gg in SS attaining the maximal norm N+(g)\mathrm{N}^{+}(g). Since most of proofs of our results in this section are very similar with those for the case of Axiom A#A^{\#} type arithmetical semigroup, we will just describe them briefly and only focus on key steps. First, we have the following equidistribution result of the largest (in terms of norm \|\cdot\|) prime factors of elements in 𝒢\mathcal{G}.

Lemma 3.1.

If S𝒫S\subseteq\mathcal{P} has a natural density δ(S)\delta(S), then

(73) 2gxQS(g)c𝒢δ(S)x.\sum_{2\leq\left\|g\right\|\leq x}Q_{S}(g)\sim c_{\mathcal{G}}\,\delta(S)x.
Proof.

The proof is similar as that for [18, Theorem 3.1]. ∎

Moreover, with an analogous argument as that for [27, Lemma 2.1] we have the following duality lemma for Axiom AA type arithmetical semigroups.

Lemma 3.2.

Suppose that f:f:\mathbb{N}\to\mathbb{C} is an arithmetic function with f(1)=0f(1)=0. Then for any g𝒢g\in\mathcal{G} we have

(74) h|gμ(h)1𝔇(𝒢,S)(h)f(N(h))=QS(g)f(N+(g)).\sum_{h|g}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(\mathrm{N}_{-}(h))=-Q_{S}(g)f(\mathrm{N}^{+}(g)).

Here 1𝔇(𝒢,S)1_{\mathfrak{D}(\mathcal{G},S)} is the indicator function on 𝔇(𝒢,S)\mathfrak{D}(\mathcal{G},S) defined as in (5).

The following two lemmas is giving Axiom AA type arithmetical semigroups analogues of bounds for partial sums involving the Möbius function that we will need in our proofs.

Lemma 3.3.

For some positive constant cc we have

(75) gxμ(g)=\displaystyle\sum_{\left\|g\right\|\leq x}\mu(g)= O(xexp{c(logx)1/2}),\displaystyle O\big{(}x\,\mathrm{exp}\{-c(\log x)^{1/2}\}\big{)},
(76) gxμ(g)g=\displaystyle\sum_{\left\|g\right\|\leq x}\frac{\mu(g)}{\left\|g\right\|}= O(exp{c(logx)1/2}).\displaystyle O\big{(}\mathrm{exp}\{-c(\log x)^{1/2}\}\big{)}.
Proof.

This directly generalizes [27, Lemma 2.2], which indicates that the proof relies on the zero-free region for the zeta function ζ𝒢(s)\zeta_{\mathcal{G}}(s). In particular, the analogous zero-free region exists for ζ𝒢(s)\zeta_{\mathcal{G}}(s); see [8, Theorem 1]. Thus we can use the classical Perron-type argument as in [11, Theorem 2.1] to obtain the desired results. ∎

Lemma 3.4.

For any x,y1x,y\geq 1, define

(77) M(x,y)\colonequalsgxN(g)>yμ(g).M(x,y)\colonequals\sum_{\begin{subarray}{c}\left\|g\right\|\leq x\\ \mathrm{N}_{-}(g)>y\end{subarray}}\mu(g).

There is a positive constant cc, which depends only on 𝒢\mathcal{G}, such that

(78) M(x,y)xexp(c(logx)1/2)Py(1P1/2)1,M(x,y)\ll x\,\mathrm{exp}(-c(\log x)^{1/2})\prod_{\left\|P\right\|\leq y}(1-\left\|P\right\|^{-1/2})^{-1},

uniformly for x,y1x,y\geq 1. In particular, we have

(79) M(x,y)xexp(c(logx)1/2+y).M(x,y)\ll x\,\mathrm{exp}(-c(\log x)^{1/2}+y).
Proof.

By using the same zero-free region as in Lemma 3.3 and the similar argument as in [18, Lemma 4.2], the desired result follows. ∎

To prove Theorem 3.7, we will also need a more general version of Axer’s theorem.

Theorem 3.5 (Axer [2]).

If ff is a complex-valued function on 𝒢\mathcal{G} satisfying

gx|f(g)|=O(x),gxf(g)=o(x),\sum_{\left\|g\right\|\leq x}|f(g)|=O(x),\qquad\sum_{\left\|g\right\|\leq x}f(g)=o(x),

then

gx(c𝒢xg𝒩(xg))f(g)=o(x).\sum_{\left\|g\right\|\leq x}\left(c_{\mathcal{G}}\frac{x}{\left\|g\right\|}-\mathcal{N}\left(\frac{x}{\left\|g\right\|}\right)\right)f(g)=o(x).

Next, with a similar process as in the proof for [18, Lemma 4.2] by using Lemma 3.4, we can get the following intermediate bound, which is needed to verify the conditions of Axer’s theorem when we apply it to prove Theorem 3.7.

Lemma 3.6.

For any bounded function ff, we have

(80) gxμ(g)1𝔇(𝒢,S)(g)f(N(g))=Of,𝒢(xloglogx).\sum_{\left\|g\right\|\leq x}\mu(g)1_{\mathfrak{D}(\mathcal{G},S)}(g)f(\mathrm{N}_{-}(g))=O_{f,\mathcal{G}}\left(\frac{x}{\log\log x}\right).

Now, we are ready to prove the following intermediate theorem.

Theorem 3.7.

Let ff be any bounded arithmetic function with f(1)=0f(1)=0. Let aa be an arithmetic function as in Theorem 1.3. The followings are equivalent:

(81) gxQS(g)f(N+(g))c𝒢δ(S)x.\sum_{\left\|g\right\|\leq x}Q_{S}(g)f(\mathrm{N}^{+}(g))\sim c_{\mathcal{G}}\delta(S)x.
(82) limx2gxg𝔇(𝒢,S)μ(g)f(N(g))g=δ(S).-\lim_{x\to\infty}\sum_{\begin{subarray}{c}2\leq\left\|g\right\|\leq x\\ g\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{\mu(g)f(\mathrm{N}_{-}(g))}{\left\|g\right\|}=\delta(S).
(83) limx2gxg𝔇(𝒢,S)μa(g)f(N(g))g=δ(S).-\lim_{x\to\infty}\sum_{\begin{subarray}{c}2\leq\left\|g\right\|\leq x\\ g\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{\mu*a(g)f(\mathrm{N}_{-}(g))}{\left\|g\right\|}=\delta(S).
Proof.

(81)(82)\eqref{Eqn: equidis of largest prime factors_Axiom A}\Leftrightarrow\eqref{Eqn: analogue of main result in global function feild_Axiom A}”: It is easy to see that the hypotheses of Theorem 3.5 are satisfied since the estimate (3) of 𝒩(x)\mathcal{N}(x), Lemma 3.6 and the assumption that ff is bounded. Then the desired result follows from Lemma 3.2 and Theorem 3.5 by applying the analogous approach in [18, Lemma 4.3]. In fact, we have

c𝒢δ(S)xgxQS(g)f(N+(g))=\displaystyle c_{\mathcal{G}}\delta(S)x\sim\sum_{\left\|g\right\|\leq x}Q_{S}(g)f(\mathrm{N}^{+}(g))= gxh|gμ(h)1𝔇(𝒢,S)(h)f(N(h))\displaystyle-\sum_{\left\|g\right\|\leq x}\sum_{h|g}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(\mathrm{N}_{-}(h))
=\displaystyle= hxμ(h)1𝔇(𝒢,S)(h)f(N(h))gx/h1\displaystyle-\sum_{\left\|h\right\|\leq x}\mu(h)1_{\mathfrak{D}(\mathcal{G},S)}(h)f(\mathrm{N}_{-}(h))\sum_{\left\|g\right\|\leq x/\left\|h\right\|}1
=\displaystyle= c𝒢x2hxh𝔇(𝒢,S)μ(h)f(N(h))h+o(x).\displaystyle-c_{\mathcal{G}}x\sum_{\begin{subarray}{c}2\leq\left\|h\right\|\leq x\\ h\in\mathfrak{D}(\mathcal{G},S)\end{subarray}}\frac{\mu(h)f(\mathrm{N}_{-}(h))}{\left\|h\right\|}+o(x).

(82)(83)\eqref{Eqn: analogue of main result in global function feild_Axiom A}\Leftrightarrow\eqref{Eqn: to write into partial sum_Axiom A}”: For any x,y1x,y\geq 1, we define

(84) R(x,y)\colonequals1gxN(g)>yμ(g)g.R(x,y)\colonequals\sum_{\begin{subarray}{c}1\leq\left\|g\right\|\leq x\\ \mathrm{N}_{-}(g)>y\end{subarray}}\frac{\mu(g)}{\left\|g\right\|}.

By the analogous arguments as in the proof of Lemma 2.7 we can obtain the same elementary bound for R(x,y)R(x,y). That is, for any x,y1x,y\geq 1 we have

(85) R(x,y)=O(1).R(x,y)=O(1).

Furthermore, for 1yloglogx1\leq y\leq\log\log x, we can obtain the following refined estimate

(86) R(x,y)exp(c(logx)1/2),R(x,y)\ll\mathrm{exp}(-c(\log x)^{1/2}),

where the positive constant cc depends only on 𝒢\mathcal{G}. In fact, for y=1y=1, the estimate (86) is just (76) in Lemma 3.3. As for 1<yloglogx1<y\leq\log\log x, the estimate (86) follows by using Lemma 3.4 and summation by parts. Finally, after going through a similar process as in the proof for “(25)(26)\eqref{Eqn: analogue of main result in global function feild}\Leftrightarrow\eqref{Eqn: to write into partial sum}” (Sect. 2.2) by using the estimates (85) and (86), the desired result follows. ∎

As a result, Theorem 1.3 follows immediately from Lemma 3.1 and Theorem 3.7 by taking f(n)=1f(n)=1 for all n2n\in\mathbb{Z}_{\geq 2}. As for the proof of Corollary 1.4, it is similar to that of Corollary 1.7.

4. Applications of main theorems

This section is devoted to give some necessary background in order to make sense of our applications stated in Sect. 1.4. Moreover, we also include some results those are not stated in Sect. 1.4 due to the limit of space.

4.1. Number field and its integral ideal semigroup

For references of the followings, one can refer to [21, Chap. I]. Recall that a number field KK is defined to be a finite extension of the rational field \mathbb{Q} and the ring of algebraic integers 𝒪K\mathcal{O}_{K} of KK is a Dedekind domain. Thus every ideal 𝔞\mathfrak{a} in 𝒪K\mathcal{O}_{K} admits a uniquely factorization into prime ideals up to order, i.e., we have

𝔞=𝔭1a1𝔭2a2𝔭mam\mathfrak{a}=\mathfrak{p}_{1}^{a_{1}}\mathfrak{p}_{2}^{a_{2}}\cdots\mathfrak{p}_{m}^{a_{m}}

with each 𝔭i\mathfrak{p}_{i} a prime ideal and each ai>0a_{i}>0. And we let K\mathcal{I}_{K} be the semigroup of integral ideals. Since every Dedekind domain has Krull dimension one, each nonzero prime ideal 𝔭\mathfrak{p} is also maximal. Thus the quotient 𝒪K/𝔭\mathcal{O}_{K}/\mathfrak{p} is an integral zero-dimensional ring, i.e., a field. In fact, 𝒪K/𝔭\mathcal{O}_{K}/\mathfrak{p} is a finite field since it can be considered as a finite extension of the finite field /(𝔭)\mathbb{Z}/(\mathbb{Z}\cap\mathfrak{p}). Hence the norm 𝔭\colonequals[𝒪k:𝔭]=#(𝒪K/𝔭)\|\mathfrak{p}\|\colonequals[\mathcal{O}_{k}:\mathfrak{p}]=\#(\mathcal{O}_{K}/\mathfrak{p}) is well-defined. An effective form of Landau’s prime ideal theorem for algebraic number fields (see [19]) implies that

πK(x)\colonequals#{𝔭:𝔭x}=xlogx+OK(xexp(cKlogx))\pi_{K}(x)\colonequals\#\{\mathfrak{p}:\|\mathfrak{p}\|\leq x\}=\frac{x}{\log x}+O_{K}(x\,\mathrm{exp}(-c_{K}\sqrt{\log x}))

for some constant cK>0c_{K}>0 depending only on KK. In particular, the set 𝒫\mathcal{P} of all prime ideals is countable. Moreover, it is well known that

𝒩(x)\colonequals#{𝔞:𝔞x}=cKx+O(xη)\mathcal{N}(x)\colonequals\#\{\mathfrak{a}:\|\mathfrak{a}\|\leq x\}=c_{K}x+O(x^{\eta})

with η=11/[K:]\eta=1-1/[K:\mathbb{Q}] (see [20]). According to all of above facts, one can see that all the conditions of Axiom AA type arithmetical semigroup are satisfied. In conclusion, an immediate application of Theorem 1.3 deduces Corollary 1.8, which is stated again here for the convenience of readers.

Corollary 4.1.

Let S𝒫S\subseteq\mathcal{P} be a subset of prime ideals with natural density δ(S)\delta(S). For any arithmetic function a:Ka:\mathcal{I}_{K}\to\mathbb{C} supported on 𝔇(K,S)\mathfrak{D}(\mathcal{I}_{K},S) with a(𝒪K)=1a(\mathcal{O}_{K})=1, and

limx2𝔞x𝔞𝔇(K,S)|a(𝔞)|𝔞loglog𝔞<.\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|\mathfrak{a}\right\|\leq x\\ \mathfrak{a}\in\mathfrak{D}(\mathcal{I}_{K},S)\end{smallmatrix}}\frac{|a(\mathfrak{a})|}{\left\|\mathfrak{a}\right\|}\log\log\left\|\mathfrak{a}\right\|<\infty.

Then

limx2𝔞x𝔞𝔇(K,S)μa(𝔞)𝔞=δ(S).-\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|\mathfrak{a}\right\|\leq x\\ \mathfrak{a}\in\mathfrak{D}(\mathcal{I}_{K},S)\end{smallmatrix}}\frac{\mu*a(\mathfrak{a})}{\left\|\mathfrak{a}\right\|}=\delta(S).

In particular, we have

limx2𝔞x𝔞𝔇(K,S)μ(𝔞)𝔞=δ(S).-\lim_{x\to\infty}\sum_{\begin{smallmatrix}2\leq\left\|\mathfrak{a}\right\|\leq x\\ \mathfrak{a}\in\mathfrak{D}(\mathcal{I}_{K},S)\end{smallmatrix}}\frac{\mu(\mathfrak{a})}{\left\|\mathfrak{a}\right\|}=\delta(S).

4.2. Algebraic varieties and their 0-cycle semigroups

Fix pp to be a prime number and let q=prq=p^{r} be a prime power. Let 𝔽q\mathbb{F}_{q} be the corresponding finite field. Let XX be a dd-dimensional projective smooth variety defined over 𝔽q\mathbb{F}_{q}. A 0-cycle AA of XX over 𝔽q\mathbb{F}_{q} is a 0-dimensional subvariety of XX which is also defined over 𝔽q\mathbb{F}_{q}. By the Zariski topology assigned to XX, this means that the base change A𝔽¯qA_{\overline{\mathbb{F}}_{q}} of AA to the algebraic closure of 𝔽q\mathbb{F}_{q} is a finite set of geometric points which is stable under the action of Galois group Gal(𝔽¯q/𝔽q)\operatorname{Gal}(\overline{\mathbb{F}}_{q}/\mathbb{F}_{q}).

For any geometric point P~X(𝔽¯q)\widetilde{P}\in X(\overline{\mathbb{F}}_{q}), the induced effective prime 0-cycle (or prime cycle for simplicity) P=P(P~)P=P(\widetilde{P}) is defined to be the finite sum of all the distinct Galois conjugations of P~\widetilde{P} over the ground field 𝔽q\mathbb{F}_{q}, i.e., if P~\widetilde{P} is defined over 𝔽qn\mathbb{F}_{q^{n}} but not over any its proper subfield, then

P\colonequalsσGal(𝔽qn/𝔽q)P~σ.P\colonequals\sum\limits_{\sigma\in\operatorname{Gal}(\mathbb{F}_{q^{n}}/\mathbb{F}_{q})}{\widetilde{P}}^{\sigma}.

It follows immediately that P(P~)=P(P~σ)P(\widetilde{P})=P(\widetilde{P}^{\sigma}) for any σGal(𝔽¯q/𝔽q)\sigma\in\operatorname{Gal}(\overline{\mathbb{F}}_{q}/\mathbb{F}_{q}). In this subsection, we denote by 𝒫\mathcal{P} the set of prime cycles (of XX). And we define 𝒜X\mathcal{A}_{X} to be the semigroup of 𝔽q\mathbb{F}_{q}-effective 0-cycles of XX, i.e., it is the semigroup consisting of all the finite sum of prime formal sum of the form

A=i=1kaiPi=a1P1+a2P2++akPk,with Pi𝒫,ai0 and k1.A=\sum_{i=1}^{k}a_{i}P_{i}=a_{1}P_{1}+a_{2}P_{2}+\cdots+a_{k}P_{k},\quad\text{with }P_{i}\in\mathcal{P},a_{i}\in\mathbb{Z}_{\geq 0}\text{ and }k\in\mathbb{Z}_{\geq 1}.

In particular, 0 is contained in 𝒜X\mathcal{A}_{X}. For each prime cycle PP, we define its degree (P)\partial(P) to be the minimal positive integer mm such that PP splits over 𝔽qm\mathbb{F}_{q^{m}}, i.e., mm is the minimal integer such that there exists P~\widetilde{P} as above such that P=P(P~)P=P(\widetilde{P}). Note that this definition is independent of the choice of P~\widetilde{P}. Then we denote by P\colonequals(qd)(P)\|P\|\colonequals(q^{d})^{\partial(P)} and call it the norm of PP444Our definition of norm is different from the classical version P=q(P)\|P\|=q^{\partial(P)}, which is more familiar by algebraic geometer. The reason we take this change is such normalization will make it easier to apply our main results. If one would like to stick with the classical norm, then one needs to replace A\|A\| by Ad\|A\|^{d} in the statement of Corollary 1.10 to get the correct result.. For any effective 0-cycle AA, its degree is defined to be the summation of its prime cycles and its norm is the product of the corresponding norms.

Recall that the Hasse-Weil zeta function of XX is defined to be

(87) ZX(T)\colonequalsexp(n=1#X(𝔽qn)Tnn)=n=1(1Tn)πX#(n),Z_{X}(T)\colonequals\exp\left(\sum_{n=1}^{\infty}\#X(\mathbb{F}_{q^{n}})\frac{T^{n}}{n}\right)=\prod_{n=1}^{\infty}(1-T^{n})^{-\pi^{\#}_{X}(n)},

where π#(n)\colonequals{A𝒜X:(A)=n}\pi^{\#}(n)\colonequals\{A\in\mathcal{A}_{X}:\partial(A)=n\} is the total number of 0-cycles of degree nn. By the Weil conjecture [33], which has been settled by a series of works of Dwork [10], Grothendieck [12] and Deligne [7] we know that

(88) ZX(T)=1i2d1,i oddFi(T)(1T)(1qdT)2i2d2,i evenFi(T),Z_{X}(T)=\frac{\prod\limits_{1\leq i\leq 2d-1,i\text{ odd}}F_{i}(T)}{(1-T)(1-q^{d}T)\prod\limits_{2\leq i\leq 2d-2,i\text{ even}}F_{i}(T)},

where each Fi(T)[T]F_{i}(T)\in\mathbb{Z}[T] is a polynomial of degree BiB_{i} equaling the iith Betti number of XX. Moreover, for each ii the corresponding Fi(T)=j=1Bi(1αijT)F_{i}(T)=\prod_{j=1}^{B_{i}}(1-\alpha_{ij}T), where αij\alpha_{ij} are algebraic integers of absolute value qi/2q^{i/2}. In particular, ZX(T)Z_{X}(T) does not have a zero at T=q1T=-q^{-1}.

Taking logarithm of both sides of (88), using (87) to express #X(𝔽qn)\#X(\mathbb{F}_{q^{n}}) in terms of π#(m)\pi^{\#}(m) with m|nm|n and applying the Möbius inversion formula, one can deduce that

πX#(n)=(qd)nn+O(q(2d1)n2n).\pi^{\#}_{X}(n)=\frac{(q^{d})^{n}}{n}+O\left(\frac{q^{\frac{(2d-1)n}{2}}}{n}\right).

It follows that the set 𝒫\mathcal{P} of prime 0-cycles is countable. Then

G#(n)\colonequals#{A𝒜X:(A)=n}=cX(qd)n+O(q2d12n)G^{\#}(n)\colonequals\#\{A\in\mathcal{A}_{X}:\partial(A)=n\}=c_{X}(q^{d})^{n}+O\left(q^{\frac{2d-1}{2}n}\right)

follows immediately as argued in [14]. Here the constant cX>0c_{X}>0 only depends on XX. Thus, by all above facts, it follows that 𝒜X\mathcal{A}_{X} an arithmetical semigroup satisfying Axiom A#A^{\#}. In particular, by Theorem 1.6 we obtain Corollary 1.10 as desired which is copied here for the convenience of readers.

Corollary 4.2.

Suppose that S𝒫S\subseteq\mathcal{P} is a subset of prime 0-cycles with natural density δ(S)\delta(S). For any arithmetic function a:𝒜Xa:\mathcal{A}_{X}\to\mathbb{C} supported on 𝔇(𝒜X,S)\mathfrak{D}(\mathcal{A}_{X},S) with a(0)=1a(0)=1, and

limn1(A)nA𝔇(𝒜X,S)|a(A)|AloglogA<.\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(A)\leq n\\ A\in\mathfrak{D}(\mathcal{A}_{X},S)\end{smallmatrix}}\frac{|a(A)|}{\|A\|}\log\log\|A\|<\infty.

Then

(89) limn1(A)nA𝔇(𝒜X,S)μa(A)A=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(A)\leq n\\ A\in\mathfrak{D}(\mathcal{A}_{X},S)\end{smallmatrix}}\frac{\mu*a(A)}{\|A\|}=\delta(S).

In particular, we have

(90) limn1(A)nA𝔇(𝒜X,S)μ(A)A=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(A)\leq n\\ A\in\mathfrak{D}(\mathcal{A}_{X},S)\end{smallmatrix}}\frac{\mu(A)}{\|A\|}=\delta(S).

Moreover, we give two more concrete examples of Corollary 1.10 and Corollary 1.7 as follows.

Example 4.3.

Let X=𝔽q1X=\mathbb{P}^{1}_{\mathbb{F}_{q}} be the projective line over 𝔽q\mathbb{F}_{q}. For a polynomial FF in the affine coordinate ring 𝔽q[x]\mathbb{F}_{q}[x], we take the effective divisor of zeros of FF to be

(F)0\colonequalsordP(F)0ordP(F)P.(F)_{0}\colonequals\sum_{{\rm{ord}}_{P}(F)\geq 0}{\rm{ord}}_{P}(F)P.

Using the fact that 𝔽q(x)\mathbb{F}_{q}(x) has class number one, the map ϕ:F(F)0\phi:F\mapsto(F)_{0} gives a bijection of the following sets.

{monic polynomials F𝔽q[x]}\displaystyle\left\{\text{monic polynomials $F\in\mathbb{F}_{q}[x]$}\right\} {effective divisors D not supported by ()}\displaystyle\Longleftrightarrow\left\{\text{effective divisors $D$ not supported by $(\infty)$}\right\}
{irreducible monic polynomials FP𝔽q[x]}\displaystyle\left\{\text{irreducible monic polynomials $F_{P}\in\mathbb{F}_{q}[x]$}\right\} {prime divisors P()}\displaystyle\Longleftrightarrow\left\{\text{prime divisors $P\neq(\infty)$}\right\}

If a monic polynomial FF admit a unique irreducible factor Pmin(F)P_{\min}(F) with multiplicity one over 𝔽q\mathbb{F}_{q}, which is the minimal with respect to degree, then we say that FF is distinguishable. Under the above bijection ϕ\phi, if S\infty\notin S, then the set 𝔇(X,S)\mathfrak{D}(X,S) can be translated to

𝔇(q,S)\colonequals{F𝔽q[x] is monic :F is distinguishable and (Pmin(F))0S}.\mathfrak{D}(q,S)\colonequals\{F\in\mathbb{F}_{q}[x]\text{ is monic }\colon F\text{ is distinguishable and }(P_{\min}(F))_{0}\in S\}.

In this situation, if one takes aa to be the identity of the convolution ring of arithmetic functions over effective divisors of 𝔽q1\mathbb{P}_{\mathbb{F}_{q}}^{1} in Corollary 1.7 and uses the fact that δ(𝒫)=δ(𝒫{})\delta(\mathcal{P})=\delta(\mathcal{P}-\{\infty\}), then one immediately obtains the following corollary.

Corollary 4.4.

For two coprime monic polynomials f,g𝔽q[x]f,g\in\mathbb{F}_{q}[x], we have

(91) limnF𝔇(q,𝒫),1degFnpmin(F)f(modg)μ(F)φ(F)=1φ(g).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}F\in\mathfrak{D}(q,\mathcal{P}),1\leq\deg F\leq n\\ p_{\min}(F)\equiv f\,(\operatorname{mod}g)\end{smallmatrix}}\frac{\mu(F)}{\varphi(F)}=\frac{1}{\varphi(g)}.
Example 4.5.

We can also apply our results to study the distribution of subspaces. Let VV be an (+1)(\ell+1)-dimensional vector space over 𝔽q\mathbb{F}_{q}. Denote by G(k+1,+1)G(k+1,\ell+1) (or G(k+1,V)G(k+1,V)) the set of (k+1)(k+1)-dimensional linear subspaces (over 𝔽¯q\overline{\mathbb{F}}_{q}) of VV. Using the Plücker embedding G(k+1,+1)(k+1V){G}(k+1,\ell+1)\hookrightarrow\mathbb{P}(\bigwedge^{k+1}V) [13, Lecture 6], we can realize G(k+1,+1)G(k+1,\ell+1) as a (k+1)(k)(k+1)(\ell-k)-dimensional smooth projective variety defined over 𝔽q\mathbb{F}_{q}, i.e., the Grassmannian variety, which is denoted by 𝔾(k,𝔽q)\mathbb{G}(k,\mathbb{P}_{\mathbb{F}_{q}}^{\ell}). By the same embedding, one can see that [26, Proposition 1.7.2]

#𝔾(k,𝔽q)(𝔽qN)=|GL(+1,𝔽qN)|qN(k+1)(k)|GL(k+1,𝔽qN)||GL(k,𝔽qN)|.\#\mathbb{G}(k,\mathbb{P}_{\mathbb{F}_{q}}^{\ell})(\mathbb{F}_{q^{N}})=\frac{|{\rm{GL}}(\ell+1,\mathbb{F}_{q^{N}})|}{q^{N(k+1)(\ell-k)}|{\rm{GL}}(k+1,\mathbb{F}_{q^{N}})|\cdot|{\rm{GL}}(\ell-k,\mathbb{F}_{q^{N}})|}.

From this, one can deduce the zeta function of 𝔾(k+1,+1)\mathbb{G}(k+1,\ell+1), which turns out to be of the form

Z𝔾(k+1,+1)(T)=1(1T)(1qT)B2(1q2T)B4(1qdT),Z_{\mathbb{G}(k+1,\ell+1)}(T)=\frac{1}{(1-T)(1-qT)^{B_{2}}(1-q^{2}T)^{B_{4}}\cdots(1-q^{d}T)},

where d=(k+1)(k)d=(k+1)(\ell-k) and BiB_{i} is the Betti number of 𝔾(k,𝔽q)\mathbb{G}(k,\mathbb{P}_{\mathbb{F}_{q}}^{\ell}) for each ii.

Now let X=𝔽qX=\mathbb{P}_{\mathbb{F}_{q}}^{\ell}. Note that its affine cone 𝔸𝔽q+1\mathbb{A}_{\mathbb{F}_{q}}^{\ell+1} can be understood as a (+1)(\ell+1)-dimensional space over 𝔽q\mathbb{F}_{q}. For every (k+1)(k+1)-dimensional subspace H~\widetilde{H} of 𝔸𝔽¯q+1=𝔸𝔽q+1×Spec𝔽¯q\mathbb{A}_{\overline{\mathbb{F}}_{q}}^{\ell+1}=\mathbb{A}_{\mathbb{F}_{q}}^{\ell+1}\times{{\rm Spec}\,\overline{\mathbb{F}}_{q}}, denote by H\colonequalsH(H~)=σH~σ{H}\colonequals H(\widetilde{H})=\cup_{\sigma}\widetilde{H}^{\sigma}, where H~σ\widetilde{H}^{\sigma} runs over all the distinct Galois conjugates of H~\widetilde{H}. It is easy to see that each H{H} is a reduced variety defined over the ground field 𝔽q\mathbb{F}_{q}. Then define

𝒫k,,q\colonequals{H|H~G(k+1,𝔸𝔽q+1)}.\mathcal{P}_{k,\ell,q}\colonequals\{{H}|\widetilde{H}\in G(k+1,\mathbb{A}_{\mathbb{F}_{q}}^{\ell+1})\}.

Consider the semigroup

𝒜k,,q={a1H1a2H2asHs|H𝒫k,,q,ai0}\mathcal{A}_{k,\ell,q}=\{a_{1}{H}_{1}\cup a_{2}{H}_{2}\cup\cdots\cup a_{s}{H}_{s}|{H}\in\mathcal{P}_{k,\ell,q},a_{i}\in\mathbb{Z}_{\geq 0}\}

generated by 𝒫k,,q\mathcal{P}_{k,\ell,q} by taking finite union in the sense of schemes (i.e. count the multiplicities). In particular, if all ai=0a_{i}=0, then we let the union to the empty set \emptyset. One can check that the above setups give rise to a well-defined abelian free semigroup structure of 𝒜k,,q\mathcal{A}_{k,\ell,q} with the identity element \emptyset and the additional operation equaling to the union operation.

Recall that d=(k+1)(k)d=(k+1)(\ell-k), for each H𝒫k,,q{H}\in\mathcal{P}_{k,\ell,q}, we define degH\deg{H} to be the number of geometric irreducible components of H{H} and take the norm H\left\|H\right\| to be qddegHq^{d\deg H}.

In order to estimate the corresponding π#(n)\pi^{\#}(n) in this case, identify G(k+1,𝔸𝔽q+1)G(k+1,\mathbb{A}_{\mathbb{F}_{q}}^{\ell+1}) with 𝔾(k,𝔽q)\mathbb{G}(k,\mathbb{P}_{\mathbb{F}_{q}}^{\ell}) as in the first paragraph of this example. One can verify that under this identification, there is a natural bijection between 𝒜k,,q\mathcal{A}_{k,\ell,q} and 𝒜𝔾(k,𝔽q)\mathcal{A}_{\mathbb{G}(k,\mathbb{P}_{\mathbb{F}_{q}}^{\ell})} (i.e., the semigroup of 0-cycles of 𝔾(k,𝔽q)\mathbb{G}(k,\mathbb{P}_{\mathbb{F}_{q}}^{\ell})) and also a bijection between 𝒫k,,q\mathcal{P}_{k,\ell,q} and 𝒫𝔾(k,𝔽q)\mathcal{P}_{\mathbb{G}(k,\mathbb{P}_{\mathbb{F}_{q}}^{\ell})}. Moreover, it is not hard to see that this identification respects to the degree maps and norm maps. Hence we can deduce that

π#(n)=q(k+1)(k)nn+O(qη(k+1)(k)nn)\pi^{\#}(n)=\frac{q^{(k+1)(\ell-k)n}}{n}+O\left(\frac{q^{\eta{(k+1)(\ell-k)n}}}{n}\right)

and G#(n)=c𝒜k,,qq(k+1)(k)n+O(qη(k+1)(k)n)G^{\#}(n)=c_{\mathcal{A}_{k,\ell,q}}q^{(k+1)(\ell-k)n}+O(q^{\eta(k+1)(\ell-k)n}) with suitable constants c𝒜k,,q>0c_{\mathcal{A}_{k,\ell,q}}>0 and 0η<10\leq\eta<1. In particular, 𝒜k,,q\mathcal{A}_{k,\ell,q} is an Axiom A#A^{\#} type arithmetical semigroup by definition. Thus, it follows from Corollary 1.10 that we obtain

limn1(H)nH𝔇(𝒜k,,q,S)μ(H)H=δ(S),-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial(H)\leq n\\ H\in\mathfrak{D}(\mathcal{A}_{k,\ell,q},S)\end{smallmatrix}}\frac{\mu(H)}{\|H\|}=\delta(S),

as long as S𝒫k,,qS\subseteq\mathcal{P}_{k,\ell,q} has natural density δ(S)\delta(S) and the definition of 𝔇(𝒜k,,q,S)\mathfrak{D}(\mathcal{A}_{k,\ell,q},S) is an analogue to (11).

4.3. Finite graphs and the semigroups of closed paths

This subsection follows from [29], especially its chapter 2. The graphs in this section will always be finite, connected and undirected. The degree of a vertex is the number of the edges connecting this vertex. Moreover, we will assume that all the graphs in this section do not contain a degree-11 vertex. Given a graph GG, we will denote by VV the vertex set of GG and by EE the edge set of GG.

In order to define the prime element and the corresponding semigroup. We orient the edges in EE and label them as follows:

e1,e2,,em,em+1=e11,,e2m=em1,e_{1},e_{2},\cdots,e_{m},e_{m+1}=e_{1}^{-1},\cdots,e_{2m}=e_{m}^{-1},

where m=#(E)m=\#(E) is the number of unoriented edges and ej1=ej+me^{-1}_{j}=e_{j+m} is the edge eje_{j} with the opposite orientation. One oriented edge (i.e., the labeled edge) aja_{j} is called to follow another oriented edge aia_{i} if the start vertex of aja_{j} (with respect to the orientation of aja_{j}) is the end vertex of aia_{i}.

A path CC is a sequence {a1,a2,,as}\{a_{1},a_{2},\cdots,a_{s}\} of oriented edges of GG such that ai+1a_{i+1} follows aia_{i}. To simplify the notation, we will write such a path as C=a1asC=a_{1}\cdots a_{s}. A closed path is a path whose starting vertex and the terminal vertex coincide. In the following, we assume that all paths are closed and do not have a backtrack or tail, i.e., ai+1ai1a_{i+1}\neq a_{i}^{-1} for all i=1,,s1i=1,\ldots,s-1 and asa11a_{s}\neq a_{1}^{-1}. A (closed) path CC is called primitive or prime if CDkC\neq D^{k} for some positive integer k>1k>1, i.e., we cannot find another path D=a1asD=a_{1}\cdots a_{s} such that C=DD=(a1as)(a1as)C=D\cdots D=(a_{1}\cdots a_{s})\cdots(a_{1}\cdots a_{s}).

For a (closed) path C=a1asC=a_{1}\cdots a_{s}, the equivalence class [C][C] means the following

[C]={a1as,a2a3asa1,,asa1as1}.[C]=\{a_{1}\cdots a_{s},a_{2}a_{3}\cdots a_{s}a_{1},\cdots,a_{s}a_{1}\cdots a_{s-1}\}.

That is, [C][C] represents the set of paths which are only differed by the starting (hence also by the ending) vertex. We say that [C][C] is a prime if so is CC. And we denote by ν(C)=ν(a1as)=s\nu(C)=\nu(a_{1}\cdots a_{s})=s to be the length of CC. We also note that [C]=[a1as][C]=[a_{1}\cdots a_{s}] and [C1]=[asa1][C^{-1}]=[a_{s}\cdots a_{1}] are considered to be distinct classes.

One can verify that with above setups, every path CC can be decomposed into primitive paths and this decomposition induces a unique decomposition of [C][C] into prime classes up to order. Thus we will consider the semigroup 𝒜G\mathcal{A}_{G} of all the path classes regarding to GG and it is clear that

𝒜G={[C]=finite sumci[Pi]:[Pi]𝒫 and ci0 for all i}.\mathcal{A}_{G}=\left\{[C]=\sum_{\text{finite sum}}c_{i}[P_{i}]\colon[P_{i}]\in\mathcal{P}\text{ and }c_{i}\in\mathbb{Z}_{\geq 0}\text{ for all }i\right\}.

Here 𝒫\mathcal{P} the set of all primes [P][P] in GG. In particular, 𝒜G\mathcal{A}_{G} is abelian since the equivalence classes of path do not care the starting vertex.

Definition 4.6.

[29, Definition 2.2] The Ihara zeta function ζG\zeta_{G} for a finite connected graph without degree-11 vertices is a complex function is defined by

ζG(z)\colonequals[P]𝒫(1zν(P))1.\zeta_{G}(z)\colonequals\prod_{[P]\in\mathcal{P}}\left(1-z^{\nu(P)}\right)^{-1}.

The Ihara theorem generalized by Bass, Hashimoto, etc [29, Theorem 2.5] describes the properties of the Ihara zeta function. To state this result, we need to introduce some matrices associated to a graph. For a graph GG as above with =#(V)\ell=\#(V), the adjacency matrix AA of GG is an ×\ell\times\ell matrix with (i,j)(i,j)th entry

aij={number of undirected edges connecting vertex i to vertex j,if ij,2× number of loops at vertex i,if i=j.a_{ij}=\begin{cases}\text{number of undirected edges connecting vertex $i$ to vertex $j$},&\text{if }i\neq j,\\ \text{$2\times$ number of loops at vertex $i$},&\text{if }i=j.\end{cases}

Also, we associate GG with another diagonal matrix QQ whose jjth diagonal entry qjq_{j} such that qj+1q_{j}+1 is the degree of the jjth vertex of GG.

Theorem 4.7 ([29], Theorem 2.5).

Let the graph GG, the adjacency matrix AA and the diagonal matrix QQ be defined as above. And let r=#(E)#(V)+1r=\#(E)-\#(V)+1. Then

(92) ζG(z)=1(1z2)r1det(IAz+Qz2).\zeta_{G}(z)=\frac{1}{(1-z^{2})^{r-1}\det(I-Az+Qz^{2})}.

Since GG is connected, we know that r10r-1\geq 0 (in fact r1=0r-1=0 if and only if GG is a tree). It follows that ζG(z)\zeta_{G}(z) does not have a zero. Moreover, let RGR_{G} be the radius of convergence of ζG\zeta_{G}, i.e., the RGR_{G} is the minimum of the absolute value of all poles of ζG\zeta_{G}. Then we have the following result of Kotani and Sunada.

Theorem 4.8.

[29, Theorem 8.1] Let GG be as above, let α\alpha and β\beta be two integers such that α+1\alpha+1 is the maximal degree among all vertices and β+1\beta+1 be the minimal degree, respectively. Then every pole uu of ζG\zeta_{G} satisfies

RG|u|1and1αRG1β.R_{G}\leq|u|\leq 1\quad\text{and}\quad\frac{1}{\alpha}\leq R_{G}\leq\frac{1}{\beta}.

From this theorem, one can deduce (applying the same arguments as in the previous section) the corresponding prime number theorem of 𝒜G\mathcal{A}_{G}. Indeed, if we take ΔG\colonequalsgcd{ν(P):[P] prime of G}\Delta_{G}\colonequals\gcd\{\nu(P)\colon[P]\text{ prime of }G\}, then (see [29, §\S 2.7])

πG#(n)\colonequals#{primes [P]:ν(P)=ΔGn}=1nRGΔGn+O(1nRGΔGηn)\pi^{\#}_{G}(n)\colonequals\#\{\text{primes }[P]\colon\nu(P)=\Delta_{G}n\}=\frac{1}{nR_{G}^{\Delta_{G}n}}+O\left(\frac{1}{nR_{G}^{\Delta_{G}\eta n}}\right)

for some 0η<10\leq\eta<1. Thus the set 𝒫\mathcal{P} of primes [P][P] is countable, and similarly as previous sections, we have

G#(n)\colonequals#{[C]:ν(C)=ΔGn}=cGRGΔGn+O(1RGΔGηn),G^{\#}(n)\colonequals\#\{[C]:\nu(C)=\Delta_{G}n\}=\frac{c_{G}}{R_{G}^{\Delta_{G}n}}+O\left(\frac{1}{R_{G}^{\Delta_{G}\eta n}}\right),

where the positive constant cGc_{G} only depends on GG. By all above facts, we see that the semigroup 𝒜G\mathcal{A}_{G} satisfies all the conditions of Axiom A#A^{\#} in Definition 1.5, hence Corollary 1.14 follows as a consequence of Theorem 1.6. As what we did in other subsections, we restate the corollary here, in which we set the degree map ([C])=ν([C])\partial([C])=\nu([C]) for consistency and [C]=(1/RG)ν([C])\left\|[C]\right\|=(1/R_{G})^{\nu([C])}. Moreover, the definition of 𝔇(𝒜G,S)\mathfrak{D}(\mathcal{A}_{G},S) is an analogue to (11).

Corollary 4.9.

Assume that S𝒫S\subseteq\mathcal{P} is a subset of primitive path classes with natural density δ(S)\delta(S). For any arithmetic function a:𝒜Ga:\mathcal{A}_{G}\to\mathbb{C} supported on 𝔇(𝒜G,S)\mathfrak{D}(\mathcal{A}_{G},S) with a([0])=1a([0])=1, and

limn1([C])n[C]𝔇(𝒜G,S)|a([C])|[C]loglog[C]<.\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial([C])\leq n\\ [C]\in\mathfrak{D}(\mathcal{A}_{G},S)\end{smallmatrix}}\frac{|a([C])|}{\|[C]\|}\log\log\|[C]\|<\infty.

Then

limn1([C])n[C]𝔇(𝒜G,S)μa([C])[C]=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial([C])\leq n\\ [C]\in\mathfrak{D}(\mathcal{A}_{G},S)\end{smallmatrix}}\frac{\mu*a([C])}{\|[C]\|}=\delta(S).

In particular, we have

limn1([C])n[C]𝔇(𝒜G,S)μ([C])[C]=δ(S).-\lim_{n\to\infty}\sum_{\begin{smallmatrix}1\leq\partial([C])\leq n\\ [C]\in\mathfrak{D}(\mathcal{A}_{G},S)\end{smallmatrix}}\frac{\mu([C])}{\|[C]\|}=\delta(S).

Acknowledgments

We would like to thank Jeff Achter, Dongchun Han, Rachel Pries, Frank Thorne and Biao Wang for their helpful comments. We would also like to thank the referee for the detailed comments and suggestions.

References

  • [1] Krishnaswami Alladi. Duality between prime factors and an application to the prime number theorem for arithmetic progressions. J. Number Theory, 9(4):436–451, 1977.
  • [2] A. Axer. Beitrag zur Kenntnis der zahlentheoretischen Funktionen μ(n)\mu(n) und λ(n)\lambda(n). Prace Matematyczno-Fizyczne, 21(1):65–95, 1910.
  • [3] Weiyan Chen. Analytic number theory for 0-cycles. Math. Proc. Cambridge Philos. Soc., 166(1):123–146, 2019.
  • [4] Stephen D. Cohen. The function field abstract prime number theorem. Math. Proc. Cambridge Philos. Soc., 106(1):7–12, 1989.
  • [5] Madeline Locus Dawsey. A new formula for Chebotarev densities. Res. Number Theory, 3:Art. 27, 2017.
  • [6] Madeline Locus Dawsey, Matthew Just, and Robert Schneider. A ”supernormal” partition statistic. arXiv preprint arXiv:2107.14284, 2021.
  • [7] Pierre Deligne. La conjecture de Weil. I. Inst. Hautes Études Sci. Publ. Math., (43):273–307, 1974.
  • [8] Harold G. Diamond, Hugh L. Montgomery, and Ulrike M. A. Vorhauer. Beurling primes with large oscillation. Math. Ann., 334(1):1–36, 2006.
  • [9] Lian Duan, Biao Wang, and Shaoyun Yi. Analogues of Alladi’s formula over global function fields. Finite Fields Appl., 74:101874, 2021.
  • [10] Bernard Dwork. On the rationality of the zeta function of an algebraic variety. Amer. J. Math., 82:631–648, 1960.
  • [11] Yusuke Fujisawa and Makoto Minamide. On partial sums of the Möbius and Liouville functions for number fields. arXiv preprint arXiv:1212.4348, 2012.
  • [12] Alexander Grothendieck. Formule de Lefschetz et rationalité des fonctions LL. In Séminaire Bourbaki, Vol. 9, pages Exp. No. 279, 41–55. Soc. Math. France, Paris, 1995.
  • [13] Joe Harris. Algebraic geometry, volume 133 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1995. A first course, Corrected reprint of the 1992 original.
  • [14] Karl-Heinz Indlekofer, Eugenijus Manstavičius, and Richard Warlimont. On a certain class of infinite products with an application to arithmetical semigroups. Arch. Math. (Basel), 56(5):446–453, 1991.
  • [15] John Knopfmacher. Abstract analytic number theory. North-Holland Publishing Co., Amsterdam-Oxford; American Elsevier Publishing Co., Inc., New York, 1975. North-Holland Mathematical Library, Vol. 12.
  • [16] John Knopfmacher. Analytic arithmetic of algebraic function fields, volume 50 of Lecture Notes in Pure and Applied Mathematics. Marcel Dekker, Inc., New York, 1979.
  • [17] John Knopfmacher and Wen-Bin Zhang. Number theory arising from finite fields, volume 241 of Monographs and Textbooks in Pure and Applied Mathematics. Marcel Dekker, Inc., New York, 2001. Analytic and probabilistic theory.
  • [18] Michael Kural, Vaughan McDonald, and Ashwin Sah. Möbius formulas for densities of sets of prime ideals. Arch. Math. (Basel), 115(1):53–66, 2020.
  • [19] J. C. Lagarias and A. M. Odlyzko. Effective versions of the Chebotarev density theorem. In Algebraic number fields: LL-functions and Galois properties (Proc. Sympos., Univ. Durham, Durham, 1975), pages 409–464, 1977.
  • [20] M. Ram Murty and Jeanine Van Order. Counting integral ideals in a number field. Expo. Math., 25(1):53–66, 2007.
  • [21] Jürgen Neukirch. Algebraic number theory, volume 322 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999. Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder.
  • [22] Ken Ono, Robert Schneider, and Ian Wagner. Partition-theoretic formulas for arithmetic densities, II. The Hardy-Ramanujan Journal (Special Issue): https://arxiv.org/abs/2010.15107, 10010.
  • [23] Ken Ono, Robert Schneider, and Ian Wagner. Partition-theoretic formulas for arithmetic densities. In Analytic number theory, modular forms and qq-hypergeometric series, volume 221 of Springer Proc. Math. Stat., pages 611–624. Springer, Cham, 2017.
  • [24] Robert Schneider. Arithmetic of partitions and the qq-bracket operator. Proc. Amer. Math. Soc., 145(5):1953–1968, 2017.
  • [25] Robert Schneider. Partition-theoretic Abelian theorems. arXiv preprint arXiv:2011.08386, 2020.
  • [26] Richard P. Stanley. Enumerative combinatorics. Volume 1, volume 49 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, second edition, 2012.
  • [27] Naomi Sweeting and Katharine Woo. Formulas for Chebotarev densities of Galois extensions of number fields. Res. Number Theory, 5(1):Paper No. 4, 13, 2019.
  • [28] Terence Tao. A remark on partial sums involving the Möbius function. Bull. Aust. Math. Soc., 81(2):343–349, 2010.
  • [29] Audrey Terras. Zeta functions of graphs, volume 128 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2011. A stroll through the garden.
  • [30] Biao Wang. An analogue of a formula for Chebotarev densities. Int. J. Number Theory, 16(7):1557–1565, 2020.
  • [31] Biao Wang. The Ramanujan sum and Chebotarev densities. The Ramanujan Journal, Jul 2020.
  • [32] Biao Wang. Analogues of Alladi’s formula. Journal of Number Theory, 221:232–246, Apr 2021.
  • [33] André Weil. Numbers of solutions of equations in finite fields. Bull. Amer. Math. Soc., 55:497–508, 1949.