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

Weighted enumeration of number fields using Pseudo and Sudo maximal orders

Gaurav Digambar Patil Department of Mathematics, University of Toronto, Bahen Centre, 40 St. George Street, Room 6290, Toronto, Ontario, Canada, M5S 2E4 [email protected]
Abstract.

We establish a fundamental theorem of orders (FTO) which allows us to express all orders uniquely as an intersection of ‘irreducible orders’ along which the index and the conductor distributes multiplicatively. We define a subclass of Irreducible orders named Pseudo maximal orders. We then consider orders (called Sudo maximal orders) whose decomposition under FTO contains only Pseudo maximal orders. These rings can be seen as being “close” to being maximal (ring of integers) and thus there is a limited number of them with bounded index (by X). We give an upper bound for this quantity. We then show that all polynomials which can be sieved using only the Ekedahl sieve correspond to Sudo Maximal Orders. We use this understanding to get a weighted count for the number of number-fields with fixed degree and bounded discriminant using the concept of weakly divisible rings.

1. Introduction

We define the quantities S(X:n)S(X:n) and N(X:n)N(X:n) as follows:

Definition 1.
S(X:n):={K:[K:]=n,Gal(K¯/)Sn,𝔡𝔦𝔰𝔠(K)<X}S(X:n):=\{K:[K:\mathbb{Q}]=n,Gal(\bar{K}/\mathbb{Q})\simeq S_{n},\mathfrak{disc}(K)<X\}
N(X:n):=#S(X:n)N(X:n):=\#S(X:n)

Malle’s Conjecture (for SnS_{n}) in [8] tells us to expect

N(X:n)cnX.N(X:n)\simeq c_{n}X.

where cnc_{n} is a constant dependent on only n.n. This conjecture is known for small values of nn, namely n5.n\leq 5. However, some lower bounds on the quantity N(X:n)N(X:n) which are known are as follows.

  • In [7] (2002), Malle showed that

    N(X:n)X1n.N(X:n)\gg X^{\frac{1}{n}}.
  • In [5](2006) , Akshay Venkatesh and Jordan Ellenberg establish

    N(X:n)X12+1n2.N(X:n)\gg X^{\frac{1}{2}+\frac{1}{n^{2}}}.
  • In [2], Bhargava-Shankar-Wang show

    N(X:n)nX12+1nN(X:n)\gg_{n}X^{\frac{1}{2}+\frac{1}{n}}
  • In [3], the same authors show

    N(X:n)nX12+1n1.N(X:n)\gg_{n}X^{\frac{1}{2}+\frac{1}{n-1}}.
  • In our previous paper [10], we showed that

    Theorem 1 (Number-field count).
    N(X:n)nX12+1n1+(n3)rn(n1)(n2)N(X:n)\gg_{n}X^{\frac{1}{2}+\frac{1}{n-1}+\frac{(n-3)r_{n}}{(n-1)(n-2)}}

    where rn=ηnn24n+32ηn(n+2n2)r_{n}=\frac{\eta_{n}}{n^{2}-4n+3-2\eta_{n}(n+\frac{2}{n-2})} and where ηn\eta_{n} is 15n\frac{1}{5n} if nn is odd and is 188n6\frac{1}{88n^{6}} when nn is even.

In that paper([10]), we also show that the number of rings with rank nn over \mathbb{Z} with bounded discriminant is

nX12+1n4/3.\gg_{n}X^{\frac{1}{2}+\frac{1}{n-4/3}}.

Since, for every prime pp the ratio of non-maximal rings is O(1/p2)O(1/p^{2}), one might expect that N(X:n)N(X:n) is also of the same order. In this paper, we show that that is true up to weighted enumeration and an ϵ\epsilon in the order. Thus, showing that the liminf of N(X:n)N(X:n) is bounded below by X12+1n4/3.X^{\frac{1}{2}+\frac{1}{n-4/3}}.

Each of the above bounds uses different classifications of rings (see Introduction of our previous paper titled “Weakly Divisible Rings” [10].).

There are two main obstructions (there are three but we will only look at two) on to getting better and better lower bounds for N(X:n)N(X:n). One side of the problem requires more general parametrizations of rings(as discussed in [10]). The first three results given above use monic polynomials parametrizing monogenic rings. Heuristically, the third bound best possible using monogenic rings. This optimality is a result of solving a sieve theoretic problem. The fourth bound is based on binary rings and their parametrization via binary forms. Heuristically, this is the best possible result using just binary rings and a result of solving the second problem which is sieve theoretic in nature. In [10] we introduce a different parametrization, namely weakly divisible rings, to get the fifth result. Our other computations heuristically suggests that the lower bound can be improved to X12+1n4/3,\gg X^{\frac{1}{2}+\frac{1}{n-4/3}}, provided we can optimize the seive that sieves for maximal (weakly divisible) rings among all (weakly divisible) rings.

To get lower bounds, one has to restrict the given parametrization to certain local conditions that force the ring under consideration to be the whole ring of integers.

Ekedahl sieve allows us to sieve out points in a parameter space that are lifts of points in a codimension two (or more) Variety over 𝔽p.\mathbb{F}_{p}. We thus refer to such a condition as Ekedahlian. In general local conditions are a combination of modp\bmod p, modp2\bmod p^{2}, etc conditions. A specific type of local condition that is not easily handled by the Ekedahl sieve (in its base form) is often the biggest deterrent to our ability to sieve for maximal rings of integers. In this paper, we show that in some such cases, using the Ekedahl sieve only we can instead get a weighted lower bound of the appropriate order. We do this by carefully looking at what it means for a polynomial to be weakly divisible in terms of the ring associated to it. The concept of weakly divisible polynomials is defined in [3] and [2].

We define

Definition 2.
N(X:n):=𝕂S(X,n)1𝔡𝔦𝔰𝔠(𝕂).N^{*}(X:n):=\sum_{\mathbb{K}\in S(X,n)}\frac{1}{\sqrt{\mathfrak{disc}(\mathbb{K})}}.

Malle’s conjecture for N(X:n)N(X:n) is clearly equivalent (via discrete integration by parts computation) to

N(X:n)2cnX12.N^{*}(X:n)\simeq 2c_{n}X^{\frac{1}{2}}.

In our previous paper we showed that, the number of “weakly divisible” rings of rank nn over \mathbb{Z} with discriminant less than or equal to XX is

nX12+1n43.\gg_{n}X^{\frac{1}{2}+\frac{1}{n-\frac{4}{3}}}.

one would expect that the number of number fields one gets sieving these rings would also be

nX12+1n43.\gg_{n}X^{\frac{1}{2}+\frac{1}{n-\frac{4}{3}}}.

This is because the product of the local probabilities of a ring under consideration having a index co-prime to pp for any given prime pp is (in all the currently known parametrizations) (1O(p2)),\prod(1-O(p^{-2})), which is bounded below by an absolute constant dependent on only n.n.

In this paper, we will show (using only the Ekedahl Sieve in its base form)

Theorem 2.
N(X:n)X1n43ϵ.N^{*}(X:n)\gg X^{\frac{1}{n-\frac{4}{3}}-\epsilon}.

As a corollary, this implies

lim supXN(X:n)X12+1n43ϵ=.\limsup_{X\longrightarrow\infty}\frac{N(X:n)}{X^{\frac{1}{2}+\frac{1}{n-\frac{4}{3}}-\epsilon}}=\infty.

If we get some good upper-bound on the size of the space of parametrization used (weakly divisible polynomials) or some sort of equi-distribution, it would follow that

N(X:n)nX12+1n43ϵN(X:n)\gg_{n}X^{\frac{1}{2}+\frac{1}{n-\frac{4}{3}}-\epsilon}

improving the result of our previous paper.

In section 2 and section 3 we discuss a Fundamental Theorem of Orders, which describes Orders being described as an intersection of irreducible orders in a unique way. For this decomposition of an order, the conductor of the given order distributes multiplicatively along the conductors of the the irreducible orders in its decomposition. The index of an order in its ring of integers also distributes (multiplicatively) along the decomposition of the order into irreducible orders. The above theorem and understanding give us a slightly more nuanced way to understand conductor ideals and give a nice proof of Furtẅangler’s condition for a conductor ideal of order in a given number field.

We will apply the context of FTO to binary rings and as a consequence show a Dedekind-Kummer Type Theorem for Binary Rings and factorization of associated binary form. This was previously shown by [4], which we found in the references of [11], after writing this paper.

We discuss small prime splitting restrictions for Binary Rings, a restriction similar to the one the small prime splitting restrictions for Monogenic Rings.

We then show a structure theorem of all ‘large’ irreducible orders. We call these orders as Pseudo Maximal Orders. We will give a more general structure theorem for irreducible orders in a future paper.

We then define Sudo Maximal Orders as orders whose decomposition in the Fundamental Theorem of Orders consists of only‘large’ irreducible orders (i.e. Pseudo Maximal orders). We have a structure theorem for such orders. We bound the number of such orders in a given number field of bounded index. This quantity can be bound in terms of only the bound on the index and the degree of the number field. Notably, this bound is independent of the discriminant of the number field. This bound can be seen as an upper bound for all orders in a number field with squarefree conductors. This is evidence that indicates that the theorem shown in [6] about orders in number fields with degree less than or equal to five, is generalizable to higher degrees. In an upcoming paper, we will do this.

We use the Ekedahl sieve on the space of polynomials and get polynomials which we refer to as Ultra-Weakly divisible polynomials. As a consequence of the Dedekind-Kummer Theorem for Binary Rings and the Ultra-Weakly Divisible condition on Binary Rings associated to Ultra-Weakly Divisible polynomials it follows that the Binary Rings (and Weakly Divisible Rings) associated to Ultra-Weakly Divisible Polynomials are all Sudo Maximal Rings.

Combining with our bound on Sudo Maximal Orders in a number field we show the result in theorem 2.

2. Dedekind Domains in number-fields

Notation 3.

Let 𝕂\mathbb{K} denote a number-field, 𝒪𝕂\mathcal{O}_{\mathbb{K}} its ring of integers. Let M(𝕂)M(\mathbb{K}) denote the set of non-Archimedean places of 𝕂\mathbb{K} i.e. the set of non-zero prime ideals in 𝒪𝕂\mathcal{O}_{\mathbb{K}} (or maximal ideals). Let \mathbb{P}_{\mathbb{Z}} denote the set of prime numbers in integers.

For R𝕂R\subseteq\mathbb{K} be a ring (not necessarily of finite rank over \mathbb{Z}) such that Frac(R)=𝕂,Frac(R)=\mathbb{K}, let R¯\bar{R} denote its integral closure.

Lemma 1.

p:|R/pR|p[𝕂:].\forall p\in\mathbb{P}_{\mathbb{Z}}:|\nicefrac{{R}}{{pR}}|\leq p^{[\mathbb{K}:\mathbb{Q}]}.

Proof.

We see that R/pR\nicefrac{{R}}{{pR}} as a /p=𝔽p\nicefrac{{\mathbb{Z}}}{{p\mathbb{Z}}}=\mathbb{F}_{p}-module.
For xRx\in R let xx^{*} denote its image under the canonical map from RR to R/pR.\nicefrac{{R}}{{pR}}. We follow our previous observation with another,

if x1,x2,xkx_{1}^{*},x_{2}^{*},\cdots x_{k}^{*} are linearly independent over 𝔽p,\mathbb{F}_{p}, then x1,x2,,xnx_{1},x_{2},\cdots,x_{n} are linearly independent over \mathbb{Z}(or equivalently, over \mathbb{Q}).

This is is easy to see since if xix_{i}’s satisfy a \mathbb{Z} linear relationship we may re write the relationship with coefficients such that the gcd of all the coefficients is 11 i.e. we may find aia_{i} such that i=1kaixi=0\sum_{i=1}^{k}a_{i}x_{i}=0 and (a1,a2,ak)=1(a_{1},a_{2},\cdots a_{k})=1 which immediately allows us to see i=1kaixi=0\sum_{i=1}^{k}a_{i}^{*}x_{i}^{*}=0 where not all ai=0a_{i}^{*}=0 as p(a1,a2,ak)=1.p\nmid(a_{1},a_{2},\cdots a_{k})=1.

Lemma 2.

p:|R/pkR|=|R/pR|k.\forall p\in\mathbb{P}_{\mathbb{Z}}:|\nicefrac{{R}}{{p^{k}R}}|=|\nicefrac{{R}}{{pR}}|^{k}.

Proof.

Multiplication by pmp^{m} from RpmRR\longrightarrow p^{m}R and pRpm+1RpR\longrightarrow p^{m+1}R are bijections and thus the canonical induced map, R/pRpmR/pm+1R\nicefrac{{R}}{{pR}}\longrightarrow\nicefrac{{p^{m}R}}{{p^{m+1}R}} is also a bijection, giving us |pmR/pm+1R|=|RpR|.|\nicefrac{{p^{m}R}}{{p^{m+1}R}}|=|\frac{R}{pR}|. The lemma follows. ∎

It follows that R/nR\nicefrac{{R}}{{nR}} is a finite module for all n.n.

If xRx\in R then one can see that the numerator of 𝔑𝔪(x)\mathfrak{Nm}(x) (say tt) can be expressed as an integral polynomial of xx and thus R/xRR/tR\nicefrac{{R}}{{xR}}\subseteq\nicefrac{{R}}{{tR}} which is a finite set. Implying that every ideal in RR is finitely generated and thus no-etherian. We also note that for every non zero prime ideal in R,R, R/ρ\nicefrac{{R}}{{\rho}} is a finite integral domain and hence a field i.e. ρ\rho is a maximal ideal. Thus, R¯\bar{R} is Noetherian, integrally closed domain with Krull dimension equal to one. It follows that R¯\bar{R} is a Dedekind Domain.

We naturally have 𝒪𝕂R¯\mathcal{O}_{\mathbb{K}}\subseteq\bar{R} for every such ring. (Since, integral closure of \mathbb{Z} in 𝕂\mathbb{K} (or 𝒪𝕂\mathcal{O}_{\mathbb{K}}) will sit in the integral closure of RR in 𝕂.\mathbb{K}.)

Every non-zero prime (ρ\rho) in R¯\bar{R} thus naturally corresponds to a prime (ρ𝒪𝕂\rho\cap\mathcal{O}_{\mathbb{K}}) in 𝒪𝕂\mathcal{O}_{\mathbb{K}}.
Furthermore, since R¯\bar{R} is a Dedekind domain R¯ρ\bar{R}_{\rho} (R¯\bar{R} localized at ρ\rho) is a DVR and hence, a prime ρ\rho in R¯\bar{R} will naturally induce a non-Archimedean metric on 𝕂\mathbb{K}. Ostrowski’s theorem tells us that M(𝕂)M(\mathbb{K}) simultaneously denotes the set of all non zero primes in 𝒪𝕂\mathcal{O}_{\mathbb{K}} and all possible non-Archimedean valuations on 𝕂\mathbb{K}. If v=ρ(𝒪𝕂)M(𝕂)v=\rho\cap(\mathcal{O}_{\mathbb{K}})\in M(\mathbb{K}), it follows that R¯ρ=(𝒪𝕂)v.\bar{R}_{\rho}=(\mathcal{O}_{\mathbb{K}})_{v}. This allows us to see non-zero prime ideals in R¯\bar{R} as a subset of M(𝕂).M(\mathbb{K}). On the other hand, given any nonempty subset (SS) of M(𝕂)M(\mathbb{K}) one can obtain a Dedekind domain from this set by defining

Notation 4.
RS:=vS(𝒪𝕂)v.R_{S}:=\bigcap_{v\in S}(\mathcal{O}_{\mathbb{K}})_{v}.

We thus get the following correspondence,

Theorem 3.
{Dedekind Domains in 𝕂}{ Non-Empty subsets of M(𝕂)}\bigg{\{}\textit{Dedekind Domains in }\mathbb{K}\bigg{\}}\leftrightarrow\bigg{\{}\textit{ Non-Empty subsets of }M(\mathbb{K})\bigg{\}}
Notation 5.

When RR is a Dedekind Domain in 𝕂\mathbb{K}, we use the notation

M(R):={vM(𝕂):R(𝒪𝕂)v}.M(R):=\{v\in M(\mathbb{K}):R\subseteq(\mathcal{O}_{\mathbb{K}})_{v}\}.

We can write the correspondence as

RM(R)\displaystyle R\leftrightarrow M(R)
RSS.\displaystyle R_{S}\leftrightarrow S.
Remark 4.

We can let gg denote the class number of 𝕂\mathbb{K} and observe that vgv^{g} can be seen as a principal ideal (tv).(t_{v}). For SM(𝕂)S\subseteq M(\mathbb{K}) we may define

A(S):={vStvnv:nv0}A(S):=\{\prod_{v\in S}^{*}t_{v}^{n_{v}}:n_{v}\in\mathbb{Z}_{\geq 0}\}

where \prod^{*} denotes finitely supported product. Then, the above correspondence from set to domain can be written as

SA(S)1𝒪𝕂.S\longrightarrow A(S)^{-1}\cdot\mathcal{O}_{\mathbb{K}}.

In particular, we notice that for any ring RR,

R¯=R𝒪𝕂.\bar{R}=R\mathcal{O}_{\mathbb{K}}.

3. Orders and Separation Lemmata

Notation 6.

For R𝕂R\subseteq\mathbb{K} a ring with Frac(R)=𝕂Frac(R)=\mathbb{K}, we define M(R)M(R) to be the set of non-zero prime ideals (same as maximal ideals in this context) of R.R.

When RR is integrally closed (and thus a Dedekind Domain) this matches up with the understanding that M(R)M(R) is the set of valuations corresponding to non-zero prime ideals of RR from above.

We note that for every ring RR satisfying Frac(R)=𝕂,Frac(R)=\mathbb{K}, M(R¯)M(\bar{R}) can be seen as a cover of M(R).M(R).

To see this, we note that for ρ,ρM(R)\rho,\rho^{\prime}\in M(R) we clearly have ρ+ρ=R\rho+\rho^{\prime}=R since they are distinct maximal ideals. Thus, valuations over ρ\rho and ρ\rho^{\prime} have to be disjoint.

Notation 7.

For ρR,\rho\in R, we define

ρ¯:={vM(R¯):Rρ(𝒪𝕂)v}.\bar{\rho}:=\{v\in M(\bar{R}):R_{\rho}\subseteq(\mathcal{O}_{\mathbb{K}})_{v}\}.

In other words, {ρ¯:ρM(R)}\{\bar{\rho}:\rho\in M(R)\} is a partition of M(R¯).M(\bar{R}).

We also note that

R=ρRRρ.R=\bigcap_{\rho\in R}R_{\rho}.

We also note that there will only exist finitely many ρM(R)\rho\in M(R) such that Rρ(𝒪𝕂)v.R_{\rho}\subsetneq(\mathcal{O}_{\mathbb{K}})_{v}. See, remark 5.

Notation 8.

We set

S(R):={ρM(R):Rρ(𝒪𝕂)v}.S(R):=\{\rho\in M(R):R_{\rho}\subsetneq(\mathcal{O}_{\mathbb{K}})_{v}\}.

For all ρS(R),\rho\notin S(R), we have |ρ¯|=1.|\bar{\rho}|=1.

Definition 9.

We say an ring RR with Frac(R)=𝕂Frac(R)=\mathbb{K} is irreducible if |S(R)|=1.|S(R)|=1.

In other words, RR is irreducible if and only if RR¯R\neq\bar{R} and !ρ\exists!\rho, a maximal ideal in RR such that ρρ\forall\rho^{\prime}\neq\rho in M(R),M(R), we have

Rρ=(𝒪𝕂)vR_{\rho^{\prime}}=(\mathcal{O}_{\mathbb{K}})_{v}

for some vM(R¯).v\in M(\bar{R}).

We call RR an irreducible order, when RR is an order that is irreducible.

We thus have every ring RR with Frac(R)=𝕂,Frac(R)=\mathbb{K}, we have

R=ρS(R)(RρR¯).R=\bigcap_{\rho\in S(R)}(R_{\rho}\cap\bar{R}). (1)
Notation 10.

For an order 𝒪,\mathcal{O}, we let c𝒪c_{\mathcal{O}} denote the conductor of the order 𝒪.\mathcal{O}. In other words,

c𝒪:=(𝒪:𝒪𝕂)𝕂={x𝕂:x𝒪𝕂𝒪}c_{\mathcal{O}}:=(\mathcal{O}:\mathcal{O}_{\mathbb{K}})_{\mathbb{K}}=\{x\in\mathbb{K}:x\mathcal{O}_{\mathbb{K}}\subseteq\mathcal{O}\}
Remark 5.

We recall some properties of the conductor ideal of an order.

  • c𝒪c_{\mathcal{O}} is simultaneously an ideal of 𝒪\mathcal{O} and 𝒪𝕂.\mathcal{O}_{\mathbb{K}}.

  • c𝒪c_{\mathcal{O}} is the largest ideal of 𝒪𝕂\mathcal{O}_{\mathbb{K}} sitting inside 𝒪\mathcal{O} as a subset.

  • 𝒪ρ\mathcal{O}_{\rho} is not a DVR (i.e. is not equal to (𝒪𝕂)v(\mathcal{O}_{\mathbb{K}})_{v} for some valuation vM(𝕂)v\in M(\mathbb{K})) if and only if c𝒪ρ.c_{\mathcal{O}}\subseteq\rho.

Thus, the radical of c𝒪c_{\mathcal{O}} in 𝕂\mathbb{K} will be the exactly equal to the product of valuations in ρS(𝒪)ρ¯.\sqcup_{\rho\in S(\mathcal{O})}\bar{\rho}.

Lemma 3 (Separation Lemma 1).

Let 𝒪\mathcal{O} is an order in 𝒪𝕂,\mathcal{O}_{\mathbb{K}}, and

𝒪=ρS(𝒪)(𝒪ρ𝒪𝕂)\mathcal{O}=\bigcap_{\rho\in S(\mathcal{O})}(\mathcal{O}_{\rho}\cap\mathcal{O}_{\mathbb{K}})

from eq. 1.

Let cρc_{\rho} denote the conductor of 𝒪ρ𝒪𝕂.\mathcal{O}_{\rho}\cap\mathcal{O}_{\mathbb{K}}. We have

  • [𝒪𝕂:𝒪]=ρS(𝒪)[𝒪𝕂:(𝒪ρ𝒪𝕂)].[\mathcal{O}_{\mathbb{K}}:\mathcal{O}]=\prod_{\rho\in S(\mathcal{O})}[\mathcal{O}_{\mathbb{K}}:(\mathcal{O}_{\rho}\cap\mathcal{O}_{\mathbb{K}})].
  • For ρ,ρS(𝒪)\rho,\rho^{\prime}\in S(\mathcal{O})

    cρ+cρ=𝒪𝕂c_{\rho}+c_{\rho^{\prime}}=\mathcal{O}_{\mathbb{K}}

    and

    c𝒪=ρS(𝒪)cρc_{\mathcal{O}}=\prod_{\rho\in S(\mathcal{O})}c_{\rho}
Proof.

Since, for distinct primes ρ,ρM(𝒪)\rho,\rho^{\prime}\in M(\mathcal{O}) we have ρ+ρ=𝒪.\rho+\rho^{\prime}=\mathcal{O}. This implies that, for any primary ideals ,\partial,\partial^{\prime} satisfying =ρ\sqrt{\partial}=\rho and =ρ,\sqrt{\partial^{\prime}}=\rho^{\prime}, we have +=𝒪.\partial+\partial^{\prime}=\mathcal{O}. In other words,

𝒪/𝒪/×𝒪/.\nicefrac{{\mathcal{O}}}{{\partial\partial^{\prime}}}\simeq\nicefrac{{\mathcal{O}}}{{\partial}}\times\nicefrac{{\mathcal{O}}}{{\partial^{\prime}}}.

It follows that

𝒪p=ρS(𝒪)(𝒪ρ𝒪𝕂)p=ρS(𝒪)𝒪ρp\mathcal{O}\otimes\mathbb{Z}_{p}=\prod_{\rho\in S(\mathcal{O})}(\mathcal{O}_{\rho}\cap\mathcal{O}_{\mathbb{K}})\otimes\mathbb{Z}_{p}=\prod_{\rho\in S(\mathcal{O})}\mathcal{O}_{\rho}\otimes\mathbb{Z}_{p}

where 𝒪ρpvρ¯(𝒪𝕂)v.\mathcal{O}_{\rho}\otimes\mathbb{Z}_{p}\subseteq\prod_{v\in\bar{\rho}}(\mathcal{O}_{\mathbb{K}})_{v}.

We can look at all of these as p\mathbb{Z}_{p} modules and compare the p\mathbb{Z}_{p} indexes. We get

vp([𝒪𝕂:𝒪])=ρS(𝒪)vp([𝒪𝕂:(𝒪ρ𝒪𝕂)]).v_{p}([\mathcal{O}_{\mathbb{K}}:\mathcal{O}])=\sum_{\rho\in S(\mathcal{O})}v_{p}([\mathcal{O}_{\mathbb{K}}:(\mathcal{O}_{\rho}\cap\mathcal{O}_{\mathbb{K}})]).

The result follows.

The other part follows naturally from Remark remark 5 as

cρ=vρ¯v\sqrt{c_{\rho}}=\prod_{v\in\bar{\rho}}v

and the fact that for any distinct primes ρ,ρM(R),\rho,\rho^{\prime}\in M(R),

ρ¯ρ¯=ϕ.\bar{\rho}\cap\bar{\rho}^{\prime}=\phi.

Lemma 4 (Separation Lemma 2).

If 𝒪=𝒪1𝒪2\mathcal{O}=\mathcal{O}_{1}\cap\mathcal{O}_{2} such that c𝒪1+c𝒪2=𝒪𝕂,c_{\mathcal{O}_{1}}+c_{\mathcal{O}_{2}}=\mathcal{O}_{\mathbb{K}}, then S(𝒪1)S(\mathcal{O}_{1}) and S(𝒪2)S(\mathcal{O}_{2}) can be naturally identified with disjoint subsets of S(𝒪)S(\mathcal{O}) such that

  • S(𝒪1)S(𝒪2)=S(𝒪)S(\mathcal{O}_{1})\sqcup S(\mathcal{O}_{2})=S(\mathcal{O})

  • ρS(𝒪1)S(𝒪)𝒪ρ=(𝒪1)ρ\rho\in S(\mathcal{O}_{1})\subseteq S(\mathcal{O})\iff\mathcal{O}_{\rho}=(\mathcal{O}_{1})_{\rho}

  • ρS(𝒪2)S(𝒪)𝒪ρ=(𝒪2)ρ.\rho\in S(\mathcal{O}_{2})\subseteq S(\mathcal{O})\iff\mathcal{O}_{\rho}=(\mathcal{O}_{2})_{\rho}.

Proof.

Since, c𝒪1+c𝒪2=𝒪𝕂c_{\mathcal{O}_{1}}+c_{\mathcal{O}_{2}}=\mathcal{O}_{\mathbb{K}} let r1+r2=1r_{1}+r_{2}=1 with r1c𝒪1r_{1}\in c_{\mathcal{O}_{1}} and r2c𝒪2.r_{2}\in c_{\mathcal{O}_{2}}.

Fix a prime ideal ρS(𝒪1)\rho\in S(\mathcal{O}_{1}) (that is, c𝒪1ρ)c_{\mathcal{O}_{1}}\subseteq\rho). Let τ=ρ𝒪.\tau=\rho\cap\mathcal{O}.

Clearly, τ\tau is a prime ideal in 𝒪\mathcal{O} and 𝒪τ(𝒪1)ρ.\mathcal{O}_{\tau}\subseteq(\mathcal{O}_{1})_{\rho}.

If s(𝒪1)ρs\in(\mathcal{O}_{1})_{\rho} then s=q/ts=q/t for some q𝒪1q\in\mathcal{O}_{1} and t𝒪1\ρt\in\mathcal{O}_{1}\backslash\rho. We simply note that since tr2c𝒪2𝒪2tr_{2}\in c_{\mathcal{O}_{2}}\subseteq\mathcal{O}_{2} and tr2=ttr1𝒪1tr_{2}=t-tr_{1}\in\mathcal{O}_{1} we get tr2𝒪.tr_{2}\in\mathcal{O}. We further note that tr2τtr_{2}\notin\tau as tr2τρtr_{2}\in\tau\subseteq\rho would imply r2ρr_{2}\in\rho and since r1c𝒪1r_{1}\in c_{\mathcal{O}_{1}} we get 1=r1+r2ρ.1=r_{1}+r_{2}\in\rho.) Similarly, qr2c𝒪2𝒪2qr_{2}\in c_{\mathcal{O}_{2}}\subseteq\mathcal{O}_{2} and qr2=qqr1𝒪1qr_{2}=q-qr_{1}\in\mathcal{O}_{1} we get qr2𝒪.qr_{2}\in\mathcal{O}.

We get s=q/t=(qr2)/(tr2)𝒪τs=q/t=(qr_{2})/(tr_{2})\in\mathcal{O}_{\tau} implying

ρS(𝒪1)(𝒪1)ρ=𝒪ρ𝒪\rho\in S(\mathcal{O}_{1})\Rightarrow(\mathcal{O}_{1})_{\rho}=\mathcal{O}_{\rho\cap\mathcal{O}}

If ρ,ρS(𝒪1)\rho,\rho^{\prime}\in S(\mathcal{O}_{1}) such that ρ𝒪=ρ𝒪,\rho\cap\mathcal{O}=\rho^{\prime}\cap\mathcal{O}, then (𝒪1)ρ=(𝒪2)ρ(\mathcal{O}_{1})_{\rho}=(\mathcal{O}_{2})_{\rho^{\prime}} and thus ρ=ρ.\rho=\rho^{\prime}.

Thus, the canonical map S(𝒪1)S(𝒪)S(\mathcal{O}_{1})\longrightarrow S(\mathcal{O}) given by ρρ𝒪\rho\longrightarrow\rho\cap\mathcal{O} is injective. Similarly we can identify S(𝒪2)S(\mathcal{O}_{2}) with a subset of S(𝒪).S(\mathcal{O}). To show that S(𝒪1)S(\mathcal{O}_{1}) and S(𝒪2)S(\mathcal{O}_{2}) are disjoint, we only need to note that if some ρS(𝒪)\rho\in S(\mathcal{O}) lies in both S(𝒪1)S(\mathcal{O}_{1}) and S(𝒪2),S(\mathcal{O}_{2}), we would immediately have r1c𝒪1ρr_{1}\in c_{\mathcal{O}_{1}}\subseteq\rho and r2c𝒪2ρ.r_{2}\in c_{\mathcal{O}_{2}}\subseteq\rho. Forcing 1=r1+r2ρ.1=r_{1}+r_{2}\in\rho. Contradicting ρ\rho is a prime ideal. ∎

Remark 6.

Note that the above lemma tells us that irreducible orders mimic Euclid’s property defining prime numbers, that is

If 𝒪\mathcal{O} is an irreducible order, then 𝒪A,𝒪B\mathcal{O}_{A},\mathcal{O}_{B} are orders with c𝒪A+c𝒪B=𝒪𝕂c_{\mathcal{O}_{A}}+c_{\mathcal{O}_{B}}=\mathcal{O}_{\mathbb{K}}, then

𝒪A𝒪B𝒪𝒪A𝒪 or 𝒪B𝒪.\mathcal{O}_{A}\cap\mathcal{O}_{B}\subseteq\mathcal{O}\Rightarrow\mathcal{O}_{A}\subseteq\mathcal{O}\textbf{ or }\mathcal{O}_{B}\subseteq\mathcal{O}.
Theorem 7 (Fundamental Theorem of Orders).

Every order 𝒪\mathcal{O} can be written as an intersection of irreducible orders in a unique way such that the conductors of the irreducible orders are pairwise co-prime. Furthermore, we get that the index (and conductor) of 𝒪\mathcal{O} in 𝒪𝕂\mathcal{O}_{\mathbb{K}} will be the product of the indices (and conductors) of the irreducible orders in 𝒪𝕂\mathcal{O}_{\mathbb{K}} in the given decomposition.

Proof.

Clearly follows from lemma 3 and lemma 4. ∎

Notation 11.

For any valuation vM(𝕂)v\in M(\mathbb{K}), we denote the ramification index and the inertial degree of vv by eve_{v} and fv,f_{v}, respectively. That is, if v=(p)v\cap\mathbb{Z}=(p) for some prime in ,\mathbb{Z}, then eve_{v} and fvf_{v} are defined by

p𝒪𝕂=pvvevp\mathcal{O}_{\mathbb{K}}=\prod_{p\in v}v^{e_{v}}

and

𝔑𝔪𝒪𝕂(v)=pfv.\mathfrak{Nm}_{\mathcal{O}_{\mathbb{K}}}(v)=p^{f_{v}}.
Definition 12.

For any order 𝒪\mathcal{O} and any non-zero prime ideal of ρM(𝒪),\rho\in M(\mathcal{O}), we define

ef(ρ):=vρ¯evfv.ef(\rho):=\sum_{v\in\bar{\rho}}e_{v}\cdot f_{v}.
Definition 13.

If 𝒪\mathcal{O} is an irreducible order with S(𝒪)={ρ},S(\mathcal{O})=\{\rho\}, we define

ef(𝒪):=ef(ρ).ef(\mathcal{O}):=ef(\rho).

4. Dedekind-Kummer type Theorems and the Fundamental theorem of Orders.

We recall the traditional versions of Dedekind-Kummer and Dedekind Criterion theorems surrounding polynomials.

4.1. Dedekind-Kummer Theorem

Theorem 8 (Dedekind-Kummer Theorem).

Suppose α¯\alpha\in\overline{\mathbb{Z}} satisfies [α]=𝕂,\mathbb{Q}[\alpha]=\mathbb{K}, and suppose that f(x)f(x) is the minimal polynomial of α\alpha (will be monic) over 𝕂\mathbb{K}.

If p[𝒪𝕂:[α]]p\nmid[\mathcal{O}_{\mathbb{K}}:\mathbb{Z}[\alpha]] and

f(x)i(fi)eimodp,f(x)\equiv\prod_{i}(f_{i})^{e_{i}}\bmod p,

then

p𝒪𝕂=i(βi)eip\mathcal{O}_{\mathbb{K}}=\prod_{i}(\beta_{i})^{e_{i}}

where βi\beta_{i} are prime ideals in 𝒪𝕂\mathcal{O}_{\mathbb{K}} sitting over p,p, with the understanding that

βi=p𝒪𝕂+fi(α)𝒪𝕂.\beta_{i}=p\mathcal{O}_{\mathbb{K}}+f_{i}(\alpha)\mathcal{O}_{\mathbb{K}}.
Notation 14.

Let g¯\overline{g} denote the image of g[x]g\in\mathbb{Z}[x] in 𝔽p[x].\mathbb{F}_{p}[x].

Dedekind’s Criterion (see Lemma 3.1 in [1]) for index adds

Theorem 9 (Dedekind’s Criterion).

We have

p|[𝒪𝕂:[α]]i:1ik,fi¯|fi(fi)eip¯.p|[\mathcal{O}_{\mathbb{K}}:\mathbb{Z}[\alpha]]\iff\exists i:1\leq i\leq k,\>\overline{f_{i}}|\overline{\frac{f-\prod_{i}(f_{i})^{e_{i}}}{p}}.

Equivalently, we can say

p|[𝒪𝕂:[α]]g[x]:f(p2,pg,g2)[x]p|[\mathcal{O}_{\mathbb{K}}:\mathbb{Z}[\alpha]]\iff\exists g\in\mathbb{Z}[x]:f\in(p^{2},pg,g^{2})\subseteq\mathbb{Z}[x]

We will unpack the standard proof of Dedekind-Kummer theorem with the Dedekind’s Criterion by talking about the ring [α]\mathbb{Z}[\alpha]rather than 𝒪𝕂\mathcal{O}_{\mathbb{K}} as follows.

Theorem 10.

If 𝒪=[α]\mathcal{O}=\mathbb{Z}[\alpha] with α¯\alpha\in\overline{\mathbb{Z}} and if f(x)f(x) the minimal polynomial of ff shows the following decomposition modulo p,p,

f(x)i(fi)eimodp,f(x)\equiv\prod_{i}(f_{i})^{e_{i}}\bmod p,

then, we get

(\p)1𝒪=i𝒪ρi(\mathbb{Z}\backslash p\mathbb{Z})^{-1}\mathcal{O}=\bigcap_{i}\mathcal{O}_{\rho_{i}}

and

p𝒪=iβip\mathcal{O}=\prod_{i}\beta_{i}

where βi\beta_{i} are primary ideals in 𝒪\mathcal{O} sitting over pp such that

βi=ρi\sqrt{\beta_{i}}=\rho_{i}

with the understanding that

ρi=p𝒪+fi(α)𝒪.\rho_{i}=p\mathcal{O}+f_{i}(\alpha)\mathcal{O}.

are all distinct prime ideals and all prime ideals in 𝒪\mathcal{O} sitting over pp and

ief(ρi)=[𝕂:].\sum_{i}ef(\rho_{i})=[\mathbb{K}:\mathbb{Q}].
Remark 11.

The theorem 10 says that, “The factorization of a monic polynomial modulo pp informs us about the local ring decomposition of [α]\mathbb{Z}[\alpha] over pp” and the theorem 9 is about when a particular factor (corresponding to a particular prime ideal) does or does not correspond to a DVR.

4.2. Dedekind-Kummer-Dedekind System for Binary Rings and Primitive Binary Forms

Theorem 12.

If 𝒪=[δ][δ1]\mathcal{O}=\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}] with δ¯\delta\in\overline{\mathbb{Q}} and if f(x,y)f(x,y) the minimal primitive binary form such that f(δ,1)=0f(\delta,1)=0 and shows the following decomposition/factorization modulo p,p,

f(x,y)i(fi)eimodp,f(x,y)\equiv\prod_{i}(f_{i})^{e_{i}}\bmod p,

then, we get

(\p)1𝒪=i𝒪ρi(\mathbb{Z}\backslash p\mathbb{Z})^{-1}\mathcal{O}=\bigcap_{i}\mathcal{O}_{\rho_{i}}

and

p𝒪=iβip\mathcal{O}=\prod_{i}\beta_{i}

where βi\beta_{i} are primary ideals in 𝒪\mathcal{O} sitting over pp such that

βi=ρi\sqrt{\beta_{i}}=\rho_{i}

with the understanding that

  • if fiY,f_{i}\neq Y, then

    ρi=(p[δ]+fi(δ,1)[δ])[δ1]\rho_{i}=(p\mathbb{Z}[\delta]+f_{i}(\delta,1)\mathbb{Z}[\delta])\cap\mathbb{Z}[\delta^{-1}]
  • and if fiX,f_{i}\neq X, then

    ρi=(p[δ1]+fi(1,δ1)[δ1])[δ]\rho_{i}=(p\mathbb{Z}[\delta^{-1}]+f_{i}(1,\delta^{-1})\mathbb{Z}[\delta^{-1}])\cap\mathbb{Z}[\delta]

such that ρi\rho_{i} are all distinct prime ideals and all prime ideals in 𝒪\mathcal{O} sitting over pp and

ief(ρi)=[𝕂:].\sum_{i}ef(\rho_{i})=[\mathbb{K}:\mathbb{Q}].
Theorem 13 (Dedekind’s Criterion Binary Rings).

We have

p|[𝒪𝕂:([δ][δ1])]i:1ik,fi¯|fi(fi)eip¯.p|[\mathcal{O}_{\mathbb{K}}:(\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}])]\iff\exists i:1\leq i\leq k,\>\overline{f_{i}}|\overline{\frac{f-\prod_{i}(f_{i})^{e_{i}}}{p}}.

Equivalently, we can say

p|[𝒪𝕂:([δ][α1])]g[x,y]m:f(p2,pg,g2)[x,y]np|[\mathcal{O}_{\mathbb{K}}:(\mathbb{Z}[\delta]\cap\mathbb{Z}[\alpha^{-1}])]\iff\exists g\in\mathbb{Z}[x,y]_{m}:f\in(p^{2},pg,g^{2})\subseteq\mathbb{Z}[x,y]_{n}

Before jumping into the proof we will review some properties of binary rings associated to primitive forms.

4.3. Localizations of Binary Rings associated to primitive forms

Let’s review Binary rings: All of the following results can be found in [11].

Let

f(X,Y):=anXn+an1Xn1Y++a0Ynf(X,Y):=a_{n}X^{n}+a_{n-1}X^{n-1}Y+\cdots+a_{0}Y^{n}

denote a binary form of degree nn. Let an0a_{n}\neq 0 and suppose that δ\delta denotes the image of XX in the algebra [X]/f(X,1).\mathbb{Q}[X]/f(X,1).

Definition 15.

When an0,a_{n}\neq 0,

Rf:=1,anδ,anδ2+an1δ,,i=0k1aniδki,,i=0n2aniδn1i.R_{f}:=\mathbb{Z}\langle 1,\>a_{n}\delta,\>a_{n}\delta^{2}+a_{n-1}\delta,\>\cdots,\>\sum_{i=0}^{k-1}a_{n-i}\delta^{k-i},\>\cdots,\>\sum_{i=0}^{n-2}a_{n-i}\delta^{n-1-i}\rangle. (2)

We set

B0,B1,,Bn1:=\displaystyle\langle B_{0},B_{1},\cdots,B_{n-1}\rangle:= 1,anδ,anδ2+an1δ,,i=0k1aniδki,,i=0n2aniδn1i\displaystyle\langle 1,\>a_{n}\delta,\>a_{n}\delta^{2}+a_{n-1}\delta,\>\cdots,\>\sum_{i=0}^{k-1}a_{n-i}\delta^{k-i},\>\cdots,\>\sum_{i=0}^{n-2}a_{n-i}\delta^{n-1-i}\rangle (3)
that is,
B0:=\displaystyle B_{0}:= 1\displaystyle 1 (4)
B1:=\displaystyle B_{1}:= anδ\displaystyle a_{n}\delta (5)
\displaystyle\cdots
\displaystyle\cdots
\displaystyle\cdots
Bk:=\displaystyle B_{k}:= anδk+an1δk1++ank+1δ\displaystyle a_{n}\delta^{k}+a_{n-1}\delta^{k-1}+\cdots+a_{n-k+1}\delta (6)
\displaystyle\cdots
Bn1:=\displaystyle B_{n-1}:= anδn1+an1δn2++a2δ\displaystyle a_{n}\delta^{n-1}+a_{n-1}\delta^{n-2}+\cdots+a_{2}\delta (7)
Definition 16.

IfI_{f} denotes the (fractional) ideal class generated by (1,δ)(1,\delta) in RfR_{f}

Theorem 14.

RfR_{f} is a ring of rank n.n.

Theorem 15.

Properties of RfR_{f}:

  1. (1)
    𝔡𝔦𝔰𝔠(Rf)=𝔡𝔦𝔰𝔠(f)\mathfrak{disc}_{\mathbb{Z}}(R_{f})=\mathfrak{disc}(f) (8)
  2. (2)

    If δ\delta is invertible, and ff is primitive, then

    Rf=[δ][δ1].R_{f}=\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}]. (9)
  3. (3)

    If ff is primitive, IfI_{f} is invertible in Rf.R_{f}.

  4. (4)

    RfR_{f} and IfI_{f} invariant a under the natural GL2()GL_{2}(\mathbb{Z}) action on binary forms of degree n.n.

  5. (5)

    We also have

    Rf+Rfδ=1,δ,B2,,Bn1R_{f}+R_{f}\delta=\mathbb{Z}\langle 1,\delta,B_{2},\cdots,B_{n-1}\rangle (10)

    and

    RfRfδ1=an,B1,B2,,Bn1R_{f}\cap R_{f}\delta^{-1}=\mathbb{Z}\langle a_{n},B_{1},B_{2},\cdots,B_{n-1}\rangle (11)

    and

    (Rf+Rfδ)(RfRfδ1)=gcd(an,an1,,a0),B1a1,B2a2,,Bn1an1.(R_{f}+R_{f}\delta)\cdot(R_{f}\cap R_{f}\delta^{-1})=\mathbb{Z}\langle gcd(a_{n},a_{n-1},\cdots,a_{0}),B_{1}-a_{1},B_{2}-a_{2},\cdots,B_{n-1}-a_{n-1}\rangle. (12)
Remark 16.

We make note of two facts. One is that the product (of modules) in eq. 12 is in fact equal to RfR_{f} when gcd(an,an1,,a0)=1gcd(a_{n},a_{n-1},\cdots,a_{0})=1 (that is, ff is primitive). The second is

Rf/RfRfδ1/an.\nicefrac{{R_{f}}}{{R_{f}\cap R_{f}\delta^{-1}}}\simeq\nicefrac{{\mathbb{Z}}}{{a_{n}\mathbb{Z}}}. (13)

This implies that the prime ideals in RfR_{f} containing RfRfδ1R_{f}\cap R_{f}\delta^{-1} are of the form pRf+(RfRfδ1)pR_{f}+(R_{f}\cap R_{f}\delta^{-1}) where pan.p\mid a_{n}. We can conclude that for each prime number dividing ana_{n} there exists exactly one prime in RfR_{f} containing pp and RfRfδ1R_{f}\cap R_{f}\delta^{-1}, and for panp\nmid a_{n} there is no prime ideal in RfR_{f} containing p.p. Furthermore, when panp\mid a_{n} the prime ideal in RfR_{f} containing pp and RfRfδ1R_{f}\cap R_{f}\delta^{-1} clearly has norm p.p.

Notation 17.

Let 𝒪=[δ][δ1].\mathcal{O}=\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}]. Let f(x,y)f(x,y) denote the minimal primitive binary form satisfying f(δ,1)=0.f(\delta,1)=0.

Thus, Rf𝒪.R_{f}\simeq\mathcal{O}.

Lemma 5.
(𝒪𝒪δ1)[δ]=[δ](\mathcal{O}\cap\mathcal{O}\delta^{-1})\mathbb{Z}[\delta]=\mathbb{Z}[\delta]
Proof.

We know that from item 5,

𝒪𝒪δ1=an,B1,B2,,Bn1.\mathcal{O}\cap\mathcal{O}\delta^{-1}=\mathbb{Z}\langle a_{n},B_{1},B_{2},\cdots,B_{n-1}\rangle.

Thus, (𝒪𝒪δ1)[δ](\mathcal{O}\cap\mathcal{O}\delta^{-1})\mathbb{Z}[\delta] is spanned by the set organized in the table,

ana_{n} B1B_{1} B2B_{2} \cdots Bn1B_{n-1}
anδa_{n}\delta B1δB_{1}\delta B2δB_{2}\delta \cdots Bn1δB_{n-1}\delta
anδ2a_{n}\delta^{2} B1δ2B_{1}\delta^{2} B2δ2B_{2}\delta^{2} \cdots Bn1δ2B_{n-1}\delta^{2}
anδ3a_{n}\delta^{3} B1δ3B_{1}\delta^{3} B2δ3B_{2}\delta^{3} \cdots Bn1δ3B_{n-1}\delta^{3}
\vdots \vdots \vdots \vdots \vdots

We note that Bi+1Biδ=aniB_{i+1}-B_{i}\delta=a_{n-i} is in the \mathbb{Z}-span of the above set for all n1i1n-1\geq i\geq 1. We can also see that a0=Bn1δa_{0}=-B_{n-1}\delta is also in the span. Thus,

1=gcd(a0,a1,a2,,an)(𝒪𝒪δ1)[δ].1=\gcd(a_{0},a_{1},a_{2},\cdots,a_{n})\in(\mathcal{O}\cap\mathcal{O}\delta^{-1})\mathbb{Z}[\delta].

Result follows. ∎

We note that if ρM(𝒪)\rho\in M(\mathcal{O}) (prime ideal in 𝒪\mathcal{O}) such that

(𝒪𝒪δ1)ρ(\mathcal{O}\cap\mathcal{O}\cdot\delta^{-1})\not\subseteq\rho

i.e. the denominator (proper) ideal of δ\delta in 𝒪\mathcal{O} is not contained in ρ,\rho, then t𝒪𝒪δ1\exists t\in\mathcal{O}\cap\mathcal{O}\cdot\delta^{-1} such that t1𝒪ρ.t^{-1}\in\mathcal{O}_{\rho}. Since, tδ𝒪t\delta\in\mathcal{O} it follows that [δ]𝒪ρ.\mathbb{Z}[\delta]\subseteq\mathcal{O}_{\rho}. For such primes ρ\rho (primes not containing 𝒪𝒪δ1\mathcal{O}\cap\mathcal{O}\cdot\delta^{-1}), we note that [δ]/ρ[δ]\nicefrac{{\mathbb{Z}[\delta]}}{{\rho\mathbb{Z}[\delta]}} contains 𝒪/ρ\nicefrac{{\mathcal{O}}}{{\rho}} and is contained in 𝒪ρ/ρ𝒪ρ𝒪/ρ\nicefrac{{\mathcal{O}_{\rho}}}{{\rho\mathcal{O}_{\rho}}}\simeq\nicefrac{{\mathcal{O}}}{{\rho}} and hence must be equal to 𝒪/ρ\nicefrac{{\mathcal{O}}}{{\rho}} forcing ρ[δ]\rho\mathbb{Z}[\delta] to be a prime ideal in [δ]\mathbb{Z}[\delta]. This identifies ρM(𝒪)\rho\in M(\mathcal{O}) with a prime in [δ]\mathbb{Z}[\delta] in a localization preserving way. That is to say,

[δ]ρ[ρ]=𝒪ρ.\mathbb{Z}[\delta]_{\rho\mathbb{Z}[\rho]}=\mathcal{O}_{\rho}.

Combining with lemma 5, we argue that the only prime ideals in 𝒪\mathcal{O} which will not be mapped to any prime ideal in [δ]\mathbb{Z}[\delta] will be exactly those prime ideals containing 𝒪𝒪δ1\mathcal{O}\cap\mathcal{O}\cdot\delta^{-1}

So, if we look at the map M([δ])M(𝒪)M(\mathbb{Z}[\delta])\longrightarrow M(\mathcal{O}) given by ρρ𝒪\rho\longrightarrow\rho\cap\mathcal{O}, then this map is clearly the inverse of the localization preserving map above.

4.4. Proof of Dedekind Kummer in Binary/One-fine Rings or theorem 12.

This is easily observed using the previous understanding of prime ideals in 𝒪=[δ][δ1]\mathcal{O}=\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}] and their relationship to prime ideals in [δ]\mathbb{Z}[\delta] and the identification

[δ][x]/(f(x,1)).\mathbb{Z}[\delta]\simeq\nicefrac{{\mathbb{Z}[x]}}{{(f(x,1))}}.

Suppose, f(x,y)f(x,y) has the following decomposition/factorization modulo p.p.

f(x,y)yki(fi)eimodp,f(x,y)\equiv y^{k}\prod_{i}(f_{i})^{e_{i}}\bmod p,

where fif_{i} are irreducible binary forms modp\bmod p such that fi(x,1)f_{i}(x,1) has the same degree in XX as the binary form fif_{i} in (x,y),(x,y), (note that this will not happen if and only if fi=yf_{i}=y). We set

ρi=(p[δ]+fi(δ,1)ei[δ]).\rho_{i}=(p\mathbb{Z}[\delta]+f_{i}(\delta,1)^{e_{i}}\mathbb{Z}[\delta]).

It follows that

[δ]/ρi[X]/(p,fi(x,1)ei,f(x,1)[X]/(p,fi(x,1))𝔽p[x]/(fi(x)ei).\nicefrac{{\mathbb{Z}[\delta]}}{{\rho_{i}}}\simeq\nicefrac{{\mathbb{Z}[X]}}{{(p,f_{i}(x,1)^{e_{i}},f(x,1)}}\simeq\nicefrac{{\mathbb{Z}[X]}}{{(p,f_{i}(x,1))}}\simeq\nicefrac{{\mathbb{F}_{p}[x]}}{{(f_{i}(x)^{e_{i}})}}.

Thus, ρi\rho_{i} are primary ideals in [δ]\mathbb{Z}[\delta] with

ρi=(p,fi(δ)).\sqrt{\rho_{i}}=(p,f_{i}(\delta)).

On the other hand, it is easy to see that

(p,f(x,1))=i(p,fi(x,1)ei)\displaystyle(p,f(x,1))=\prod_{i}(p,f_{i}(x,1)^{e_{i}})
where ij(p,fi(x,1)ei)+(p,fj(x,1)ej)=[x],\displaystyle\textit{where }i\neq j\implies(p,f_{i}(x,1)^{e_{i}})+(p,f_{j}(x,1)^{e_{j}})=\mathbb{Z}[x],

since fi(x,1)modpf_{i}(x,1)\bmod p and fj(x,1)modpf_{j}(x,1)\bmod p are irreducible distinct polynomials in 𝔽p[x].\mathbb{F}_{p}[x]. Thus,

[δ]/(p)i𝔽p[x]/(fi(x,1)ei).\nicefrac{{\mathbb{Z}[\delta]}}{{(p)}}\simeq\prod_{i}\nicefrac{{\mathbb{F}_{p}[x]}}{{(f_{i}(x,1)^{e_{i}})}}.

We can conclude that ρi\sqrt{\rho_{i}} are all distinct prime ideals and all prime ideals in [δ]\mathbb{Z}[\delta] containing pp. This concludes the proof of the Dedekind Kummer theorem for Binary rings associated to primitive forms.

4.5. Proof of Dedekind’s Criterion for Primitive binary forms or theorem 13

The proof is very much the same as we can look at [δ]\mathbb{Z}[\delta] and [δ1]\mathbb{Z}[\delta^{-1}] separately.

Suppose that f(x,y)f(x,y) is the minimal primitive binary form satisfying f(δ,1)=0.f(\delta,1)=0.

Notation 18.

Given a polynomial h(x)[x]h(x)\in\mathbb{Z}[x] we let its image in 𝔽p[x]\mathbb{F}_{p}[x] be denoted by h¯(x).\bar{h}(x).

Thus, if f(x,y)f(x,y) shows the following decomposition/factorization modulo p,p,

f(x,y)yki(fi)eimodp,f(x,y)\equiv y^{k}\prod_{i}(f_{i})^{e_{i}}\bmod p,

where fif_{i} are irreducible binary forms modp\bmod p such that fi(x,1)f_{i}(x,1) has the same degree in xx as the binary form fif_{i} in (x,y),(x,y), (we note that this will not happen if and only if fi=yf_{i}=y) then we may set

βi=(p[δ]+fi(δ,1)[δ])\beta_{i}=(p\mathbb{Z}[\delta]+f_{i}(\delta,1)\mathbb{Z}[\delta])

as the prime ideal associated to fi.f_{i}.

It follows that

βi2=(p2[δ]+pfi(δ,1)[δ]+fi(δ,1)2[δ])\beta_{i}^{2}=(p^{2}\mathbb{Z}[\delta]+pf_{i}(\delta,1)\mathbb{Z}[\delta]+f_{i}(\delta,1)^{2}\mathbb{Z}[\delta])

Based on the structure of the prime ideal, we conclude a random element

t=m(δ)βi if and only if fi¯(x,1)|m¯(x).t=m(\delta)\in\beta_{i}\textit{ if and only if }\bar{f_{i}}(x,1)|\bar{m}(x). (14)

We will consider two cases:

fi(δ,1)βi2[δ]βif_{i}(\delta,1)\in\beta_{i}^{2}\mathbb{Z}[\delta]_{\beta_{i}} and fi(δ,1)βi2[δ]βi.f_{i}(\delta,1)\not\in\beta_{i}^{2}\mathbb{Z}[\delta]_{\beta_{i}}.

  • Case: fi(δ,1)βi2[δ]βi.f_{i}(\delta,1)\in\beta_{i}^{2}\mathbb{Z}[\delta]_{\beta_{i}}.

    fi(δ,1)βi2[δ]βi\displaystyle f_{i}(\delta,1)\in\beta_{i}^{2}\mathbb{Z}[\delta]_{\beta_{i}}\implies fi(δ,1)=p2A+pfi(δ,1)B+fi(δ,1)2C\displaystyle f_{i}(\delta,1)=p^{2}A+pf_{i}(\delta,1)B+f_{i}(\delta,1)^{2}C
    \displaystyle\iff fi(δ,1)(1pBfi(δ,1)C)=p2A\displaystyle f_{i}(\delta,1)(1-pB-f_{i}(\delta,1)C)=p^{2}A

    Since, 1pBfi(δ,1)Cβi,1-pB-f_{i}(\delta,1)C\notin\beta_{i}, it is a unit this means that p2R(δ)=fi(δ,1)S(δ)p^{2}R(\delta)=f_{i}(\delta,1)S(\delta) where S(δ)S(\delta) is a unit in [δ]βi.\mathbb{Z}[\delta]_{\beta_{i}}.

    In this case, [δ]βi\mathbb{Z}[\delta]_{\beta_{i}} is a DVR where the uniformizer can be taken to be pp as it generates the maximal ideal in [δ]βi.\mathbb{Z}[\delta]_{\beta_{i}}. Furthermore, since the uniformizer of [δ]βip\mathbb{Z}[\delta]_{\beta_{i}}\otimes\mathbb{Z}_{p} is the same as the uniformizer of p\mathbb{Z}_{p}, [δ]βip\mathbb{Z}[\delta]_{\beta_{i}}\otimes\mathbb{Z}_{p} is an un-ramified extension of p.\mathbb{Z}_{p}. This forces ei=1,e_{i}=1, which implies f(x,y)(p2,pfi,(fi)2).f(x,y)\notin(p^{2},pf_{i},(f_{i})^{2}).

  • Case: fi(δ,1)βi2[δ]βi.f_{i}(\delta,1)\notin\beta_{i}^{2}\mathbb{Z}[\delta]_{\beta_{i}}.

    In this case, fi(δ,1)βi[δ]βi\βi2[δ]βi.f_{i}(\delta,1)\in\beta_{i}\mathbb{Z}[\delta]_{\beta_{i}}\backslash\beta_{i}^{2}\mathbb{Z}[\delta]_{\beta_{i}}. Thus, if [δ]βi\mathbb{Z}[\delta]_{\beta_{i}} is a DVR then fi(δ,1)f_{i}(\delta,1) may be chosen as a uniformizer.

    We see [δi]p\mathbb{Z}[\delta_{i}]\otimes\mathbb{Z}_{p} as the ring of integers of a totally ramified extension of a totally un-ramified extension of p.\mathbb{Z}_{p}. Since, [δ]/βi\nicefrac{{\mathbb{Z}[\delta]}}{{\beta_{i}}} is a deg(fi)deg(f_{i}) extension of 𝔽p.\mathbb{F}_{p}. Thus, the ramification degree of [δi]p\mathbb{Z}[\delta_{i}]\otimes\mathbb{Z}_{p} is deg(fi)deg(f_{i}) and the ramification index is eie_{i}. This will imply that

    fi(δ,1)ei=puf_{i}(\delta,1)^{e_{i}}=pu

    for some unit uu.

    Now substituting (δ,1)(\delta,1) in f(x,y)=ykjfj(x,y)ei+ph(x,y)f(x,y)=y^{k}\prod_{j}f_{j}(x,y)^{e_{i}}+ph(x,y) we get

    jfj(δ,1)ej=ph(δ,1)\prod_{j}f_{j}(\delta,1)^{e_{j}}=-ph(\delta,1)

    Since for ji,j\neq i, fj(δ,1)f_{j}(\delta,1) are all units in [δ]βi,\mathbb{Z}[\delta]_{\beta_{i}}, as they, by definition, are distinct irreducible polynomials over 𝔽p.\mathbb{F}_{p}.

    This means that h(δ,1)h(\delta,1) has to be a unit in [δ]βi\mathbb{Z}[\delta]_{\beta_{i}} i.e. f¯i(x,1)h¯(x,1)\bar{f}_{i}(x,1)\nmid\bar{h}(x,1) (using eq. 14) which is equivalent to h(x,y)(p,fi(x,y))f(x,y)(p2,pfi,fi2).h(x,y)\notin(p,f_{i}(x,y))\iff f(x,y)\notin(p^{2},pf_{i},f_{i}^{2}).

To show the other direction, we note that, if ei=1e_{i}=1 then clearly f(x,y)(p2,pfi,fi2)f(x,y)\notin(p^{2},pf_{i},f_{i}^{2}). At the same time,

jfj(δ,1)ej=ph(δ,1)\prod_{j}f_{j}(\delta,1)^{e_{j}}=-ph(\delta,1)

and for ji,j\neq i, fj(δ,1)f_{j}(\delta,1) are all units in [δ]βi\mathbb{Z}[\delta]_{\beta_{i}} which means that βi[δ]βi=p[δ]βi\beta_{i}\mathbb{Z}[\delta]_{\beta_{i}}=p\mathbb{Z}[\delta]_{\beta_{i}} making it a principal ideal and thus [δ]βi\mathbb{Z}[\delta]_{\beta_{i}} a DVR.

If ei>1e_{i}>1 and f(p2,pfi,fi2)f\notin(p^{2},pf_{i},f_{i}^{2}) then h(x,y)(p,fi(x,y)).h(x,y)\notin(p,f_{i}(x,y)). Subbing in δ\delta we see that this means that h(δ,1)h(\delta,1) is a unit in [δ]βi.\mathbb{Z}[\delta]_{\beta_{i}}. At the same time,

jfj(δ,1)ej=ph(δ,1)\prod_{j}f_{j}(\delta,1)^{e_{j}}=-ph(\delta,1)

and for ji,j\neq i, fj(δ,1)f_{j}(\delta,1) are all units in [δ]βi\mathbb{Z}[\delta]_{\beta_{i}} which means that βi[δ]βi=fi(δ,1)[δ]βi\beta_{i}\mathbb{Z}[\delta]_{\beta_{i}}=f_{i}(\delta,1)\mathbb{Z}[\delta]_{\beta_{i}} making it a principal ideal and thus [δ]βi\mathbb{Z}[\delta]_{\beta_{i}} a DVR. This completes the proof.

5. Small prime splitting restrictions.

Just as a small prime splitting poses an obstruction for a ring to be monogenic, a small prime splitting poses a restriction for a ring to be binary. In this section we will quatify this restriction.

Let 𝕂\mathbb{K} denote a finite extension of \mathbb{Q}. Let M𝕂M_{\mathbb{K}} denote the set of non-archimedean places of K.K. We fix a prime pp in \mathbb{Z} for this section. Let 𝒪\mathcal{O} denote an order in 𝒪𝕂\mathcal{O}_{\mathbb{K}} Let p𝒪=δ1δ2δrp\mathcal{O}=\delta_{1}\cdot\delta_{2}\cdots\delta_{r} denote the unique m-primary ideal factorization of the ideal p𝒪p\mathcal{O} in 𝒪\mathcal{O} where δi+δj=𝒪.\delta_{i}+\delta_{j}=\mathcal{O}. Let τi\tau_{i} denote the radical of the ideal δi\delta_{i} in R.R. These will be all the maximal ideals of 𝒪\mathcal{O} containing pp. For any non zero proper ideal II of 𝒪\mathcal{O} we define

Norm𝒪(I)=|𝒪/I|.Norm_{\mathcal{O}}(I)=|\nicefrac{{\mathcal{O}}}{{I}}|.

And let

T𝒪(p,f):={τi:Norm(τi)=pf}H(p,f):=1fd|fpdμ(fd)T_{\mathcal{O}}(p,f):=\{\tau_{i}:Norm(\tau_{i})=p^{f}\}\qquad H(p,f):=\frac{1}{f}\sum_{d|f}p^{d}\mu(\frac{f}{d})
Remark 17.

H(p,f)H(p,f) is the number of prime ideals in 𝔽p[x]\mathbb{F}_{p}[x] with the norm pf.p^{f}.

Theorem 18.

If 𝒪\mathcal{O} is binary ring, then

H(p,1)+1|T𝒪(p,f)|\displaystyle H(p,1)+1\geq|T_{\mathcal{O}}(p,f)|
f2H(p,f)|T𝒪(p,f)|.\displaystyle f\geq 2\implies H(p,f)\geq|T_{\mathcal{O}}(p,f)|.
Proof.

Let 𝒪=[δ][δ1]\mathcal{O}=\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}]. From Dedekind Kummer Theorem for Binary forms (theorem 12), we have a neat (norm preserving) correspondence from all prime ideals of [δ]\mathbb{Z}[\delta] containing pp to either all or all but one prime ideal(that of norm pp) of 𝒪\mathcal{O}. We identify [δ]\mathbb{Z}[\delta] with [X]/g(x),\nicefrac{{\mathbb{Z}[X]}}{{g(x)}}, where g(x)=f(x,1).g(x)=f(x,1). Let A=|T𝒪(p,f)|.A=|T_{\mathcal{O}}(p,f)|.

Consider the following diagram of canonical quotient maps,

[x]{\mathbb{Z}[x]}[δ][x](g(x)){\mathbb{Z}[\delta]\simeq\frac{\mathbb{Z}[x]}{(g(x))}}𝔽p[x]{\mathbb{F}_{p}[x]}𝔽p[δ¯][x](p,g(x)){\mathbb{F}_{p}[\bar{\delta}]\simeq\frac{\mathbb{Z}[x]}{(p,g(x))}}Z[δ]τAτ(𝔽pf)|A|{\dfrac{Z[\delta]}{\displaystyle\prod_{\tau\in A}\tau}\simeq(\mathbb{F}_{p^{f}})^{|A|}}π\scriptstyle{\pi}

Clearly each arrow is surjective. Consider the lower row of maps culminating in the following surjective ring homo-morphism

𝔽p[x](𝔽pf)|A|\mathbb{F}_{p}[x]\longrightarrow(\mathbb{F}_{p^{f}})^{|A|}

In particular, there must exist at least |A||A| distinct prime ideals in 𝔽p[x]\mathbb{F}_{p}[x] with norm pf.p^{f}. This corresponds to surjective maps from 𝔽p[x]\mathbb{F}_{p}[x] to 𝔽pf\mathbb{F}_{p^{f}} up-to Gal(𝔽pf/𝔽p)Gal(\mathbb{F}_{p^{f}}/\mathbb{F}_{p}) action. In other words, the number of degree ff elements in 𝔽pf\mathbb{F}_{p^{f}} up-to Gal(𝔽pf/𝔽p)Gal(\mathbb{F}_{p^{f}}/\mathbb{F}_{p}) action. This is H(p,f).H(p,f). Thus, H(p,f)|A|H(p,f)\geq|A|. Combining with Dedekind Kummer Theorem for Binary rings, we see that,

f2\displaystyle f\geq 2\implies H(p,f)|T𝒪(p,f)|\displaystyle H(p,f)\geq|T_{\mathcal{O}}(p,f)|
H(p,1)+1|T𝒪(p,f)|.\displaystyle H(p,1)+1\geq|T_{\mathcal{O}}(p,f)|.

Remark 19.

If

(p,f):=#{ prime ideals in (𝔽¯p) with norm pf},\mathbb{P}(p,f):=\#\{\textit{ prime ideals in }\mathbb{P}(\overline{\mathbb{F}}_{p})\textit{ with norm }p^{f}\},

then,

(p,1)=H(p,1)+1\mathbb{P}(p,1)=H(p,1)+1

and

f2(p,f)=H(p,f).f\geq 2\implies\mathbb{P}(p,f)=H(p,f).
Remark 20.

Binary rings are best viewed as a subset of orders which are locally monogenic, more specifically there are locally monogenic rings which are almost globally monogenic. Monogenic rings are globally monogenic (simultaneously across all primes). In case of n=2n=2 the concept of freely locally monogenic is equivalent to the version of almost globally monogenic used here. We compare this to, “Every ideal class in a rank two ring occurs in binary quadratic forms.” When n=3n=3, this concept of locally monogenic is the same as almost globally monogenic used here, but not freely locally monogenic. We compare this to, “Every ring of integers occurs as a binary ring not every ideal class in ring of integers occurs as a binary ring and associated invertible ideal.” We generalize the concept in our thesis and in upcoming papers and relate it to other classification of ring theorems.

Remark 21.

We compare this condition with the small prime splitting condition in monogenic rings, which is

f1H(p,f)|T𝒪(p,f)|.f\geq 1\implies H(p,f)\geq|T_{\mathcal{O}}(p,f)|.

In-spite of a very minimal weakening on this condition we note that the rings captured as monogenic rings are expected to be proportion 0 in binary rings (almost globally monogenic rings).

6. Pseudo Maximal Orders and Sudo Maximal orders.

Remark 22.

In general, if we have an irreducible order 𝒪\mathcal{O} with S(𝒪)={ρ}S(\mathcal{O})=\{\rho\} such that ρ¯={v1,v2,,vk}\overline{\rho}=\{v_{1},v_{2},\cdots,v_{k}\} and if we let the conductor of 𝒪\mathcal{O} to be c𝒪,c_{\mathcal{O}}, an ideal in 𝒪𝕂,\mathcal{O}_{\mathbb{K}}, then we know that

c𝒪=(v1)av1(v2)av2(vk)avkc_{\mathcal{O}}=(v_{1})^{a_{v_{1}}}(v_{2})^{a_{v_{2}}}\cdots(v_{k})^{a_{v_{k}}}

where ai1.a_{i}\geq 1.

Furthermore, the since c𝒪𝒪𝒪𝕂c_{\mathcal{O}}\subseteq\mathcal{O}\subseteq\mathcal{O}_{\mathbb{K}} we see that 𝒪\mathcal{O} is completely determined by 𝒪/c𝒪\nicefrac{{\mathcal{O}}}{{c_{\mathcal{O}}}} a sub-ring of 𝒪𝕂/c𝒪.\nicefrac{{\mathcal{O}_{\mathbb{K}}}}{{c_{\mathcal{O}}}}. Hence, we can search for sub-rings of the appropriate complete local finite extension of pp-adic integers.

From here it is easy to get the Furtwrangler’s condition for a conductor ideals by just looking at the ideal (pm,c𝒪)(p^{m},c_{\mathcal{O}}) in a reduced order where 𝒪\mathcal{O} where pm+1c𝒪p^{m+1}\in c_{\mathcal{O}} and pmc𝒪p^{m}\notin c_{\mathcal{O}} and observing when it will be an ideal 𝒪𝕂\mathcal{O}_{\mathbb{K}}. This ideal can be an ideal of 𝒪𝕂\mathcal{O}_{\mathbb{K}} if and only if c𝒪c_{\mathcal{O}} cannot be a conductor ideal. In fact, combining with Fundamental Theorem of Orders or theorem 7 it gives a more nuanced answer about orders and associated conductors and what can exist. For example, it tells us that there cannot exist an order 𝒪\mathcal{O} with conductor vwvw with fv=ev=fw=ew=1f_{v}=e_{v}=f_{w}=e_{w}=1 and 𝒪/vwp2.\nicefrac{{\mathcal{O}}}{{vw}}\simeq\mathbb{Z}_{p}^{2}. This one is obvious, by things we already know.

6.1. Classification of Irreducible orders 𝒪\mathcal{O} with ef(𝒪)=1ef(\mathcal{O})=1 and ef(𝒪)=2ef(\mathcal{O})=2

6.1.1. ef(𝒪)=1ef(\mathcal{O})=1

There exist no irreducible order 𝒪\mathcal{O} in 𝒪K\mathcal{O}_{K} with ef(𝒪)=1ef(\mathcal{O})=1 as this would be akin to finding proper unital sub-rings of the p-adic integers p\mathbb{Z}_{p}.

6.1.2. ef(𝒪)=2ef(\mathcal{O})=2

Lemma 6.

If A=p+ptA=\mathbb{Z}_{p}+\mathbb{Z}_{p}\cdot t is a algebra of rank 22 over p\mathbb{Z}_{p} then all sub-rings of AA which have rank 22 over p\mathbb{Z}_{p} and containing p\mathbb{Z}_{p} are given by p+pk(A)\mathbb{Z}_{p}+p^{k}(A) for some k0.k\in\mathbb{Z}^{\geq 0}. The index of p+pk(A)\mathbb{Z}_{p}+p^{k}(A) in AA is pk.p^{k}.

Proof.

If MM is a sub-ring of AA which is rank 22 over p\mathbb{Z}_{p} may be seen as p+pr\mathbb{Z}_{p}+\mathbb{Z}_{p}r where r=a+btr=a+bt with a,bpa,b\in\mathbb{Z}_{p}. We further write b=pkub=p^{k}u where upu\in\mathbb{Z}_{p} is a unit, i.e. up=pu\mathbb{Z}_{p}=\mathbb{Z}_{p} Clearly,

p+pr=p+pbt=p+p+pupkt=p+p(pkt)=p+pkA.\mathbb{Z}_{p}+\mathbb{Z}_{p}r=\mathbb{Z}_{p}+\mathbb{Z}_{p}bt=\mathbb{Z}_{p}+\mathbb{Z}_{p}+\mathbb{Z}_{p}up^{k}t=\mathbb{Z}_{p}+\mathbb{Z}_{p}(p^{k}t)=\mathbb{Z}_{p}+p^{k}A.

Remark 23.

There is no irreducible(non 𝒪𝕂\mathcal{O}_{\mathbb{K}}) order 𝒪\mathcal{O} with ef(𝒪)=1.ef(\mathcal{O})=1. Consequence of the Furtwangler condition for an ideal to be a conductor ideal. We give a structure theorem for irreducible orders with ef(𝒪)=2ef(\mathcal{O})=2. Similar Structure theorems can be given easily using corresponding ring classification theories for irreducible orders with ef(𝒪)4.ef(\mathcal{O})\leq 4. With some difficulty, the authors also expect such theorems possible for ef(𝒪)=5ef(\mathcal{O})=5 by extending Bhargava’s work for quintic rings to p.\mathbb{Z}_{p}.

Definition 19.

A Pseudo Maximal Order in 𝕂\mathbb{K} is an irreducible order 𝒪\mathcal{O} satisfying ef(𝒪)=2.ef(\mathcal{O})=2. See definition 13 for the definition of ef.ef.

If 𝒪\mathcal{O} is an irreducible order with S(𝒪)={ρ}S(\mathcal{O})=\{\rho\}, then its conductor is made up of prime ideals (in 𝕂\mathbb{K}) in ρ¯\overline{\rho}. Thus, when

Lemma 7.

𝒪\mathcal{O} is a Pseudo Maximal Order, then one of the following must be true:

ρ¯=\displaystyle\overline{\rho}= {v,w} with fv=fw=ev=ew=1\displaystyle\{v,w\}\textit{ with }f_{v}=f_{w}=e_{v}=e_{w}=1
=\displaystyle= {v} with fv=2=2ev\displaystyle\{v\}\textit{ with }f_{v}=2=2e_{v}
=\displaystyle= {w} with 2fv=2=ev.\displaystyle\{w\}\textit{ with }2f_{v}=2=e_{v}.
Theorem 24 (A).

If {v,w}M(𝕂) with fv=fw=ev=ew=1,\{v,w\}\subseteq M(\mathbb{K})\textit{ with }f_{v}=f_{w}=e_{v}=e_{w}=1, then the irreducible orders with conductor vawbv^{a}w^{b} are given by

  • if a=b=ra=b=r then there is a unique irreducible order 𝒪\mathcal{O} with conductor (vw)r(vw)^{r} given by +(vw)r.\mathbb{Z}+(vw)^{r}. This order will have index in 𝒪𝕂\mathcal{O}_{\mathbb{K}} given by [𝒪𝕂:+(vw)r]=pr.[\mathcal{O}_{\mathbb{K}}:\mathbb{Z}+(vw)^{r}]=p^{r}.

  • if aba\neq b then there is no irreducible (or otherwise) order with conductor vawb.v^{a}w^{b}.

Proof.

Clearly, we are looking for irreducible orders in pp\mathbb{Z}_{p}\oplus\mathbb{Z}_{p} which is treated an extension of p\mathbb{Z}_{p} via diagonal embedding. So, we may write it as p+pt\mathbb{Z}_{p}+\mathbb{Z}_{p}t where t=(0,1)t=(0,1). Thus every order here must be of the form p+pk(pp).\mathbb{Z}_{p}+p^{k}(\mathbb{Z}_{p}\oplus\mathbb{Z}_{p}). Coordinate wise p=(p,p)p=(p,p) due to diagonal embedding. Thus, the sub-ring is given by p+((p,1)k(1,p)k)(pp)\mathbb{Z}_{p}+((p,1)^{k}\cdot(1,p)^{k})(\mathbb{Z}_{p}\oplus\mathbb{Z}_{p}). Intersecting with 𝒪𝕂\mathcal{O}_{\mathbb{K}} we get the order is given by +(vw)k\mathbb{Z}+(vw)^{k} whose conductor is easily seen as (vw)k.(vw)^{k}.

Theorem 25 (B).

If {v}M(𝕂) with fv=2=2ev,\{v\}\subseteq M(\mathbb{K})\textit{ with }f_{v}=2=2e_{v}, then there is a unique irreducible order with conductor vav^{a} which is given by +va\mathbb{Z}+v^{a} with [𝒪𝕂:(+va)]=pa[\mathcal{O}_{\mathbb{K}}:(\mathbb{Z}+v^{a})]=p^{a}

Proof.

Clearly, we are looking for irreducible orders in the ring of integers of the unique totally un-ramified extension of degree 2 over p.\mathbb{Z}_{p}. Thus, pp may be seen as the uniformizer for this completion. So, we may write it as Ov:=p+ptO_{v}:=\mathbb{Z}_{p}+\mathbb{Z}_{p}t. Thus every order here must be of the form p+pk(𝒪v).\mathbb{Z}_{p}+p^{k}(\mathcal{O}_{v}). Thus, the sub-ring is given by p+pk(Ov)\mathbb{Z}_{p}+p^{k}(O_{v}). Intersecting with 𝒪𝕂\mathcal{O}_{\mathbb{K}} we get the order is given by +vk\mathbb{Z}+v^{k} whose conductor is easily seen as vk.v^{k}.

Theorem 26 (C).

If {v}M(𝕂) with 2fv=2=ev,\{v\}\subseteq M(\mathbb{K})\textit{ with }2f_{v}=2=e_{v}, then

  • if a2a\geq 2 and aa is even then there is a unique irreducible order with conductor vav^{a} which is given by +va\mathbb{Z}+v^{a} with [𝒪𝕂:(+va)]=pa/2.[\mathcal{O}_{\mathbb{K}}:(\mathbb{Z}+v^{a})]=p^{a/2}.

  • if a1a\geq 1 and aa is odd then there is no irreducible (or otherwise) order with conductor va.v^{a}.

Proof.

Clearly, we are looking for irreducible orders in the ring of integers of some totally ramified extension of degree 2 over p.\mathbb{Z}_{p}. Thus, pp may be seen as the square of the uniformizer for this completion. So, we may write it as Ov:=p+ptO_{v}:=\mathbb{Z}_{p}+\mathbb{Z}_{p}t where tt is the uniformizer(since it is totally ramified. Thus, every order here must be of the form p+pk(𝒪v).\mathbb{Z}_{p}+p^{k}(\mathcal{O}_{v}). Thus, the sub-ring is given by p+pk(Ov)\mathbb{Z}_{p}+p^{k}(O_{v}). Intersecting with 𝒪𝕂\mathcal{O}_{\mathbb{K}} we get the order is given by +v2k\mathbb{Z}+v^{2k} (as (p)=v2(p)=v^{2}, as it is a totally ramified extension) whose conductor is easily seen as v2k.v^{2k}.

A corollary we will use for counting purposes is,

Corollary 1.

Given a set of valuations SM(𝕂)S\subseteq M(\mathbb{K}) such that vSevfv=2\sum_{v\in S}e_{v}f_{v}=2 then given r1r\geq 1 there exists a unique irreducible order 𝒪\mathcal{O} with S(𝒪)={ρ}S(\mathcal{O})=\{\rho\} such that ρ¯=S\overline{\rho}=S and [𝒪𝕂:𝒪]=pr.[\mathcal{O}_{\mathbb{K}}:\mathcal{O}]=p^{r}.

Definition 20.

A Sudo Maximal Order in 𝕂\mathbb{K} is an order 𝒪\mathcal{O} such that

ρS(𝒪):ef(ρ)=2.\displaystyle\forall\rho\in S(\mathcal{O}):ef(\rho)=2.

We say that this is Restricted Sudo Maximal Order if for every prime pp\in\mathbb{Z} we have at most one prime ideal in S(𝒪)S(\mathcal{O}) which contains p.p.

Theorem 27.

The number of Restricted Sudo Maximal Orders in 𝕂\mathbb{K} with index X\leq X is

AnXlog(X)(n2)1\leq A_{n}X\log(X)^{\binom{n}{2}-1}

where AnA_{n} is a constant only dependent on n.n.

Proof.

We note that if C𝕂,pkC_{\mathbb{K},p^{k}} denote the number of Restricted Sudo Maximal Order in 𝕂\mathbb{K} with index pk,p^{k}, we may look at

S𝕂,p(s):=k=0C𝕂,pkpksS_{\mathbb{K},p}(s):=\sum_{k=0}^{\infty}C_{\mathbb{K},p^{k}}p^{-ks}

and look at the L-function

L𝕂(s):=pS𝕂,pL_{\mathbb{K}}(s):=\prod_{p\in\mathbb{P}}S_{\mathbb{K},p}

Since we know from the above section that

C𝕂,pk=\displaystyle C_{\mathbb{K},p^{k}}= |{(v,w):v,wM(𝕂),fv=fw=1,vw,v(p)=w(p)=1}|\displaystyle|\{(v,w):v,w\in M(\mathbb{K}),f_{v}=f_{w}=1,v\neq w,v(p)=w(p)=1\}|
+|{vM(𝕂):fv=2v(p)=1}|+|{vM(𝕂):fv=1v(p)=2}|.\displaystyle+|\{v\in M(\mathbb{K}):f_{v}=2v(p)=1\}|+|\{v\in M(\mathbb{K}):f_{v}=1v(p)=2\}|.

Thus, C𝕂,pk(n2).C_{\mathbb{K},p^{k}}\leq\binom{n}{2}.

It follows that the Dirichlet coefficients of L𝕂(s)L_{\mathbb{K}}(s) are all between 0 and the Dirichlet coefficients of ζ(s)(n2)\zeta(s)^{\binom{n}{2}}.

Sum of Dirichlet coefficients up to XX of ζ(s)(n2)\zeta(s)^{\binom{n}{2}} grows asymptotically like

Xlog(X)(n2)1((n2)1)!.\frac{X\log(X)^{\binom{n}{2}-1}}{(\binom{n}{2}-1)!}.

Thus, the number of orders in 𝕂\mathbb{K} with index X\leq X is

AnXlog(X)(n2)1((n2)1)!\leq A_{n}\frac{X\log(X)^{\binom{n}{2}-1}}{(\binom{n}{2}-1)!}

for some constant AnA_{n} only dependant on n.n.

Remark 28.

We note that this gives a strong indication that the number of orders in 𝕂\mathbb{K} with index bounded by XX should be of the order of X.X.

6.2. Strongly divisible-ness and Ultra-weakly divisible polynomials

Remark 29.

Recall that Bhargava-Shankar-Wang define ff is strongly divisible by pp (in [3] and [2]) if p2|𝔡𝔦𝔰𝔠(f+pg)p^{2}|\mathfrak{disc}(f+p\cdot g) for any choice of gg such that deg(f)deg(g)deg(f)\geq deg(g).

Recall the theorem.

Theorem 30.

p2|𝔡𝔦𝔰𝔠(f)p^{2}|\mathfrak{disc}(f) if and only if one of the following holds:

  • ff is strongly divisible by pp if and only if one of the following is true.

    1. (1)

      fmodpf\bmod p has a triple root in (𝔽p)\mathbb{P}(\mathbb{F}_{p})

    2. (2)

      fmodpf\bmod p has two double roots in (𝔽¯p).\mathbb{P}(\overline{\mathbb{F}}_{p}).

  • If ff is not strongly divisible by pp, then one of the following holds

    1. (1)

      ff is weakly divisible by p.p. This is the case where the linear double root in (𝔽p)\mathbb{P}(\mathbb{F}_{p}) is in 𝔽p\mathbb{F}_{p} (seen as a base affine component) and not at the point at infinity.

    2. (2)

      The lead coefficient is divisible by p2p^{2} and the second lead is divisible by pp. In other words, the palindromic reverse of ff is weakly divisible by pp at 0.0.This is the case where the linear double root in (𝔽p)\mathbb{P}(\mathbb{F}_{p}) is at the point at infinity. (We will try and ignore this case by making sure that the lead coefficient of our polynomial is squarefree or making sure that the leading two coefficients are co-prime)

Proof.
  • We start by showing that if

    • fmodpf\bmod p has a triple root in (𝔽p)\mathbb{P}(\mathbb{F}_{p}) or

    • fmodpf\bmod p has two double roots in (𝔽¯p)\mathbb{P}(\overline{\mathbb{F}}_{p})

    then p2|𝔡𝔦𝔰𝔠(f).p^{2}|\mathfrak{disc}(f). Note that this automatically implies that ff is strongly divisible by p.p.

    Since, we just proved the Dedekind-Kummer Theorem for Binary Rings (theorem 12), let us use it here.

    If f(x,y)f(x,y) shows the following decomposition/factorization modulo p,p,

    f(x,y)i(fi)eimodp,f(x,y)\equiv\prod_{i}(f_{i})^{e_{i}}\bmod p,

    then we know (fi)ei(f_{i})^{e_{i}} corresponds to either a DVR or it corresponds to a proper subring in associated DVR.

    Now if some (fi)ei(f_{i})^{e_{i}} corresponds to a proper subring of a DVR, then pp will divide the index of corresponding local ring in DVR. This will imply p2|𝔡𝔦𝔰𝔠(f).p^{2}|\mathfrak{disc}(f).

    On the other hand, if it does not correspond to an irreducible order, then RfR_{f} localized at that prime is a DVR, which corresponds to a degree eie_{i} totally ramified extension of the totally un-ramified extension of degree fif_{i} of the local field p.\mathbb{Q}_{p}. This means that pei1p^{e_{i}-1} divided the discriminant of the corresponding local field extension and thus piei1p^{\sum_{i}e_{i}-1} divides the discriminant of 𝕂f\mathbb{K}_{f}. In this case, however, the power of pp dividing the discriminant of 𝕂f\mathbb{K}_{f} and ff is the same.

    Thus, p2|𝔡𝔦𝔰𝔠(f+pg)p^{2}|\mathfrak{disc}(f+pg) for any choice of gg provided that for some i,i, deg(fi)2&ei2.deg(f_{i})\geq 2\And e_{i}\geq 2.

    Similarly, p2|𝔡𝔦𝔰𝔠(f+pg)p^{2}|\mathfrak{disc}(f+pg) for any choice of gg provided that for some i,i, deg(fi)1&ei3.deg(f_{i})\geq 1\And e_{i}\geq 3.

  • If p2|𝔡𝔦𝔰𝔠(f),p^{2}|\mathfrak{disc}(f), then p|𝔡𝔦𝔰𝔠f.p|\mathfrak{disc}{f}. Thus, ff has a double root in (𝔽¯p).\mathbb{P}(\overline{\mathbb{F}}_{p}). If this root is non linear then we fall in the case discussed above which implies ff is strongly divisible by p.p. Similarly, if the root is linear and the multiplicity is greater than 33 we again fall in the above case.

    Thus, since ff is not strongly divisible by pp, we may assume that ff has only one double root in 1(𝔽¯p).\mathbb{P}^{1}(\bar{\mathbb{F}}_{p}). Now if this linear double root is in 𝔽p,\mathbb{F}_{p}, we will translate appropriately so that the double root is at zero. If the linear root is a point at infinity we take the palindromic inverse or the reciprocal polynomial to shift its double root to 0.0. Then, showing initial polynomial is weakly divisible by pp as in definition 32 at the root will be same as showing new ff is weakly divisible by pp at 0.0.

    We note that, using lemma 15, the discriminant of ff can be written as

    𝔡𝔦𝔰𝔠(f)=2f(0)(f(2)(0)2!)3Δ1+f(0)2Δ2+f(0)f(0)Δ3+f(0)2Δ4\mathfrak{disc}(f)=2f(0)(\frac{f^{(2)}(0)}{2!})^{3}\cdot\Delta_{1}+f^{\prime}(0)^{2}\cdot\Delta_{2}+f(0)f^{\prime}(0)\cdot\Delta_{3}+f(0)^{2}\Delta_{4}

    where

    Δ1=𝔡𝔦𝔰𝔠(ff(0)Ynf(0)XYn1XY).\Delta_{1}=\mathfrak{disc}(\frac{f-f(0)Y^{n}-f^{\prime}(0)XY^{n-1}}{XY}).

    Since, ff is not strongly divisible by p,p, we see that pΔ1p\nmid\Delta_{1} (no two double roots) and pf(2)(0)2!p\nmid\frac{f^{(2)}(0)}{2!} (no triple linear root).

    Since, p|f(0)p|f(0) and p|f(0)p|f^{\prime}(0) and p2|𝔡𝔦𝔰𝔠fp^{2}|\mathfrak{disc}{f} and pΔ1p\nmid\Delta_{1} and pf(2)(0)2!p\nmid\frac{f^{(2)}(0)}{2!} it follows that p2|f(0).p^{2}|f(0).

See appendix.

Definition 21.

We say a ff ultra weakly divisible if

p2|𝔡𝔦𝔰𝔠(f)f is weakly divisible by p.p^{2}|\mathfrak{disc}(f)\Rightarrow f\textit{ is weakly divisible by }p.

If we treat 𝔡𝔦𝔰𝔠(f)\mathfrak{disc}(f) as a function of its coefficients (say aia_{i} by abuse of notation), then this condition corresponds to p2|𝔡𝔦𝔰𝔠(f)p^{2}|\mathfrak{disc}(f) and p|(𝔡𝔦𝔰𝔠)ai(f)p|\frac{\partial(\mathfrak{disc})}{\partial a_{i}}(f) for all 0in0\leq i\leq n (fmodpf\bmod p is in the singular locus of 𝔡𝔦𝔰𝔠(f)=0modp\mathfrak{disc}(f)=0\bmod p).

Remark 31.

While it may seem that the singular locus has a very large co-dimension, it has a component of co-dimension 2.

By definition, if ff is ultra-weakly divisible, then ff does not have non linear double root or a linear triple root modulo any prime p.p.

Theorem 32.

If ff is ultra-weakly divisible, then there exist a maximal mfm_{f} such that ff is weakly divisible by mfm_{f} (at some lfl_{f}) and R(f,mf,lf)R^{\prime}_{(f,m_{f},l_{f})} is maximal i.e. R(f,mf,lf)R^{\prime}_{(f,m_{f},l_{f})} is THE ring of integers for 𝕂f[x]/(f).\mathbb{K}_{f}\simeq\mathbb{Q}[x]/(f). In fact, if 𝔡𝔦𝔰𝔠(f)=st2\mathfrak{disc}(f)=st^{2} where ss is squarefree, then mf=t.m_{f}=t.

Proof.

Let pp denote a prime such that p2|𝔡𝔦𝔰𝔠(f).p^{2}|\mathfrak{disc}(f). And let 0l0<p0\leq l_{0}<p denote the linear double root in 𝔽p\mathbb{F}_{p} of f.f. Then, we note that if

f(x+l)=f(l)+f(l)x+f(2)(l)2!x2++a0xnf(x+l)=f(l)+f^{\prime}(l)\cdot x+\frac{f^{(2)}(l)}{2!}\cdot x^{2}+\cdots+a_{0}x^{n}

then using lemma 15 we see that,

𝔡𝔦𝔰𝔠(f)=f(l)Δ1+f(l)2Δ2+f(l)f(l)Δ3.\mathfrak{disc}(f)=f(l)\cdot\Delta_{1}+f^{\prime}(l)^{2}\cdot\Delta_{2}+f(l)f^{\prime}(l)\cdot\Delta_{3}.

Furthermore, if ll0modpl\equiv l_{0}\bmod p then we can realize Δ1\Delta_{1} as

Δ14(f(2)(l)2!)3𝔡𝔦𝔰𝔠(f)modp\Delta_{1}\equiv 4(\frac{f^{(2)}(l)}{2!})^{3}\mathfrak{disc}(f^{*})\bmod p

where f=f(x)¯(xl0)2¯𝔽p[x]f^{*}=\frac{\overline{f(x)}}{\overline{(x-l_{0})^{2}}}\in\mathbb{F}_{p}[x] .

Since ff is ultra-weakly divisible by p,p, we have ff is not strongly divisible by p.p. This means that if ll0modpl\equiv l_{0}\bmod p then pf2(l)2p\nmid\frac{f^{2}(l)}{2} as this will force ff to not have triple linear root at l0l_{0} in 𝔽p.\mathbb{F}_{p}. We can further say that that pΔ1.p\nmid\Delta_{1}. If pΔ1p\mid\Delta_{1} then it is easy to see that ff will have at least two double roots in 𝔽¯p.\bar{\mathbb{F}}_{p}.

Now, by Hensels lemma for any kk we can find lkl_{k} such that pk|f(lk).p^{k}|f^{\prime}(l_{k}).

Let D=vp(𝔡𝔦𝔰𝔠(f))D=v_{p}(\mathfrak{disc}(f)) and llkmodpkl\equiv l_{k}\bmod p^{k} with 0l<pk0\leq l<p^{k} where k=[D2].k=[\frac{D}{2}].

Then, we note that pr|f(l)p^{r}|f(l) and r<2[D2].r<2[\frac{D}{2}]. It follows that 𝔡𝔦𝔰𝔠(f)prΔ1modpr+[D/2]\mathfrak{disc}(f)\equiv p^{r}\cdot\Delta_{1}\bmod p^{r+[D/2]} which contradicts the fact that pD||𝔡𝔦𝔰𝔠(f).p^{D}||\mathfrak{disc}(f).

Thus, p2[D/2]|f(l)p^{2\cdot[D/2]}|f(l) and p[D/2]|f(l)p^{[D/2]}|f^{\prime}(l) or ff is weakly divisible by p[D/2].p^{[D/2]}. Thus, ff is weakly divisible by mfm_{f} where 𝔡𝔦𝔰𝔠(f)=s(mf)2\mathfrak{disc}(f)=s(m_{f})^{2} and ss is squarefree. This tells us that R(f,mf,lf)R^{\prime}_{(f,m_{f},l_{f})} is pD2[D/2]p^{D-2\cdot[D/2]} which is at most one. This makes R(f,mf,lf)R^{\prime}_{(f,m_{f},l_{f})} the ring of integers of 𝕂f\mathbb{K}_{f} as its discriminant is squarefree. ∎

7. Polynomials that correspond to distinct rings


Definition 22.
W(s:t):={i=0naixni[x]:1a1<s/2<a0s,(a0,a1)=1,|ai|sti for all 2in}W(s:t):=\Big{\{}\sum_{i=0}^{n}a_{i}x^{n-i}\in\mathbb{Z}[x]:1\leq a_{1}<s/2<a_{0}\leq s,(a_{0},a_{1})=1,|a_{i}|\leq st^{i}\text{ for all }2\leq i\leq n\Big{\}}
Remark 33.

The above set is a subset of a fundamental domain for action of integers by translation on polynomials. That is, for any polynomial ff there exists at most one integer ll such that f(x+l)W(s:t).f(x+l)\in W(s:t).

Definition 23.
W(s:t:m):={(f,l):fW(s:t),0l<m,f is weakly divisible by m at l}W(s:t:m):=\{(f,l):f\in W(s:t),0\leq l<m,f\textit{ is weakly divisible by }m\textit{ at }l\}

with the understanding that each element (f,l)W(s:t:m)(f,l)\in W(s:t:m) where ff is Reduced-mm-Polynomial will give distinct rings, see Theorem theorem 50.

Lemma 8.

Given a0,a1,,an2,l,ma_{0},a_{1},\cdots,a_{n-2},l,m there is a unique choice of an1a_{n-1} and ana_{n} satisfying

  • B+1an1B+mB+1\leq a_{n-1}\leq B+m
  • C+1an1C+m2C+1\leq a_{n-1}\leq C+m^{2}
  • a0xn+a1xn1++an is weakly divisible by m at l.a_{0}x^{n}+a_{1}x^{n-1}+\cdots+a_{n}\textit{ is weakly divisible by $m$ at $l.$}
Proof.

Follows from the definition directly. ∎

Discussion 24.

Let BϵB_{\epsilon} denote a compact set in

{f[x]:H(f)1,𝔡𝔦𝔰𝔠(f)0}\{f\in\mathbb{R}[x]:H(f)\leq 1,\mathfrak{disc}_{\mathbb{Z}}(f)\neq 0\}

such that

Vol(Bϵ)(1ϵ)Vol({f:H(f)1}).Vol(B_{\epsilon})\geq(1-\epsilon)Vol(\{f:H(f)\leq 1\}).

Then for this region we can construct a ρB\rho_{B} such that all polynomials ff with height ρB\geq\rho_{B} satisfying

H(f)nf(XH(f))BϵH(f)^{n}f(\frac{X}{H(f)})\in B_{\epsilon}

are Reduced-11-Polynomials.

From theorem 50, we can say that polynomials of this type with height

m1/(n2)ρB\geq m^{1/(n-2)}\rho_{B}

are Reduced-mm-Polynomials.

Furthermore, we may consider BϵB_{\epsilon} to be a finite union of disjoint boxes BiB_{i} (depending on epsilon).

Thus, correspondingly, if di,rd_{i,r} are the dimensions of the box Bi,B_{i}, we define W(s:t:m)BiW(s:t:m)_{B_{i}} to be those polynomials that satisfy

H(f)nsf(XH(f))Bi\frac{H(f)^{n}}{s}f(\frac{X}{H(f)})\in B_{i}

and

f(2)(0)2!sρBn2m.\frac{f^{(2)}(0)}{2!}\geq s\rho_{B}^{n-2}m.

The latter condition an2sρB(n2)ma_{n-2}\geq s\rho_{B}^{(n-2)}m ensures that polynomials we count are of ‘large enough’ (see lemma 14) height to be a Reduced-mm-Polynomial.

Definition 25.
W(s:t:m)Red:={(f,l)W(s,t,m):f is a Reduced-m-Polynomial.}W(s:t:m)^{Red}:=\{(f,l)\in W(s,t,m):f\textit{ is a Reduced-}m\textit{-Polynomial.}\}

7.1. Ekedahlian Sieve- Quantitative version

Let VnV_{n} denote the variety in 𝔸n+1\mathbb{A}^{n+1} which describes the Singular locus of the curve given by Gn=0G_{n}=0 where

Gn:=𝔡𝔦𝔰𝔠X(X0Xn+X1Xn1+Xn).G_{n}:=\mathfrak{disc}_{X}(X_{0}X^{n}+X_{1}X^{n-1}+\cdots X_{n}). (15)

We describe the structure of the discriminant polynomial as a polynomial of Xn.X_{n}.

Lemma 9.
𝔡𝔦𝔰𝔠X(X0Xn+X1Xn1\displaystyle\mathfrak{disc}_{X}(X_{0}X^{n}+X_{1}X^{n-1} ++Xn)=\displaystyle+\cdots+X_{n})=
nn(X0Xn)n1\displaystyle n^{n}(X_{0}X_{n})^{n-1}
+i=1n2XniΔi\displaystyle+\sum_{i=1}^{n-2}X_{n}^{i}\Delta_{i}
+Xn12𝔡𝔦𝔰𝔠X(X0Xn1+X1Xn2++Xn1)\displaystyle+X_{n-1}^{2}\mathfrak{disc}_{X}(X_{0}X^{n-1}+X_{1}X^{n-2}+\cdots+X_{n-1})
Proof.

Obvious from the the understanding that X0𝔡𝔦𝔰𝔠X(X0Xn+X1Xn1++Xn)X_{0}\mathfrak{disc}_{X}(X_{0}X^{n}+X_{1}X^{n-1}+\cdots+X_{n}) is the resultant of X0Xn+X1Xn1++XnX_{0}X^{n}+X_{1}X^{n-1}+\cdots+X_{n} and X(X0Xn+X1Xn1++Xn)\frac{\partial}{\partial X}(X_{0}X^{n}+X_{1}X^{n-1}+\cdots+X_{n}) can the consideration of this resultant as the determinant of the Sylvester matrix. See eq. 28 and lemma 15. ∎

Let

Fn:=𝔡𝔦𝔰𝔠Xn(𝔡𝔦𝔰𝔠X(X0Xn+X1Xn1++Xn))F_{n}:=\mathfrak{disc}_{X_{n}}(\mathfrak{disc}_{X}(X_{0}X^{n}+X_{1}X^{n-1}+\cdots+X_{n})) (16)

Thus, for any prime p,p, and integer tuple (a0,a1,,an)(a_{0},a_{1},\cdots,a_{n}),

(a0,a1,,an)Vn(𝔽p)p|(Gn(a0,a1,,an),Fn(a0,a1,,an1)).(a_{0},a_{1},\cdots,a_{n})\in V_{n}(\mathbb{F}_{p})\Rightarrow p|(G_{n}(a_{0},a_{1},\cdots,a_{n}),F_{n}(a_{0},a_{1},\cdots,a_{n-1)}). (17)

This follows from the fact that Gn,FnI(Vn).G_{n},F_{n}\in I(V_{n}).

We note that

Remark 34.

A polynomial ff is UWD or ultra-weakly divisible if and only if

p:fmodpVn(𝔽p).\forall p\in\mathbb{P}:f\bmod p\notin V_{n}(\mathbb{F}_{p}).

We will now look for UWD polynomials in W(s:t:m).W(s:t:m). We will then look for Reduced-mm-polynomials in this set which will allow us to count UWD Rings. We note that UWD polynomials are easily sieve-able via the Ekedahl sieve.

Discussion 26.

Given f=i=0naixnif=\sum_{i=0}^{n}a_{i}x^{n-i} such that f(x)f(x) is weakly divisible by mm at 0 and (a0,a1)=1(a_{0},a_{1})=1,

𝔡𝔦𝔰𝔠(f)m22(an2)3𝔡𝔦𝔰𝔠(f)modm\frac{\mathfrak{disc}(f)}{m^{2}}\equiv 2(a_{n-2})^{3}\mathfrak{disc}(f^{*})\bmod m

where f=i=0n2aixni2.f^{*}=\sum_{i=0}^{n-2}a_{i}x^{n-i-2}. This follows from the structure of the discriminant polynomial in lemma 15.

Thus, the only way ff can be strongly divisible by pp for some p|mp|m if

  • either p|an2p|a_{n-2}(this will force a triple root at 0)

  • or p|𝔡𝔦𝔰𝔠(f)p|\mathfrak{disc}(f^{*})(this will force some other double root in 𝔽¯p\bar{\mathbb{F}}_{p}).

Furthermore, applying the structure of the discriminant (lemma 15) to ff^{*} we can say that for every a0,a1,,an3a_{0},a_{1},\cdots,a_{n-3} and p|mp|m there are at most n+1n+1 solutions modp\bmod p for an2a_{n-2} such that

2(an2)3𝔡𝔦𝔰𝔠(i=0n2aixni2)0modp.2(a_{n-2})^{3}\mathfrak{disc}(\sum_{i=0}^{n-2}a_{i}x^{n-i-2})\equiv 0\bmod p.

We will use the traditional form of the Ekedahl sieve. This will serve us better. Our parameter of mm which interferes with the sieving in the main term. And since we wish to maximize mm to be used, we wish to use the fact that an2a_{n-2} freely moves in a much larger range than other coefficients and based on the structure of GnG_{n} can be used to give a usable lower bound for fVn(𝔽p)f\notin V_{n}(\mathbb{F}_{p}) for p|mp|m for much larger values of m.m.

Theorem 35.

If sstnm2,s\gg\frac{st^{n}}{m^{2}}, then

|{(f,l)W(s:t:m):p,p>M,pm,fVn(𝔽p)}||\{(f,l)\in W(s:t:m):\exists p\in\mathbb{P}_{\mathbb{Z}},p>M,p\nmid m,f\in V_{n}(\mathbb{F}_{p})\}|
=O(sntn(n+1)21m2)(1MlogM+m2log(stn)stn)=O(\frac{s^{n}t^{\frac{n(n+1)}{2}-1}}{m^{2}})\cdot(\frac{1}{M\log M}+\frac{m^{2}\log(st^{n})}{st^{n}})
Proof.

As usual we put this set into the union of 3 sets:

  • S1:={(f,l)W(s:t:m):Fn(f)=0}S_{1}:=\{(f,l)\in W(s:t:m):F_{n}(f)=0\}

    For any choice of (a0,a1,,an2,l)(a_{0},a_{1},\cdots,a_{n-2},l) such that FnF_{n} is non-degenerate as a polynomial in an1,a_{n-1}, then we can have at most n(n1)n(n-1) possible choices for an1.a_{n-1}.

    Inductively, the number of choices for (a0,a1,,an2)(a_{0},a_{1},\cdots,a_{n-2}) for which the polynomial FnF_{n} is degenerate as a polynomial in an1a_{n-1} would be space cut out by all coefficient polynomials in 𝔸n1.\mathbb{A}^{n-1}. Thus, the number of elements in S1S_{1} is

    On(sntn(n+1)21m2)(1s+mstn1).O_{n}(\frac{s^{n}t^{\frac{n(n+1)}{2}-1}}{m^{2}})\cdot(\frac{1}{s}+\frac{m}{st^{n-1}}).
  • S2:={(f,l)W(s:t:m):p,stnm2>p>M,pm,fVn(𝔽p)}S_{2}:=\{(f,l)\in W(s:t:m):\exists p\in\mathbb{P}_{\mathbb{Z}},\frac{st^{n}}{m^{2}}>p>M,p\nmid m,f\in V_{n}(\mathbb{F}_{p})\}

    For every such prime pp the number of solutions (modp\bmod p) to

    Gn(f)Fn(f)0modpG_{n}(f)\equiv F_{n}(f)\equiv 0\bmod p

    is O(pn1)O(p^{n-1}). Thus, the total number of solutions possible fW(s:t:m)f\in W(s:t:m) is

    O(pn1)m(sp+O(1))2(2st2p+O(1))(2stkp+O(1)(2stn2p+O(1))\displaystyle O(p^{n-1})\cdot m\cdot(\frac{s}{p}+O(1))^{2}\cdot(\frac{2st^{2}}{p}+O(1))\cdots(\frac{2st^{k}}{p}+O(1)\cdots(\frac{2st^{n-2}}{p}+O(1))
    (2stn1pm+O(1))(2stnpm2+O(1))\displaystyle\cdot(\frac{2st^{n-1}}{pm}+O(1))\cdot(\frac{2st^{n}}{pm^{2}}+O(1))
    =O(sn+1tn(n+1)21m2)(1p2).\displaystyle=O(\frac{s^{n+1}t^{\frac{n(n+1)}{2}-1}}{m^{2}})(\frac{1}{p^{2}}).

    Summing this value over the given range for primes, we get that |S2||S_{2}| is

    O(sn+1tn(n+1)21m2MlogM)O(\frac{s^{n+1}t^{\frac{n(n+1)}{2}-1}}{m^{2}M\log M})
  • S3:={(f,l)W(s:t:m):Fn(f)0,p,stnm2<p,pm,fVn(𝔽p)}S_{3}:=\{(f,l)\in W(s:t:m):F_{n}(f)\neq 0,\exists p\in\mathbb{P}_{\mathbb{Z}},\frac{st^{n}}{m^{2}}<p,p\nmid m,f\in V_{n}(\mathbb{F}_{p})\}

    In this case, we first pick any a0,a1,,an1a_{0},a_{1},\cdots,a_{n-1}. FnF_{n} is not zero. We then pick a prime pp such that p|Fn(a0,a1,,an1)p|F_{n}(a_{0},a_{1},\cdots,a_{n-1}) and p>stnm2p>\frac{st^{n}}{m^{2}}. The size of FnF_{n} is less than (stn)kn(st^{n})^{k_{n}} for some knk_{n}. Thus, the number of distinct prime divisors of FnF_{n} is

    knlog(stn).\ll k_{n}\log(st^{n}).

    Since, if pa0,p\nmid a_{0}, then we can deduce from lemma 9 that GnG_{n} is a polynomial in ana_{n} of degree n1.n-1. If p|a0,p|a_{0}, then pa1p\nmid a_{1} (as (a0,a1)=1(a_{0},a_{1})=1) and thus

    Gn(a0,,an)a12Gn1(a1,a2,,an)modp.G_{n}(a_{0},\cdots,a_{n})\equiv a_{1}^{2}G_{n-1}(a_{1},a_{2},\cdots,a_{n})\bmod p.

    This follows directly from applying lemma 9 to the palindromic inverse of ff. We basically do not have to worry about GnmodpG_{n}\bmod p being degenerate as a polynomial in ana_{n}. Thus, there will be at most n1n-1 possibilities for ana_{n} as the range it moves through is of size less than p.p. Therefore, |S3||S_{3}| will be

    O(sn+1tn(n+1)21m2)m2stnlog(stn).O(\frac{s^{n+1}t^{\frac{n(n+1)}{2}-1}}{m^{2}})\cdot\frac{m^{2}}{st^{n}}\cdot\log(st^{n}).

Definition 27.

Let WnW_{n} denote the variety which is union of VnV_{n} with the variety given by X0=X1=0.X_{0}=X_{1}=0. Let

cp:=1|Wn(𝔽p)|pn+1.c_{p}:=1-\frac{|W_{n}(\mathbb{F}_{p})|}{p^{n+1}}.

Since, for every pp we have polynomials with discriminant co-prime to p,p, cpc_{p} is never zero. Furthermore, since VnV_{n} is a variety contained in V(Fn,Gn)V(F_{n},G_{n}) of co-dimension 22 and X0=X1=0X_{0}=X_{1}=0 is also co-dimension 2 we can conclude that cp(1rnp2)c_{p}\geq(1-\frac{r_{n}}{p^{2}}) for some constant rn.r_{n}. For pm,p\nmid m, cpc_{p} will serve as the measure of how many polynomials mod p are not strongly divisible. The addition of this second variety will save us some trouble as the condition (a0,a1)=1(a_{0},a_{1})=1 will be taken care of internally.

7.2. Actual sieve

Definition 28.
W(s:t:m)={(f,l)W(s,t,m):p|m,fVn(𝔽p)}W^{*}(s:t:m)=\{(f,l)\in W(s,t,m):\forall p|m,f\notin V_{n}(\mathbb{F}_{p})\}

Before moving forward, we refer to 24. Let BiB_{i} denote boxes as there, and let

Definition 29.
W(s:t:m)Bi:=W(s:t:m)BiW(s:t:m)W^{*}(s:t:m)_{B_{i}}:=W(s:t:m)_{B_{i}}\cap W^{*}(s:t:m)

and

W(s:t:m)BiUWD:={(f,l)W(s:t:m)Bi:p,fVn(𝔽p)}W^{*}(s:t:m)_{B_{i}}^{UWD}:=\{(f,l)\in W^{*}(s:t:m)_{B_{i}}:\forall p\in\mathbb{P},f\notin V_{n}(\mathbb{F}_{p})\}

and

W(s:t:m:M)Bi:={(f,l)W(s:t:m)Bi:p<M,fVn(𝔽p)}.W^{*}(s:t:m:M)_{B_{i}}:=\{(f,l)\in W(s:t:m)^{*}_{B_{i}}:\forall p<M,f\notin V_{n}(\mathbb{F}_{p})\}.

We begin with defining M=p<Mpmp.M^{\prime}=\prod_{\begin{subarray}{c}p<M\\ p\nmid m\end{subarray}}p. We will search for solutions to p<M,fVn(𝔽p)\forall p<M,f\notin V_{n}(\mathbb{F}_{p}) by looking at the system modulo MM^{\prime}. So we let

CM=p<Mpmcp.C_{M^{\prime}}=\prod_{\begin{subarray}{c}p<M\\ p\nmid m\end{subarray}}c_{p}.

We will restrict ourselves to squarefree mm. For p|m,p|m, we will look at an2a_{n-2} only and note that for any choice of a0,a1,,an3a_{0},a_{1},\cdots,a_{n-3} there are at most nω(m)n^{\omega(m)} possible solutions modulo mm for an2a_{n-2} for ff to be strongly divisible by m.m. See 26.

Thus, making boxes of size M,M^{\prime}, we see that

|\displaystyle| W(s:t:m:M)Bi|\displaystyle W^{*}(s:t:m:M)_{B_{i}}|
=CMm(s2M+O(1))2(2st2M+O(1))(2stkM+O(1)(2stn2M+O(nω(m))+O(sρBm))\displaystyle=C_{M}\cdot m\cdot(\frac{s}{2M^{\prime}}+O(1))^{2}\cdot(\frac{2st^{2}}{M^{\prime}}+O(1))\cdots(\frac{2st^{k}}{M^{\prime}}+O(1)\cdots(\frac{2st^{n-2}}{M^{\prime}}+O(n^{\omega(m)})+O(s\rho_{B}m))
(2stn1Mm+O(1))(2stnMm2+O(1))(M)n+1.\displaystyle\cdot(\frac{2st^{n-1}}{M^{\prime}m}+O(1))\cdot(\frac{2st^{n}}{M^{\prime}m^{2}}+O(1))\cdot(M^{\prime})^{n+1}.
If we have Mmin{s,stnm2,tn2m}, we get,\displaystyle\textit{ If we have }M^{\prime}\leq\min\{s,\frac{st^{n}}{m^{2}},\frac{t^{n-2}}{m}\},\textit{ we get,}
=2n3Vol(Bi)CMsn+1tn(n+1)21m2(1+O(m2Mstn)+O(Mmtn2)+O(Ms)).\displaystyle=2^{n-3}Vol(B_{i})C_{M}\cdot\frac{s^{n+1}t^{\frac{n(n+1)}{2}-1}}{m^{2}}(1+O(\frac{m^{2}M^{\prime}}{st^{n}})+O(\frac{M^{\prime}m}{t^{n-2}})+O(\frac{M^{\prime}}{s})).

This will not be a problem as we will be taking MM^{\prime} to be small.

We quickly write |W(s:t:m)BiUWD||W^{*}(s:t:m)_{B_{i}}^{UWD}| as

|W(s:t:m:M)Bi|+O(|{(f,l)W(s:t:m):p,p>M,pm,fVn(𝔽p)}|)|W^{*}(s:t:m:M)_{B_{i}}|+O(|\{(f,l)\in W(s:t:m):\exists p\in\mathbb{P}_{\mathbb{Z}},p>M,p\nmid m,f\in V_{n}(\mathbb{F}_{p})\}|)

Let C=limMCM.C=\lim_{M\longrightarrow\infty}C_{M}. Using theorem 35, we get

Theorem 36.

If sstnm2,s\gg\frac{st^{n}}{m^{2}}, then

|W(s:t:m)BiUWD|\displaystyle|W^{*}(s:t:m)_{B_{i}}^{UWD}|\gg 2n3Vol(Bi)Csn+1tn(n+1)21m2(1+O(m2eMstn)+O(meMtn2))\displaystyle 2^{n-3}Vol(B_{i})C\cdot\frac{s^{n+1}t^{\frac{n(n+1)}{2}-1}}{m^{2}}(1+O(\frac{m^{2}e^{M}}{st^{n}})+O(\frac{me^{M}}{t^{n-2}}))
+O(sntn(n+1)21m2)(1MlogM+m2log(stn)stn)\displaystyle+O(\frac{s^{n}t^{\frac{n(n+1)}{2}-1}}{m^{2}})\cdot(\frac{1}{M\log M}+\frac{m^{2}\log(st^{n})}{st^{n}})

Setting eM=(stn)ϵ/2,e^{M}=(st^{n})^{\epsilon/2}, we get the following corollary.

Corollary 2.

If (stn)ϵs(st^{n})^{\epsilon}\ll s then

|W(s:t:m)BiUWD|2n3Vol(Bi)Csn+1tn(n+1)21m2(1+Oϵ(1log(stn))+O(m2(stn)ϵstn)+O(m(stn)ϵtn2)).|W^{*}(s:t:m)_{B_{i}}^{UWD}|\gg 2^{n-3}Vol(B_{i})C\cdot\frac{s^{n+1}t^{\frac{n(n+1)}{2}-1}}{m^{2}}(1+O_{\epsilon}(\frac{1}{\log(st^{n})})+O(\frac{m^{2}(st^{n})^{\epsilon}}{st^{n}})+O(\frac{m(st^{n})^{\epsilon}}{t^{n-2}})).

Summing over all boxes BiB_{i} we get the following corollary.

Corollary 3.

If (stn)ϵs(st^{n})^{\epsilon}\ll s then

|W(s:t:m)|(1ϵ)2n3Csn+1tn(n+1)21m2(1+Oϵ(1log(stn))+O(m2(stn)ϵstn)+O(m(stn)ϵtn2)).|W^{*}(s:t:m)|\gg(1-\epsilon)2^{n-3}C\frac{s^{n+1}t^{\frac{n(n+1)}{2}-1}}{m^{2}}(1+O_{\epsilon}(\frac{1}{\log(st^{n})})+O(\frac{m^{2}(st^{n})^{\epsilon}}{st^{n}})+O(\frac{m(st^{n})^{\epsilon}}{t^{n-2}})).

We set X=s2n2tn(n1)m2X=\frac{s^{2n-2}t^{n(n-1)}}{m^{2}}.

Let S(X,s,m):=W(s:t:m)S(X,s,m):=W^{*}(s:t:m).

Note that this set counts distinct rings.

It follows that

Theorem 37.

For n5,n\geq 5,

S(X,s,m)nX12+1ns2nm12nS(X,s,m)\gg_{n}\frac{X^{\frac{1}{2}+\frac{1}{n}}\cdot s^{\frac{2}{n}}}{m^{1-\frac{2}{n}}} (18)

provided that

  1. (1)

    sϵ(stn)ϵs\gg_{\epsilon}(st^{n})^{\epsilon}

  2. (2)

    X1ϵϵs2n2mn41n2X^{1-\epsilon}\gg_{\epsilon}s^{2n-2}\cdot m^{n-4-\frac{1}{n-2}} (Condition corresponding to tn2ϵm1+ϵt^{n-2}\gg_{\epsilon}m^{1+\epsilon} - required for injectivity).

  3. (3)

    X1ϵϵsn1m2n4X^{1-\epsilon}\gg_{\epsilon}s^{n-1}\cdot m^{2n-4} (Condition corresponding to (stn)1ϵϵm2(st^{n})^{1-\epsilon}\gg_{\epsilon}m^{2}-required so the final coefficient has enough space to very to have polynomials which are weakly divisible by mm)

To count ultra-weakly divisible rings which are weakly divisible by mm rings, we wish to choose ss appropriately. Note that within the constraints for a fixed XX, the rings counted will all be distinct for any choice of mm and then any choice of s.s.

We can chose ss such that stnt2n4.st^{n}\simeq t^{2n-4}. This will make the condition on mm a singular condition. May not be optimal.

stnt2n4s(m2X)n4(n1)(3n8)st^{n}\simeq t^{2n-4}\iff s\simeq(m^{2}X)^{\frac{n-4}{(n-1)(3n-8)}}

Both conditions combine to

X1ϵ(m2X)n43n8m2n4Xn23n213n+12ϵmX^{1-\epsilon}\geq(m^{2}X)^{\frac{n-4}{3n-8}}m^{2n-4}\iff X^{\frac{n-2}{3n^{2}-13n+12}-\epsilon}\geq m

Thus, if we let S(X:m)S(X:m) denote the set of all UWD Rings, weakly divisible by m,m, then we have

S(X:m)X12+1n+2(n4)n(n1)(3n8)m2n+4(n4)n(n1)(3n8)mS(X:m)\gg\frac{X^{\frac{1}{2}+\frac{1}{n}+\frac{2(n-4)}{n(n-1)(3n-8)}}m^{\frac{2}{n}+\frac{4(n-4)}{n(n-1)(3n-8)}}}{m}
T(X):=mXn23n213n+12ϵS(X,m)\displaystyle T(X):=\sum_{m\leq X^{\frac{n-2}{3n^{2}-13n+12}-\epsilon}}S(X,m) mXn23n213n+12ϵX12+1n+2(n4)n(n1)(3n8)m2n+4(n4)n(n1)(3n8)m\displaystyle\gg\sum_{m\leq X^{\frac{n-2}{3n^{2}-13n+12}-\epsilon}}\frac{X^{\frac{1}{2}+\frac{1}{n}+\frac{2(n-4)}{n(n-1)(3n-8)}}m^{\frac{2}{n}+\frac{4(n-4)}{n(n-1)(3n-8)}}}{m}
nX12+1n+2(n4)n(n1)(3n8)(Xn23n213n+12ϵ)2n+4(n4)n(n1)(3n8)\displaystyle\gg_{n}X^{\frac{1}{2}+\frac{1}{n}+\frac{2(n-4)}{n(n-1)(3n-8)}}(X^{\frac{n-2}{3n^{2}-13n+12}-\epsilon})^{\frac{2}{n}+\frac{4(n-4)}{n(n-1)(3n-8)}}
ϵX12+1n43ϵ\displaystyle\sim_{\epsilon}X^{\frac{1}{2}+\frac{1}{n-\frac{4}{3}}-\epsilon}
Remark 38.

We are summing over squarefree mm.

7.3. Number-field Counting Strategy

The strategy now comes into play by noting that

Theorem 39.

A UWD ring is a Restricted Sudo maximal Ring.

Proof.

Using Dedekind Kummer for Binary Rings i.e.theorem 12, we see that factorization of the corresponding UWD polynomial modulo pp captures the irreducible decomposition of the associated binary ring. We see that if ff is Ultra-weakly divisible then at every prime pp it can never have 2 double roots or a triple root. This means that ff can at most have one double root modulo p.p. The corresponding factor will correspond to either a maximal ring at pp or an irreducible order with ef(𝒪)=2.ef(\mathcal{O})=2. Thus, RfR_{f} is restricted Sudo maximal Ring. Recall that Sudo maximal rings were Rings whose theorem 7 decomposition only consist of irreducible orders 𝒪,\mathcal{O}, satisfying ef(𝒪)=2.ef(\mathcal{O})=2. Our description of Restricted Sudo Maximal Rings and Pseudo maximal rings immediately tells us that any intermediate Ring that sits in between a Restricted Sudo maximal Ring and its integral Closure(the ring of integers it sits in) is also a Restricted Sudo maximal Ring. ∎

Let A(X)A(X) denote the set of all Restricted Sudo maximal Rings with discriminant X.\leq X. Let a(X):=|A(X)|a(X):=|A(X)|. Thus, as a corollary we have

Corollary 4.

a(X)T(X).a(X)\geq T(X).

Let

B(X):={frac(R):RA(X)}.B(X):=\{frac(R):R\in A(X)\}.

and b(X):=|B(X)|.b(X):=|B(X)|.

On the other hand, we may bound a(X)a(X) above by using theorem 27 as follows.

a(X)\displaystyle a(X) 𝕂B(X)S𝕂(X)𝕂B(X)XD𝕂log(XD𝕂)cn\displaystyle\leq\sum_{\mathbb{K}\in B(X)}S_{\mathbb{K}}(X)\leq\sum_{\mathbb{K}\in B(X)}\sqrt{\frac{X}{D_{\mathbb{K}}}}\log(\frac{X}{D_{\mathbb{K}}})^{c_{n}}
X𝕂B(X)1D𝕂log(X)cn.\displaystyle\leq\sqrt{X}\sum_{\mathbb{K}\in B(X)}\frac{1}{\sqrt{D_{\mathbb{K}}}}\log(X)^{c_{n}}.

Combine with the fact that

a(X)T(X)X12+1n43ϵa(X)\gg T(X)\gg X^{\frac{1}{2}+\frac{1}{n-\frac{4}{3}}-\epsilon}

We get

N(X:n)𝕂B(X)1D𝕂X1n43ϵN^{*}(X:n)\geq\sum_{\mathbb{K}\in B(X)}\sqrt{\frac{1}{D_{\mathbb{K}}}}\gg X^{\frac{1}{n-\frac{4}{3}}-\epsilon}
Corollary 5.
lim supXN(X:n)X12+1n43ϵ=.\limsup_{X\longrightarrow\infty}\frac{N(X:n)}{X^{\frac{1}{2}+\frac{1}{n-\frac{4}{3}}-\epsilon}}=\infty.
Remark 40.

We note that if the number of UWD polynomial tuples (f,m)(f,m) with discriminant <X<X is of the expected order then we can easily convert this to be an appropriate lower bound for N(X:n).N(X:n).

Appendix A Defining Weakly Divisible Rings


Most of this section is from our previous paper on Weakly Divisible Rings.

We recall the following theorem from [11] defining Binary Rings.

Let

f(X,Y):=anXn+an1Xn1Y++a0Ynf(X,Y):=a_{n}X^{n}+a_{n-1}X^{n-1}Y+\cdots+a_{0}Y^{n}

denote a binary form of degree nn. Let an0a_{n}\neq 0 and suppose that δ\delta denotes the image of XX in the algebra [X]/(f(X,1)).\mathbb{Q}[X]/(f(X,1)).

We recall eq. 3.

Remark 41.

We refer to the basis given for RfR_{f} in eq. 3 as the canonical basis attached to f.f.

Theorem 42.

When ff is integral(i.e. ff is a binary form of degree nn with integer coefficients), RfR_{f} is a ring of rank nn over .\mathbb{Z}.

Definition 30.

We define IfI_{f} as the (fractional) ideal class generated by (1,δ)(1,\delta) over Rf,R_{f}, when ff is integral.

Definition 31.

When ff is integral, RfR_{f} is known as the binary ring associated to the binary form f.f.

These are a few properties of binary rings.

Proposition 1.

Properties of RfR_{f} (when ff is integral):

  1. (1)
    𝔡𝔦𝔰𝔠(Rf)=𝔡𝔦𝔰𝔠(f).\mathfrak{disc}_{\mathbb{Z}}(R_{f})=\mathfrak{disc}(f). (19)
  2. (2)

    If δ\delta is invertible, and ff is primitive, then

    Rf:=[δ][δ1].R_{f}:=\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}]. (20)
  3. (3)

    If ff is primitive, IfI_{f} is invertible in Rf.R_{f}.

  4. (4)

    Both RfR_{f} and IfI_{f} are invariant under the natural GL2()GL_{2}(\mathbb{Z}) action on binary forms of degree n.n. In particular, for δ¯\\delta\in\overline{\mathbb{Q}}\backslash\mathbb{Q} this means that if λ=aδ+bcδ+d\lambda=\frac{a\delta+b}{c\delta+d} with a,b,c,da,b,c,d\in\mathbb{Z} and adbc=±1ad-bc=\pm 1 then

    [δ][δ1]=[λ][λ1].\mathbb{Z}[\delta]\cap\mathbb{Z}[\delta^{-1}]=\mathbb{Z}[\lambda]\cap\mathbb{Z}[\lambda^{-1}].

You can find the proof of all of these statements in [11].

Remark 43.

Note that one can write down the multiplication table for the defining basis of RfR_{f} in terms of coefficients of ff explicitly. Using this table to define binary rings, one can give a definition for RfR_{f} without the need for the condition “an0a_{n}\neq 0”.

We will note the following part of the multiplication table of the canonical basis of RfR_{f} (eq. 3), which is easily verified.

Lemma 10.

Let f(X,Y):=anXn+an1Xn1Y++a0Ynf(X,Y):=a_{n}X^{n}+a_{n-1}X^{n-1}Y+\cdots+a_{0}Y^{n} denote a binary form of degree n,n, and let B0,B1,,Bn1\langle B_{0},B_{1},\cdots,B_{n-1}\rangle denote the canonical basis for RfR_{f} associated to ff as in eq. 3. Then, we have

Bn1Bni=a0Bni1+aiBn1\displaystyle B_{n-1}\cdot B_{n-i}=-a_{0}\cdot B_{n-i-1}+a_{i}\cdot B_{n-1} (21)
Proof.

We make note of the fact, Bn1=a0δ1.B_{n-1}=-a_{0}\delta^{-1}.
Thus,

Bn1Bni=a0δ1(j=0nianjδj)\displaystyle B_{n-1}\cdot B_{n-i}=-a_{0}\delta^{-1}\cdot(\sum_{j=0}^{n-i}a_{n-j}\delta^{j})
=a0j=0nianjδj1\displaystyle=-a_{0}\cdot\sum_{j=0}^{n-i}a_{n-j}\delta^{j-1}
=a0(Bni1+aiδ)\displaystyle=-a_{0}\cdot(B_{n-i-1}+\frac{a_{i}}{\delta})
=a0Bni1+ai(a0δ)\displaystyle=-a_{0}\cdot B_{n-i-1}+a_{i}\cdot(\frac{-a_{0}}{\delta})
=a0Bni1+aiBn1.\displaystyle=-a_{0}\cdot B_{n-i-1}+a_{i}\cdot B_{n-1}.

Definition 32.

We say a binary form f(x,y)[x,y]f(x,y)\in\mathbb{Z}[x,y] of degree nn is weakly divisible by m>0m\in\mathbb{Z}_{>0} if there exists an \ell\in\mathbb{Z} such that

f(l,1)\displaystyle f(l,1) 0modm2\displaystyle\equiv 0\bmod{m^{2}}
fx(l,1)\displaystyle\frac{\partial f}{\partial x}(l,1) 0modm\displaystyle\equiv 0\bmod m

When appropriate we say ff is weakly divisible by mm at ll.

Remark 44.

We use “weakly divisible” as these polynomials are defined in [3] and [2] and the main parts of these papers is about establishing an upper bound for the number of polynomials which are weakly divisible by mm and bounded height, for all large m.m.

Definition 33.

Given a binary form ff we set fl=fl(x,y):=f(x+ly,y).f_{l}=f_{l}(x,y):=f(x+ly,y).

The condition that “ff is weakly divisible by mm at ll” is equivalent to having integers a0,a1,,ana_{0},a_{1},\cdots,a_{n} such that

fl=anxn+an1xn1y++a2x2yn2+ma1xyn1+m2a0yn.f_{l}=a_{n}x^{n}+a_{n-1}x^{n-1}y+\cdots+a_{2}x^{2}y^{n-2}+ma_{1}xy^{n-1}+m^{2}a_{0}y^{n}.
Definition 34.

We define,

R(f,m):=B0,B1,,Bn2,Bn1m.R^{\prime}_{(f,m)}:=\mathbb{Z}\langle B_{0},B_{1},\cdots,B_{n-2},\frac{B_{n-1}}{m}\rangle.
Theorem 45.

If ff is weakly divisible by mm at l,l, then R(fl,m)R^{\prime}_{(f_{l},m)} is a ring.

Proof.

Let B0,B1,,Bn1\langle B_{0},B_{1},\cdots,B_{n-1}\rangle denote the canonical (old) basis of RflR_{f_{l}} associated to fl.f_{l}. Since, ff is weakly divisible by mm at l,l, we can write

fl=anxn+an1xn1y++a2x2yn2+ma1xyn1+m2a0yn,f_{l}=a_{n}x^{n}+a_{n-1}x^{n-1}y+\cdots+a_{2}x^{2}y^{n-2}+ma_{1}xy^{n-1}+m^{2}a_{0}y^{n},

where ai,m.a_{i},m\in\mathbb{Z}.

We will show that R(fl,m)R^{\prime}_{(f_{l},m)} is a ring, by showing that product of any two elements in the (new) basis given by B0,B1,,Bn2,Bn1m\langle B_{0},B_{1},\cdots,B_{n-2},\frac{B_{n-1}}{m}\rangle (basis for R(fl,m))R^{\prime}_{(f_{l},m)}) is in the \mathbb{Z}-span of itself. We will achieve this by comparing the product of every two elements in the old basis with the product of the corresponding two elements in the new basis.)

We note that BiBjB_{i}\cdot B_{j} for i,jn1i,j\neq n-1 can be written as a \mathbb{Z}-linear combination of Bi\langle B_{i}\rangle. This follows directly from the fact that the \mathbb{Z}-span of Bi\langle B_{i}\rangle forms a ring (the prodigal binary ring) (we can also refer to the multiplication tables given in [9] or section 2.1 in [11]).

It immediately follows that BiBjB_{i}\cdot B_{j} for i,jn1i,j\neq n-1 can also be written as a \mathbb{Z}-linear combination of our new basis B0,B1,,Bn2,Bn1m\langle B_{0},B_{1},\cdots,B_{n-2},\frac{B_{n-1}}{m}\rangle (which only differs from the original basis at the index n1n-1) by simply replacing Bn1B_{n-1} with mBn1mm\cdot\frac{B_{n-1}}{m} in the multiplication table of the old basis.

On the other hand, lemma 10 immediately tells us that, when i1i\neq 1

Bn1mBni=ma0Bni1+aiBn1m\displaystyle\frac{B_{n-1}}{m}\cdot B_{n-i}=-ma_{0}\cdot B_{n-i-1}+a_{i}\cdot\frac{B_{n-1}}{m}

and

Bn1mBn1m=a0Bn2+a1Bn1m.\displaystyle\frac{B_{n-1}}{m}\cdot\frac{B_{n-1}}{m}=-a_{0}\cdot B_{n-2}+a_{1}\cdot\frac{B_{n-1}}{m}.

Thus, the products of elements in this new basis are \mathbb{Z}-linear combinations of the same new basis. It follows that R(fl,m)R^{\prime}_{(f_{l},m)} is in-fact a ring. ∎

Definition 35.

We say R(fl,m)R^{\prime}_{(f_{l},m)} is the weakly divisible ring (at ll with respect to mm) associated to ff, when ff is weakly divisible by mm at ll. When appropriate we will also represent this ring as R(f,m,l).R^{\prime}_{(f,m,l)}.

Remark 46.

Every binary ring is a weakly divisible ring at every value with respect to 1.

Remark 47.

We note that weakly divisible rings may also defined by multiplication tables to avoid dependence on a condition like “an0a_{n}\neq 0 or a00a_{0}\neq 0”.

Appendix B Effective injectivity of the map: (f,m,l)R(fl,m)(f,m,l)\longrightarrow R^{\prime}_{(f_{l},m)}

B.1. Matrix of transformation

For convenience sake, we define (na):=0\binom{n}{a}:=0 if n0n\geq 0 and a<0.a<0.

We mention a useful matrix of transformation for base change from the canonical basis of RfR_{f} associated to f(X,Y)f(X,Y) to the canonical basis of RflR_{f_{l}} associated to fl.f_{l}.

Theorem 48.

We let f(X,Y)=A0Xn+A1Xn1Y++AnYnf(X,Y)=A_{0}X^{n}+A_{1}X^{n-1}Y+\cdots+A_{n}Y^{n} and B0,B1,,Bn1\langle B_{0},B_{1},\cdots,B_{n-1}\rangle be the canonical basis of RfR_{f} associated to f(X,Y)f(X,Y) and C0,C1,,Cn1\langle C_{0},C_{1},\cdots,C_{n-1}\rangle be the canonical basis of RflR_{f_{l}} associated to fl,f_{l}, then we have

C0,C1,,Cn1=B0,B1,,Bn1[1(n11)lA0(n12)l2A0(n1k1)lk1A0(n1n1)ln1A001(n21)l(n2k2)l2(n2n2)ln2001(n3k3)lk3(n3n3)ln3000(nkkk)lkk(nknk)lnk000(1kn+1)lkn+1l000(0kn)lkn1]\displaystyle\begin{matrix}\langle C_{0},C_{1},\cdots,C_{n-1}\rangle=\\ \\ \\ \\ \\ \\ \\ \\ \\ \end{matrix}\begin{matrix}\langle B_{0},B_{1},\cdots,B_{n-1}\rangle\\ \\ \\ \\ \\ \\ \\ \\ \\ \end{matrix}\begin{bmatrix}1&\binom{n-1}{1}lA_{0}&\binom{n-1}{2}l^{2}A_{0}&\cdots&\binom{n-1}{k-1}l^{k-1}A_{0}&\cdots&\binom{n-1}{n-1}l^{n-1}A_{0}\\ 0&1&\binom{n-2}{1}l&\cdots&\binom{n-2}{k-2}l^{2}&\cdots&\binom{n-2}{n-2}l^{n-2}\\ 0&0&1&\cdots&\binom{n-3}{k-3}l^{k-3}&\cdots&\binom{n-3}{n-3}l^{n-3}\\ \vdots&\vdots&\vdots&\ddots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&\binom{n-k^{\prime}}{k-k^{\prime}}l^{k-k^{\prime}}&\cdots&\binom{n-k^{\prime}}{n-k^{\prime}}l^{n-k^{\prime}}\\ \vdots&\vdots&\vdots&\ddots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&\binom{1}{k-n+1}l^{k-n+1}&\cdots&l\\ 0&0&0&\cdots&\binom{0}{k-n}l^{k-n}&\cdots&1\end{bmatrix}

B.2. Proof


 Let F(x,y)=i=0naixniyi denote an irreducible integral binary form.Let δ denote the root of f(x):=F(x,1).Let Fl(x,y):=f(x+ly,y)=i=0nbixniyi.Note that δl is the root of fl(x):=Fl(x,1).Let si(x):=j=0i1ajxij=a0xi+a1xi1++ai1x.Let mi(x):=j=0i1bjxij=b0xi+b1xi1++bi1x.\begin{split}&\text{ Let }F(x,y)=\sum_{i=0}^{n}a_{i}x^{n-i}y^{i}\text{ denote an irreducible integral binary form.}\\ &\text{Let }\delta\text{ denote the root of }f(x):=F(x,1).\\ &\text{Let }F_{l}(x,y):=f(x+ly,y)=\sum_{i=0}^{n}b_{i}x^{n-i}y^{i}.\\ &\text{Note that }\delta-l\text{ is the root of }f_{l}(x):=F_{l}(x,1).\\ &\text{Let }s_{i}(x):=\sum_{j=0}^{i-1}a_{j}x^{i-j}=a_{0}x^{i}+a_{1}x^{i-1}+...+a_{i-1}x.\\ &\text{Let }m_{i}(x):=\sum_{j=0}^{i-1}b_{j}x^{i-j}=b_{0}x^{i}+b_{1}x^{i-1}+...+b_{i-1}x.\end{split} (22)

Thus the basis for RFlR_{F_{l}} will be given by

1,m1(δl)+b1,m2(δl)+b2,,mk(δl)+bk,,mn1(δl)+bn1\displaystyle\mathbb{Z}\langle 1,\>m_{1}(\delta-l)+b_{1},\>m_{2}(\delta-l)+b_{2},\>\cdots,\>m_{k}(\delta-l)+b_{k},\>\cdots,\>m_{n-1}(\delta-l)+b_{n-1}\rangle

and that of RfR_{f} will be given by

1,s1(δ)+a1,s2(δ)+a2,,sk(δ)+ak,,sn1(δ)+an1\displaystyle\mathbb{Z}\langle 1,\>s_{1}(\delta)+a_{1},\>s_{2}(\delta)+a_{2},\>\cdots,\>s_{k}(\delta)+a_{k},\>\cdots,\>s_{n-1}(\delta)+a_{n-1}\rangle
Lemma 11.
bk=i=0k(nink)aib_{k}=\sum_{i=0}^{k}{{n-i}\choose{n-k}}a_{i}\\ (23)
Proof.
f(x+l)=i=0nai(x+l)ni=i=0naij=0ni(nij)xjlnijf(x+l)=\sum_{i=0}^{n}a_{i}(x+l)^{n-i}=\sum_{i=0}^{n}a_{i}\sum_{j=0}^{n-i}{n-i\choose j}x^{j}l^{n-i-j}
=i=0nj=0niai(nij)xjlnij=j=0nxji=0njai(nij)lnij=\sum_{i=0}^{n}\sum_{j=0}^{n-i}a_{i}{n-i\choose j}x^{j}l^{n-i-j}=\sum_{j=0}^{n}x^{j}\sum_{i=0}^{n-j}a_{i}{n-i\choose j}l^{n-i-j}

Thus,

bk=i=0kai(nink)lkib_{k}=\sum_{i=0}^{k}a_{i}{n-i\choose n-k}l^{k-i}

Lemma 12.
mk(x)+bk\displaystyle m_{k}(x)+b_{k} =x(mk1(x)+bk1)+bk\displaystyle=x(m_{k-1}(x)+b_{k-1})+b_{k}
sk(x)+ak\displaystyle s_{k}(x)+a_{k} =x(sk1(x)+ak1)+ak\displaystyle=x(s_{k-1}(x)+a_{k-1})+a_{k}
Proof.

By definition. ∎

Proposition 2.
mk(xl)+bk=j=0k1(nk+j1j)(skj(x)+akj)lj+(n1k)lka0m_{k}(x-l)+b_{k}=\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)+a_{k-j})l^{j}+{n-1\choose k}l^{k}a_{0}\\ (24)
Proof.

We proceed by induction on kk,
Base Case.
m1(xl)=b0(xl)+b1=(a0)(xl)+a1+na0lm_{1}(x-l)=b_{0}(x-l)+b_{1}=(a_{0})(x-l)+a_{1}+na_{0}l
=(a0x+a1)+(n11)a0l=(a_{0}x+a_{1})+{n-1\choose 1}a_{0}l
=(n20)(s1(x)+a1)+(n11)a0={n-2\choose 0}(s_{1}(x)+a_{1})+{n-1\choose 1}a_{0}

Induction Hypothesis:

mk(xl)+bk=j=0k1(nk+j1j)(skj(x)+akj)lj+(n1k)lka0m_{k}(x-l)+b_{k}=\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)+a_{k-j})l^{j}+{n-1\choose k}l^{k}a_{0}

We have from lemma 12,

mk+1(xl)=(xl)(mk(xl)+bk).m_{k+1}(x-l)=(x-l)(m_{k}(x-l)+b_{k}).

Substituting in our induction hypothesis, we get

mk+1(xl)\displaystyle m_{k+1}(x-l)
=(xl)(j=0k1(nk+j1j)(skj(x)+akj)lj+(n1k)lka0)\displaystyle=(x-l)(\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)+a_{k-j})l^{j}+{n-1\choose k}l^{k}a_{0})
=j=0k1(nk+j1j)((skj(x)+akj)ljx(skj(x)+akj)lj+1)+(n1k)a0lkx(n1k)lk+1a0\displaystyle=\sum_{j=0}^{k-1}{n-k+j-1\choose j}((s_{k-j}(x)+a_{k-j})l^{j}x-(s_{k-j}(x)+a_{k-j})l^{j+1})+{n-1\choose k}a_{0}l^{k}x-{n-1\choose k}l^{k+1}a_{0}
=(j=0k1(nk+j1j)(skj(x)x+akjx)lj+(n1k)a0xlk)\displaystyle=(\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)x+a_{k-j}x)l^{j}+{n-1\choose k}a_{0}xl^{k})
 (j=0k1(nk+j1j)(skj(x)+akj)lj+1+(n1k)a0lk+1)\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-(\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)+a_{k-j})l^{j+1}+{n-1\choose k}a_{0}l^{k+1})
=(j=0k1(nk+j1j)(skj+1(x))lj+(n1k)s1(x)lk)\displaystyle=(\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j+1}(x))l^{j}+{n-1\choose k}s_{1}(x)l^{k})
 (j=0k1(nk+j1j)(skj(x)+akj)lj+1+(n1k)a0lk+1)\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-(\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)+a_{k-j})l^{j+1}+{n-1\choose k}a_{0}l^{k+1})
=(j=0k(nk+j1j)(skj+1(x))lj(j=0k1(nk+j1j)(skj(x)+akj)lj+1+(n1k)a0lk+1)\displaystyle=(\sum_{j=0}^{k}{n-k+j-1\choose j}(s_{k-j+1}(x))l^{j}-(\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)+a_{k-j})l^{j+1}+{n-1\choose k}a_{0}l^{k+1})
=(j=0k(nk+j1j)(skj+1(x)+akj+1)lj(j=0k1(nk+j1j)(skj(x)+akj)lj+1)\displaystyle=(\sum_{j=0}^{k}{n-k+j-1\choose j}(s_{k-j+1}(x)+a_{k-j+1})l^{j}-(\sum_{j=0}^{k-1}{n-k+j-1\choose j}(s_{k-j}(x)+a_{k-j})l^{j+1})
 (n1k)a0lk+1j=0k1(nk+j1j)akj+1lj\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-{n-1\choose k}a_{0}l^{k+1}-\sum_{j=0}^{k-1}{n-k+j-1\choose j}a_{k-j+1}l^{j}
=j=0k((nk+j1j)(nk+j2j1))(sk+1j+ak+1j)lj)\displaystyle=\sum_{j=0}^{k}({n-k+j-1\choose j}-{n-k+j-2\choose j-1})(s_{k+1-j}+a_{k+1-j})l^{j})
 (n1k)a0lk+1j=0k1(nk+j1j)akj+1lj\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-{n-1\choose k}a_{0}l^{k+1}-\sum_{j=0}^{k-1}{n-k+j-1\choose j}a_{k-j+1}l^{j}
=(j=0k(nk2+jj)(sk+1j+ak+1j)lj)(n1k)a0lk+1j=0k(nk+j1j)akj+1lj\displaystyle=(\sum_{j=0}^{k}{n-k-2+j\choose j}(s_{k+1-j}+a_{k+1-j})l^{j})-{n-1\choose k}a_{0}l^{k+1}-\sum_{j=0}^{k}{n-k+j-1\choose j}a_{k-j+1}l^{j}
=(j=0k(nk2+jj)(sk+1j+ak+1j)lj+(n1k+1)a0lk+1)\displaystyle=(\sum_{j=0}^{k}{n-k-2+j\choose j}(s_{k+1-j}+a_{k+1-j})l^{j}+{n-1\choose k+1}a_{0}l^{k+1})
 (n1k)a0lk+1(j=0k(nk+j1j)akj+1lj+(n1k+1)a0lk+1)\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-{n-1\choose k}a_{0}l^{k+1}-(\sum_{j=0}^{k}{n-k+j-1\choose j}a_{k-j+1}l^{j}+{n-1\choose k+1}a_{0}l^{k+1})
=(j=0k(nk2+jj)(sk+1j+ak+1j)lj+(n1k+1)a0lk+1)\displaystyle=(\sum_{j=0}^{k}{n-k-2+j\choose j}(s_{k+1-j}+a_{k+1-j})l^{j}+{n-1\choose k+1}a_{0}l^{k+1})
 (j=0k(nk+j1j)akj+1lj+((n1k+1)+(n1k))a0lk+1)\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-(\sum_{j=0}^{k}{n-k+j-1\choose j}a_{k-j+1}l^{j}+({n-1\choose k+1}+{n-1\choose k})a_{0}l^{k+1})
=(j=0k(nk2+jj)(sk+1j+ak+1j)lj+(n1k+1)a0lk+1)\displaystyle=(\sum_{j=0}^{k}{n-k-2+j\choose j}(s_{k+1-j}+a_{k+1-j})l^{j}+{n-1\choose k+1}a_{0}l^{k+1})
 (j=0k(nk+j1j)akj+1lj+(nk+1)a0lk+1)\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-(\sum_{j=0}^{k}{n-k+j-1\choose j}a_{k-j+1}l^{j}+{n\choose k+1}a_{0}l^{k+1})
=(j=0k(nk2+jj)(sk+1j+ak+1j)lj+(n1k+1)a0lk+1)\displaystyle=(\sum_{j=0}^{k}{n-k-2+j\choose j}(s_{k+1-j}+a_{k+1-j})l^{j}+{n-1\choose k+1}a_{0}l^{k+1})
 (j=0k(nk+j1nk1)akj+1lj+(nk+1)a0lk+1)\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-(\sum_{j=0}^{k}{n-k+j-1\choose n-k-1}a_{k-j+1}l^{j}+{n\choose k+1}a_{0}l^{k+1})
=(j=0k(nk2+jj)(sk+1j+ak+1j)lj+(n1k+1)a0lk+1)\displaystyle=(\sum_{j=0}^{k}{n-k-2+j\choose j}(s_{k+1-j}+a_{k+1-j})l^{j}+{n-1\choose k+1}a_{0}l^{k+1})
 bk+1\displaystyle\text{\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt}-b_{k+1}

Substituting x=δx=\delta in the above Proposition we get the result.

B.3.

Given ff a binary form, we let B0,B1,,Bn1\langle B_{0},B_{1},\cdots,B_{n-1}\rangle be the canonical basis of RfR_{f} associated to ff as in eq. 3. Let v1,v2,,vn\langle v_{1},v_{2},\cdots,v_{n}\rangle denote a real basis for n.\mathbb{R}^{n}. We perform a Gram-Schmidt reduction on the basis using some canonical distance form. We write

v1,v2,,vn=BM[(t1,t2,,tn)]\langle v_{1},v_{2},\cdots,v_{n}\rangle=\langle B\rangle M[(t_{1},t_{2},...,t_{n})]

where BB is an orthonormal ordered basis of vectors, all of which are of the same size. MM is an upper triangular uni-potent matrix. And [(t1,t2,..,tn)][(t_{1},t_{2},..,t_{n})] denotes a diagonal matrix with tit_{i} placed in the (i,i)th(i,i)^{th} place.

Let viv_{i}^{\prime} denote the projection of viv_{i} to the space orthogonal to the space spanned by {v1,v2,,vi1}\{v_{1},v_{2},\cdots,v_{i-1}\}. Then, Gram-Schmidt process forces

|ti|=vi.|t_{i}|=||v_{i}^{\prime}||. (25)

We note that for each 1in1\leq i\leq n, and any choice of coefficients a1,,aia_{1},\cdots,a_{i} we have

aivi+j=1i1ajvj|ai||ti|.||a_{i}v_{i}+\sum_{j=1}^{i-1}a_{j}v_{j}||\ \geq\ |a_{i}||t_{i}|.

This follows by looking at the component of the given vector (on RHS) along vi.v^{\prime}_{i}. Moreover, for each 1in1\leq i\leq n, there exist b1,,bi1b_{1},\cdots,b_{i-1}\in\mathbb{Z} so that

vi+j=1i1bjvj|ti|+|ti1|++|t1|.||v_{i}+\sum_{j=1}^{i-1}b_{j}v_{j}||\leq|t_{i}|+|t_{i-1}|+\cdots+|t_{1}|.

One can simply choose bjb_{j} such that the j=1i1bjvj\sum_{j=1}^{i-1}b_{j}v_{j} approximates the vector viviv_{i}-v^{\prime}_{i} which clearly lies in the space spanned by v1,v2,,vi1.\langle v_{1},v_{2},\cdots,v_{i-1}\rangle.

Now if ti+1/ti>2t_{i+1}/{t_{i}}>2 for all ii, one can see that inductively the ithi^{th} vector in the Minkowski reduced basis of the lattice spanned by {v1,v2,vn}\{v_{1},v_{2}\cdots,v_{n}\} will have the form ±vi+j=1i1bjvj.\pm v_{i}+\sum_{j=1}^{i-1}b_{j}v_{j}.

In fact, having

ti/ti1>1+1i for all 1int_{i}/{t_{i-1}}>\sqrt{1+\frac{1}{i}}\textit{ for all }1\leq i\leq n

is sufficient to conclude this.

Furthermore, if we just know that,

tit22 for all 3in and t2t12\frac{t_{i}}{t_{2}}\geq 2\textit{ for all }3\leq i\leq n\textit{ and }\frac{t_{2}}{t_{1}}\geq 2 (26)

then, the first two vectors (which will be unique up to sign) in the Minkowski reduced basis for the lattice v1,v2,,vn\langle v_{1},v_{2},\cdots,v_{n}\rangle will have the above form. That is

{±v1,±(v2+av1)}\{\pm v_{1},\pm(v_{2}+a\cdot v_{1})\}

will be the first two elements in the Minkowski reduced basis for this lattice spanned by v1,v2,,vn.\langle v_{1},v_{2},\cdots,v_{n}\rangle.

Definition 36.

If v1,v2,,vn\langle v_{1},v_{2},\cdots,v_{n}\rangle is a basis for n\mathbb{R}^{n} and viv_{i}^{\prime} denotes the projection of viv_{i} to the space orthogonal to the space spanned by {v1,v2,,vi1}\{v_{1},v_{2},\cdots,v_{i-1}\} and ti=vi,t_{i}=||v_{i}^{\prime}||, then we say v1,v2,,vn\langle v_{1},v_{2},\cdots,v_{n}\rangle is Normally Minkowski Reduced if tit_{i} satisfy

tit22 for all 3in and t2t12\frac{t_{i}}{t_{2}}\geq 2\textit{ for all }3\leq i\leq n\textit{ and }\frac{t_{2}}{t_{1}}\geq 2 (27)

Our discussion above gives us the following lemma:

Lemma 13.

If B1,B2,,Bn\langle B_{1},B_{2},\cdots,B_{n}\rangle is Normally Minkowski Reduced, then

  1. (1)

    B1B_{1} is the unique smallest vector in LL up to sign.

  2. (2)

    a\exists a\in\mathbb{Z} such that, B2+aB1B_{2}+a\cdot B_{1} is the smallest vector in LL which is not in B1\mathbb{Z}\cdot B_{1} up to sign.

B.4. Reduced-m-Polynomials

Definition 37.

We say a tuple (f,m,l)(f,m,l) is a Reduced-mm-Polynomial (at ll) if

  • The canonical basis for R(fl,m)R^{\prime}_{(f_{l},m)} is Normally Minkowski Reduced when seen in R(fl,m)rsR^{\prime}_{(f_{l},m)}\otimes\mathbb{R}\simeq\mathbb{R}^{r}\oplus\mathbb{C}^{s} under the canonical norm.

Remark 49.

We will ignore the ll part as (f,m,l)(f,m,l) is a Reduced-mm-polynomial (at ll) \iff (f,m,r)(f,m,r) is a Reduced-mm-Polynomial (at rr) for any real value r.r. This is easy to see from the fact that the matrix in theorem 48 is upper triangular and hence will leave relative sizes of normal components unchanged.

Theorem 50.

Given f,gf,g polynomials of degree n4n\geq 4 such that ff is a Reduced-mm-polynomial and ff is weakly divisible by mm at ee and gg is a Reduced-mm^{\prime}-polynomial and gg is weakly divisible by mm^{\prime} at dd (m,m1m,m^{\prime}\geq 1) with R(fe,m)=R(gd,m)R^{\prime}_{(f_{e},m)}=R^{\prime}_{(g_{d},m)} then

m=m and r:g(x)=f(x+mrd+e).m=m^{\prime}\textit{ and }\exists r\in\mathbb{Z}:g(x)=f(x+mr-d+e).
Proof.

Let β\beta and α\alpha denote roots of ff and gg respectively (rs\in\mathbb{R}^{r}\oplus\mathbb{C}^{s}). Let a0a_{0} and b0b_{0} denote the positive leading coefficients of ff and gg and a1a_{1} and b1b_{1} denote the second leading coefficient of ff and gg respectively.

We know that the smallest two elements of any Normally Minkowski Reduced basis are unique. Since R(fe,m)=R(gd,m),R^{\prime}_{(f_{e},m)}=R^{\prime}_{(g_{d},m^{\prime})}, comparing the smallest two elements elements in these, we get

1,a0β[ab0c]=1,b0α\langle 1,a_{0}\cdot\beta\rangle\begin{bmatrix}a&b\\ 0&c\end{bmatrix}=\langle 1,b_{0}\cdot\alpha\rangle

where a=±1a=\pm 1 and c=±1.c=\pm 1. Clearly, a=1a=1 and thus it follows that

α=±a0b0β+bb0.\alpha=\frac{\pm a_{0}}{b_{0}}\cdot\beta+\frac{b}{b_{0}}.
f(±a0b0x+bb0)a0=(±a0b0)ng(x)b0.\frac{f(\frac{\pm a_{0}}{b_{0}}\cdot x+\frac{b}{b_{0}})}{a_{0}}=(\pm\frac{a_{0}}{b_{0}})^{n}\frac{g(x)}{b_{0}}.

This immediately tells us that the matrix of transfer from 1,a0β,a0β2+a1β\langle 1,a_{0}\beta,a_{0}\beta^{2}+a_{1}\beta\rangle to 1,b0α,b0α2+b1α\langle 1,b_{0}\alpha,b_{0}\alpha^{2}+b_{1}\alpha\rangle is

[1b0c00a0b0].\begin{bmatrix}1&b&*\\ 0&c&*\\ 0&0&\frac{a_{0}}{b_{0}}\end{bmatrix}.

Since n4n\geq 4 the above matrix must also be integral and invertible. It follows that a0=b0a_{0}=b_{0} and thus 𝔡𝔦𝔰𝔠(f)=𝔡𝔦𝔰𝔠(g)\mathfrak{disc}(f)=\mathfrak{disc}(g).

Since R(fe,m)=R(gd,m)R^{\prime}_{(f_{e},m)}=R^{\prime}_{(g_{d},m^{\prime})} we also have

𝔡𝔦𝔰𝔠(f)m2=𝔡𝔦𝔰𝔠(g)m2,\frac{\mathfrak{disc}(f)}{m^{2}}=\frac{\mathfrak{disc}(g)}{m^{\prime 2}},

and we immediately get m=m.m=m^{\prime}.

Now, without loss of generality, we may assume c=1c=1, for if c=1c=-1 we may change g(x)g(x) to (1)ng(x).(-1)^{n}g(-x). We let l=bb0.l=\frac{b}{b_{0}}\in\mathbb{Q}. Thus, f(x+l)=g(x).f(x+l)=g(x).

We make note of the fact that ff and gg are weakly divisible by the same value mm.

Observing the matrix of transformation from RfeR_{f_{e}} to RgdR_{g_{d}} given using LABEL:l_matrix_of_tranformation, we notice that the entry in the (n1)th(n-1)^{th} row and nthn^{th} column of the matrix of transfer for the canonical bases of RfeR_{f_{e}} to RgdR_{g_{d}} is de+ld-e+l.

Thus, the entry in the (n1)th(n-1)^{th} row and nthn^{th} column of the matrix of transfer for the canonical bases of R(fe,m)R^{\prime}_{(f_{e},m)} to R(gd,m)R^{\prime}_{(g_{d},m)} is de+lm\frac{d-e+l}{m}. Thus r=de+lmr=\frac{d-e+l}{m} must be an integer. It follows that r:g(x)=f(x+mrd+e).\exists r\in\mathbb{Z}:g(x)=f(x+mr-d+e).

Lemma 14.

If ff is a real monic polynomial, then there exists a ρf\rho_{f}\in\mathbb{R} (continuously varying with ff) such that λρnf(xρ)\lambda\rho^{n}f(\frac{x}{\rho}) is a Reduced-11-polynomial for all ρρf\rho\geq\rho_{f} and λ1\lambda\geq 1.

If ρm1/(n2)ρf\rho\geq m^{1/(n-2)}\rho_{f} and λ1\lambda\geq 1 the polynomial λρnf(xρ)\lambda\rho^{n}f(\frac{x}{\rho}) is a Reduced-mm-polynomial.

Proof.

We follow the argument of Bhargava-Shankar-Wang proving in Lemma 5.2 in[2].

Set

f(x)=xn+an1xn1++akxnk++a0f(x)=x^{n}+a_{n-1}x^{n-1}+\cdots+a_{k}x^{n-k}+\cdots+a_{0}

and set an=1.a_{n}=1.

We perform a Gram-Schmidt reduction of the basis for RfR_{f} using the canonical distance form on RfrsR_{f}\bigotimes\mathbb{R}\simeq\mathbb{R}^{r}\bigoplus\mathbb{C}^{s}(r+2s=nr+2s=n). We write

1,δ,δ2+an1δ,,i=0k1aniδki,,i=0n2aniδki=BM[(t1,t2,,tn)]\langle 1,\>\delta,\>\delta^{2}+a_{n-1}\delta,\>\cdots,\>\sum_{i=0}^{k-1}a_{n-i}\delta^{k-i},\>\cdots,\>\sum_{i=0}^{n-2}a_{n-i}\delta^{k-i}\rangle=\langle B\rangle M[(t_{1},t_{2},...,t_{n})]

where BB is an orthonormal ordered basis of vectors, MM is an upper triangular unipotent matrix, and [(t1,t2,..,tn)][(t_{1},t_{2},..,t_{n})] denotes a diagonal matrix with those entries along the diagonal. Now we note that if

1,δ,\displaystyle\langle 1,\>\delta,\> δ2+an1δ,,i=0k1aniδki,,i=0n2aniδn1i\displaystyle\delta^{2}+a_{n-1}\delta,\>\cdots,\>\sum_{i=0}^{k-1}a_{n-i}\delta^{k-i},\>\cdots,\>\sum_{i=0}^{n-2}a_{n-i}\delta^{n-1-i}\rangle
=BM[(t1,t2,,tk,,tn)]\displaystyle=\langle B\rangle M[(t_{1},t_{2},\cdots,t_{k},\cdots,t_{n})]

then, using eq. 25, we see that

1,λ(ρδ),\displaystyle\langle 1,\>\lambda(\rho\delta),\> λ((ρδ)2+an1ρδ),,λ(i=0k1ani(ρδ)ki),,λ(i=0n2ani(ρδ)n1i)\displaystyle\lambda((\rho\delta)^{2}+a_{n-1}\rho\delta),\>\cdots,\>\lambda(\sum_{i=0}^{k-1}a_{n-i}(\rho\delta)^{k-i}),\>\cdots,\>\lambda(\sum_{i=0}^{n-2}a_{n-i}(\rho\delta)^{n-1-i}\rangle)
=BMρ[(t1,λρt2,,λρk1tk,,λρn1tn)].\displaystyle=\langle B\rangle M_{\rho}[(t_{1},\lambda\rho t_{2},\cdots,\lambda\rho^{k-1}t_{k},\dots,\lambda\rho^{n-1}t_{n})].

Thus, for each polynomial ff one may find ρf\rho_{f} which is a continuous function of ff such that λρnf(xρ)\lambda\rho^{n}f(\frac{x}{\rho}) is Minkowski reduced for all λ1\lambda\geq 1 and ρρf\rho\geq\rho_{f}. One can simply take

ρf:=max{maxi3{(2t2ti)1/(i2)},2t1t2}.\rho_{f}:=\max\{\max_{i\geq 3}\{(\frac{2t_{2}}{t_{i}})^{1/(i-2)}\},\frac{2t_{1}}{t_{2}}\}.

Furthermore, we note that translating the polynomial changes the canonical basis by an upper triangular matrix.

Thus, if g(x)=λρnf(Xρ),g(x)=\lambda\rho^{n}f(\frac{X}{\rho}), then the canonical basis of R(gl,m)R^{\prime}_{(g_{l},m)} after Gram-Schmidt process will look like

BMρ,m,l[(t1,λρt2,,λρk1tk,,λρn2tn1,λρn1mtn)]\langle B\rangle M_{\rho,m,l}[(t_{1},\lambda\rho t_{2},\cdots,\lambda\rho^{k-1}t_{k},\cdots,\lambda\rho^{n-2}t_{n-1},\frac{\lambda\rho^{n-1}}{m}t_{n})]

where Mρ,m,lM_{\rho,m,l} is a uni-potent upper triangular matrix.

It follows that for ρm1/(n2)ρf\rho\geq m^{1/(n-2)}\rho_{f} and λ1,\lambda\geq 1, λρnf(Xρ)\lambda\rho^{n}f(\frac{X}{\rho}) is Reduced-mm-polynomial. ∎

Appendix C The structure of the discriminant polynomial

Let

f(X,Y):=a0Xn+a1Xn1Y++anYnf(X,Y):=a_{0}X^{n}+a_{1}X^{n-1}Y+\cdots+a_{n}Y^{n}

denote a binary form of degree nn. Let an0a_{n}\neq 0. Let

𝔡𝔦𝔰𝔠(a0,a1,,an):=𝔡𝔦𝔰𝔠(f)\mathfrak{disc}(a_{0},a_{1},\cdots,a_{n}):=\mathfrak{disc}(f)

denote the discriminant as a polynomial in the coefficients of f.f. Then, we know that

a0𝔡𝔦𝔰𝔠(f)=Res(f,Xf).a_{0}\mathfrak{disc}(f)=Res(f,\frac{\partial}{\partial X}f).

Since, the resultant can be seen as the determinant of the Sylvester matrix. The Sylvester matrix for these polynomials is

[a0a1a2an1an00000a0a1an2an1an00000a0an3an2an1000000a3a4a5an00000a2a3a4an1an0000a1a2a3an2an1anna0(n1)a1(n2)a2an1000000na0(n1)a12an2an1000000na03an32an2an1000000(n2)a2(n3)a3an100000(n1)a1(n2)a2(n3)a32an2an10000na0(n1)a1(n2)a23an32an2an1].\setcounter{MaxMatrixCols}{11}\begin{bmatrix}a_{0}&a_{1}&a_{2}&\cdots&a_{n-1}&a_{n}&0&\cdots&0&0&0\\ 0&a_{0}&a_{1}&\cdots&a_{n-2}&a_{n-1}&a_{n}&\cdots&0&0&0\\ 0&0&a_{0}&\cdots&a_{n-3}&a_{n-2}&a_{n-1}&\cdots&0&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots\\ 0&0&0&\cdots&a_{3}&a_{4}&a_{5}&\cdots&a_{n}&0&0\\ 0&0&0&\cdots&a_{2}&a_{3}&a_{4}&\cdots&a_{n-1}&a_{n}&0\\ 0&0&0&\cdots&a_{1}&a_{2}&a_{3}&\cdots&a_{n-2}&a_{n-1}&a_{n}\\ na_{0}&(n-1)a_{1}&(n-2)a_{2}&\cdots&a_{n-1}&0&0&\cdots&0&0&0\\ 0&na_{0}&(n-1)a_{1}&\cdots&2a_{n-2}&a_{n-1}&0&\cdots&0&0&0\\ 0&0&na_{0}&\cdots&3a_{n-3}&2a_{n-2}&a_{n-1}&\cdots&0&0&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots\\ 0&0&0&\cdots&(n-2)a_{2}&(n-3)a_{3}&\cdots&\cdots&a_{n-1}&0&0\\ 0&0&0&\cdots&(n-1)a_{1}&(n-2)a_{2}&(n-3)a_{3}&\cdots&2a_{n-2}&a_{n-1}&0\\ 0&0&0&\cdots&na_{0}&(n-1)a_{1}&(n-2)a_{2}&\cdots&3a_{n-3}&2a_{n-2}&a_{n-1}\end{bmatrix}. (28)

The following lemmata are easy to observe from the following:

Lemma 15.
𝔡𝔦𝔰𝔠(f)=anΔ1+anan1Δ2+an12Δ3+an2Δ4.\mathfrak{disc}(f)=a_{n}\Delta_{1}+a_{n}a_{n-1}\Delta_{2}+a_{n-1}^{2}\Delta_{3}+a_{n}^{2}\Delta_{4}.

where

Δ1=4an23𝔡𝔦𝔰𝔠(a0Xn2+a1Xn3Y++an2Yn2)\Delta_{1}=4a_{n-2}^{3}\cdot\mathfrak{disc}(a_{0}X^{n-2}+a_{1}X^{n-3}Y+\cdots+a_{n-2}Y^{n-2})

and

Δ3=an12𝔡𝔦𝔰𝔠(a0Xn1+a1Xn2Y++an2XYn2+an1Yn1).\Delta_{3}=a_{n-1}^{2}\mathfrak{disc}(a_{0}X^{n-1}+a_{1}X^{n-2}Y+\cdots+a_{n-2}XY^{n-2}+a_{n-1}Y^{n-1}).
Proof.

We carefully expand the determinant of the Sylvester Matrix along the last two columns. We can clearly see we can write the determinant as

2anan2Δ1+an12Δ2+anan1Δ3+an2Δ4.2a_{n}a_{n-2}\Delta_{1}+a_{n-1}^{2}\Delta_{2}+a_{n}a_{n-1}\Delta_{3}+a_{n}^{2}\Delta_{4}. (29)

We will pick Δ1[a0,a1,,an2]\Delta_{1}\in\mathbb{Z}[a_{0},a_{1},\cdots,a_{n-2}] and as the components of Δ1\Delta_{1} containing an1a_{n-1} and ana_{n} can be pushed inside Δ2,Δ3,Δ4\Delta_{2},\Delta_{3},\Delta_{4}. One can then choose Δ3[a0,a1,,an1]\Delta_{3}\in\mathbb{Z}[a_{0},a_{1},\cdots,a_{n-1}] as components of Δ2\Delta_{2} containing ana_{n} can be pushed inside Δ2\Delta_{2} and Δ4.\Delta_{4}.

Then, Δ1\Delta_{1} can be seen as the determinant of (i.e. the minor corresponding to an(2an2)a_{n}\cdot(2a_{n-2}) where ana_{n} and an1a_{n-1} are set to zero looks as follows)

[a0a1a200000a0a1an200000a0an3an200000a3a4a50000a2a3a40na0(n1)a1(n2)a200000na0(n1)a12an200000na03an32an200000(n2)a2(n3)a30000(n1)a1(n2)a2(n3)a32an2].\begin{bmatrix}a_{0}&a_{1}&a_{2}&\cdots&0&0&0&\cdots&0\\ 0&a_{0}&a_{1}&\cdots&a_{n-2}&0&0&\cdots&0\\ 0&0&a_{0}&\cdots&a_{n-3}&a_{n-2}&0&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&a_{3}&a_{4}&a_{5}&\cdots&0\\ 0&0&0&\cdots&a_{2}&a_{3}&a_{4}&\cdots&0\\ na_{0}&(n-1)a_{1}&(n-2)a_{2}&\cdots&0&0&0&\cdots&0\\ 0&na_{0}&(n-1)a_{1}&\cdots&2a_{n-2}&0&0&\cdots&0\\ 0&0&na_{0}&\cdots&3a_{n-3}&2a_{n-2}&0&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&(n-2)a_{2}&(n-3)a_{3}&\cdots&\cdots&0\\ 0&0&0&\cdots&(n-1)a_{1}&(n-2)a_{2}&(n-3)a_{3}&\cdots&2a_{n-2}\end{bmatrix}.

Now doing operations RnR1,Rn+1R2,Rn+1R3,,R2n2Rn2R_{n}-R_{1},R_{n+1}-R_{2},R_{n+1}-R_{3},\cdots,R_{2n-2}-R_{n-2} we get the following matrix with the same determinant.

[a0a1a200000a0a1an200000a0an3an200000a3a4a50000a2a3a40(n1)a0(n2)a1(n3)a200000(n1)a0(n2)a1an200000(n1)a02an3an200000(n3)a2(n4)a30000(n1)a1(n2)a2(n3)a32an2].\begin{bmatrix}a_{0}&a_{1}&a_{2}&\cdots&0&0&0&\cdots&0\\ 0&a_{0}&a_{1}&\cdots&a_{n-2}&0&0&\cdots&0\\ 0&0&a_{0}&\cdots&a_{n-3}&a_{n-2}&0&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&a_{3}&a_{4}&a_{5}&\cdots&0\\ 0&0&0&\cdots&a_{2}&a_{3}&a_{4}&\cdots&0\\ (n-1)a_{0}&(n-2)a_{1}&(n-3)a_{2}&\cdots&0&0&0&\cdots&0\\ 0&(n-1)a_{0}&(n-2)a_{1}&\cdots&a_{n-2}&0&0&\cdots&0\\ 0&0&(n-1)a_{0}&\cdots&2a_{n-3}&a_{n-2}&0&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&\cdots&(n-3)a_{2}&(n-4)a_{3}&\cdots&\cdots&0\\ 0&0&0&\cdots&(n-1)a_{1}&(n-2)a_{2}&(n-3)a_{3}&\cdots&2a_{n-2}\end{bmatrix}.

We note that the above matrix is very close to Sylvester matrix for computing the discriminant of

f(x,y)anYnan1XYn1Y=X(a0Xn2+a1Xn3Y++an2Yn2).\frac{f(x,y)-a_{n}Y^{n}-a_{n-1}XY^{n-1}}{Y}=X(a_{0}X^{n-2}+a_{1}X^{n-3}Y+\cdots+a_{n-2}Y^{n-2}).

More particularly, the determinant of the above matrix is twice of the discriminant of the above form. This can be easily observed by expanding the determinant of the above matrix and that of the Sylvester Matrix along the final column.

Clearly, since an2a_{n-2} is the Resultant of XX and a0Xn2+a1Xn3Y++an2Yn2,a_{0}X^{n-2}+a_{1}X^{n-3}Y+\cdots+a_{n-2}Y^{n-2}, we get

𝔡𝔦𝔰𝔠(f(x,y)anYnan1XYn1Y)=an22𝔡𝔦𝔰𝔠(a0Xn2+a1Xn3Y++an2Yn2).\mathfrak{disc}(\frac{f(x,y)-a_{n}Y^{n}-a_{n-1}XY^{n-1}}{Y})=a_{n-2}^{2}\cdot\mathfrak{disc}(a_{0}X^{n-2}+a_{1}X^{n-3}Y+\cdots+a_{n-2}Y^{n-2}).

Substituting back into eq. 29 we get the appropriate structure for Δ1\Delta_{1}.

To get the structure of Δ3\Delta_{3} we see that

an12Δ3=𝔡𝔦𝔰𝔠(f(X,Y)anYn)a_{n-1}^{2}\Delta_{3}=\mathfrak{disc}(f(X,Y)-a_{n}Y^{n})

by definition. ∎

References

  • [1] T. Z. Avner Ash, Jos Brakenhoff. Equality of polynomial and field discriminants. Experimental Mathematics, 16(3):367–374, 2007.
  • [2] M. Bhargava, A. Shankar, and X. Wang. Squarefree values of polynomial discriminants I. Invent. Math., 228(3):1037–1073, 2022.
  • [3] M. Bhargava, A. Shankar, and X. Wang. Squarefree values of polynomial discriminants II, 2022.
  • [4] I. D. Corso, R. Dvornicich, and D. Simon. Decomposition of primes in non-maximal orders. Acta Arithmetica, 120:231–244, 2005.
  • [5] J. S. Ellenberg and A. Venkatesh. The number of extensions of a number field with fixed degree and bounded discriminant. Annals of Mathematics, 163:723–741, 2003.
  • [6] N. Kaplan, J. Marcinek, and R. Takloo-Bighash. Distribution of orders in number fields. Research in the Mathematical Sciences, 2:1–57, 2014.
  • [7] G. Malle. On the distribution of galois groups. Journal of Number Theory, 92(2):315–329, 2002.
  • [8] G. Malle. On the Distribution of Galois groups, II. Experimental Mathematics, 13(2):129 – 136, 2004.
  • [9] G. D. Patil. Rings of finite rank over integers,PhD Thesis-University of Toronto. 2023.
  • [10] G. D. Patil. Weakly divisible rings, 2024.
  • [11] M. M. Wood. Rings and ideals parameterized by binary nn-ic forms. Journal of the London Mathematical Society, 83(1):208–231, 2011.