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

On a Special Metric in Cyclotomic Fields

Katerina Saettone, Alexandru Zaharescu, Zhuo Zhang Katerina Saettone: Department of Mathematics, University of Illinois Urbana-Champaign, Altgeld Hall, 1409 W. Green Street, Urbana, IL, 61801, USA [email protected] Alexandru Zaharescu: Department of Mathematics, University of Illinois Urbana-Champaign, Altgeld Hall, 1409 W. Green Street, Urbana, IL, 61801, USA and Simion Stoilow Institute of Mathematics of the Romanian Academy, P. O. Box 1-764, RO-014700 Bucharest, Romania [email protected] Zhuo Zhang: Department of Mathematics, University of Illinois Urbana-Champaign, Altgeld Hall, 1409 W. Green Street, Urbana, IL, 61801, USA [email protected]
Abstract.

Let pp be an odd prime, and let ω\omega be a primitive ppth root of unity. In this paper, we introduce a metric on the cyclotomic field K=(ω)K=\mathbb{Q}(\omega). We prove that this metric has several remarkable properties, such as invariance under the action of the Galois group. Furthermore, we show that points in the ring of integers 𝒪K\mathcal{O}_{K} behave in a highly uniform way under this metric. More specifically, we prove that for a certain hypercube in 𝒪K\mathcal{O}_{K} centered at the origin, almost all pairs of points in the cube are almost equi-distanced from each other, when pp and NN are large enough. When suitably normalized, this distance is exactly 1/61/\sqrt{6}.

Key words and phrases:
Cyclotomic fields, Lattice points
2020 Mathematics Subject Classification:
Primary 11R18; Secondary 11B99, 11P21

1. Introduction

Cyclotomic fields play an essential role in algebra and number theory, particularly in understanding the behaviour of prime numbers, and the solutions to Diophantine equations. In this paper, we uncover properties of cyclotomic fields equipped with a special metric, which we study from both algebraic and probabilistic standpoints.

Let pp be an odd prime and let ω\omega be a primitive pp-th root of unity. The extension of \mathbb{Q} generated by ω\omega in the field of complex numbers is the pp-th cyclotomic field K=(ω)K=\mathbb{Q}(\omega). We shall denote by TrK/\operatorname{Tr}_{K/\mathbb{Q}} the trace map of the number field KK (the precise definition of TrK/\operatorname{Tr}_{K/\mathbb{Q}} will be reviewed in Section 2).

For αK\alpha\in K, we denote by vαv_{\alpha} the vector in p1\mathbb{Q}^{p-1} whose jjth component is TrK/(αωj)\operatorname{Tr}_{K/\mathbb{Q}}(\alpha\omega^{j}), for 1jp11\leq j\leq p-1. In this paper, we define d(α,β)d(\alpha,\beta), the distance between α\alpha and β\beta in KK, as the Euclidean distance between the vectors vαv_{\alpha} and vβv_{\beta} in p1\mathbb{Q}^{p-1}. We shall show that dd is a metric on KK, where positive-definiteness is the only nontrivial property. Note that dd is canonically defined and is independent of the choice of ω\omega.

We aim to investigate this metric dd from several perspectives. In Section 3, we show that dd has certain nice properties that are related to the algebraic and number-theoretic structure of (ω)\mathbb{Q}(\omega). For instance, the metric dd is invariant under the action of the Galois group G=Gal(K/)G=\operatorname{Gal}(K/\mathbb{Q}). In turn, this gives us an analogy of Krasner’s lemma within the context of cyclotomic fields equipped with the metric dd.

In Section 4, we derive an explicit formula for the metric in terms of the coordinates under the canonical basis {ω,,ωp1}\{\omega,\ldots,\omega^{p-1}\} of KK.

In the rest of the paper, we build on the ideas of [1] and [2] to study the metric dd from a statistical point of view. More specifically, for a positive integer NN, we denote by B(p,N)B(p,N) the symmetric box of cyclotomic lattice points:

B(p,N):={a1ω++ap1ωp1:a1,,ap1[N,N]},B(p,N):=\big{\{}a_{1}\omega+\cdots+a_{p-1}\omega^{p-1}:a_{1},\dots,a_{p-1}\in[-N,N]\cap\mathbb{Z}\big{\}},

which lies in the ring of integers 𝒪K\mathcal{O}_{K}. In Section 5, we normalize the metric dd so that the diameter of B(p,N)B(p,N) is exactly 1 in the sense of metric spaces, i.e., the points furthest apart in B(p,N)B(p,N) are at a distance of exactly 11 from each other. This gives us a scaled distance, denoted by 𝔡p,N(α,β)\operatorname{\mathfrak{d}}_{p,N}(\alpha,\beta), which serves as a unitary means of comparing the spacing of points in different hypercubes B(p,N)B(p,N), as pp and NN vary. Our main theorem states that points in B(p,N)B(p,N) are almost equi-distanced from each other in the following sense.

Theorem 1.1.

For any ε>0\varepsilon>0, there exists an absolute and effectively computable constant A(ε)A(\varepsilon) such that if N,p>A(ε)N,p>A(\varepsilon), then

1#B(p,N)2#{(α,β)B(p,N)×B(p,N):|𝔡p,N(α,β)16|>ε}<ε.\frac{1}{\#B(p,N)^{2}}\#\left\{(\alpha,\beta)\in B(p,N)\times B(p,N):\left|\mathfrak{d}_{p,N}(\alpha,\beta)-\frac{1}{\sqrt{6}}\right|>\varepsilon\right\}<\varepsilon.

Theorem 1.1 reveals a surprising uniformity in the spacing of points among the high-dimensional lattice points in KK. It provides insight into a certain “statistical regularity” in the geometric properties of cyclotomic fields when viewed through the lens of this particular metric. Theorem 1.1 will follow from Theorem 6.8, which is an explicit quantitative version that we shall prove in Section 6. Our methods rely on calculating the various moments of distances between points in B(p,N)B(p,N).

2. Notations and definition of the metric

2.1. Notations and setup

In this subsection, we set up some notations and recall some preliminary facts from algebraic number theory that will be needed in the later discussions. More details can be found in [5], [6], and [11].

Throughout this paper, let pp be an odd prime, and let ω\omega be a primitive ppth root of unity, say ω=e2πi/p\omega=e^{2\pi i/p}. Let K=(ω)K=\mathbb{Q}(\omega) be the ppth cyclotomic field. It is well known that the Galois group G:=Gal(K/)G:=\operatorname{Gal}(K/\mathbb{Q}) is isomorphic to the group (/p)×(\mathbb{Z}/p\mathbb{Z})^{\times}, which is cyclic of order p1p-1.

We denote by 𝒪K\mathcal{O}_{K} the ring of integers of KK, that is, the integral closure of \mathbb{Z} in KK. It is well known that rings of integers have integral bases, and in this case, an integral basis of 𝒪K\mathcal{O}_{K} is given by {ω,,ωp1}\{\omega,\ldots,\omega^{p-1}\}. Therefore,

𝒪K={a1ω++ap1ωp1:a1,,ap1}.\mathcal{O}_{K}=\{a_{1}\omega+\ldots+a_{p-1}\omega^{p-1}:a_{1},\ldots,a_{p-1}\in\mathbb{Z}\}.

Many key properties of number fields can be studied via the trace map. Since cyclotomic fields are always Galois over \mathbb{Q}, the trace map TrK/\operatorname{Tr}_{K/\mathbb{Q}} has a simple definition in this case, which is

(1) TrK/(α)=σGal(K/)σ(α),αK.\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)=\sum_{\sigma\in\operatorname{Gal}(K/\mathbb{Q})}\sigma(\alpha),\quad\alpha\in K.

It can be proved that TrK/(α)\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)\in\mathbb{Q} for all αK\alpha\in K. Furthermore, if α𝒪K\alpha\in\mathcal{O}_{K}, then TrK/(α)\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)\in\mathbb{Z}.

Finally, for complex-valued functions ff and gg, we write fgf\ll g or f=O(g)f=O(g) to indicate that there exists an absolute an effectively computable constant CC such that |f|C|g||f|\leq C|g| for all inputs.

2.2. Definition of the metric

We now formally define the metric dd mentioned in the introduction. The metric is, in fact, induced by a norm on KK as a \mathbb{Q}-vector space. The norm is defined as follows:

Definition 2.1.

For any αK\alpha\in K, we define

α=j=1p1(TrK/(αωj))2=vαE,\|\alpha\|=\sqrt{\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}(\alpha\omega^{j})\right)^{2}}=\|v_{\alpha}\|_{E},

where vαp1v_{\alpha}\in\mathbb{Q}^{p-1} is the vector whose jjth component is Tr(αωj)\operatorname{Tr}(\alpha\omega^{j}) and E\|\cdot\|_{E} denotes the usual Euclidean norm on p1\mathbb{Q}^{p-1}.

Definition 2.2.

For α,βK\alpha,\beta\in K, we define their distance d(α,β)d(\alpha,\beta) to be αβ\|\alpha-\beta\|.

Theorem 2.3.

The function \|\cdot\| defined as in Definition 2.1 is a norm on KK.

Proof.

We verify the three conditions of a norm. The triangle inequality follows immediately from the usual triangle inequality in Euclidean spaces p1p1\mathbb{Q}^{p-1}\subseteq\mathbb{R}^{p-1}.

For any αK\alpha\in K and λ\lambda\in\mathbb{Q}, we need to prove that λα=|λ|α\|\lambda\alpha\|=|\lambda|\|\alpha\|. This follows from the \mathbb{Q}-linearity of trace, since it implies that vλα=λvαv_{\lambda\alpha}=\lambda v_{\alpha}.

It remains to prove positive-definiteness. Clearly α0\|\alpha\|\geq 0. Suppose α=0\|\alpha\|=0. Then vαp1v_{\alpha}\in\mathbb{Q}^{p-1} is the zero vector. Hence,

TrK/(αωj)=0,\operatorname{Tr}_{K/\mathbb{Q}}(\alpha\omega^{j})=0,

for all j=1,,p1j=1,\ldots,p-1. Suppose α0\alpha\neq 0. Then we may write

1α=c1ω++cp1ωp1,where ci.\frac{1}{\alpha}=c_{1}\omega+\ldots+c_{p-1}\omega^{p-1},\quad\text{where }c_{i}\in\mathbb{Q}.

Therefore,

1=c1αω++cp1αωp1.1=c_{1}\alpha\omega+\ldots+c_{p-1}\alpha\omega^{p-1}.

Taking the trace of both sides, and using the fact that trace is \mathbb{Q}-linear, we have

p1=j=1p1cjTrK/(αωj)=0,p-1=\sum_{j=1}^{p-1}c_{j}\operatorname{Tr}_{K/\mathbb{Q}}(\alpha\omega^{j})=0,

which is a contradiction. Hence, α=0\alpha=0. ∎

It follows that the function dd is indeed a metric on KK. The distance defined in this manner closely resembles the Euclidean distance in vector spaces but also has properties that are well-suited to the study of cyclotomic fields. This will be further explored in Section 3.

We remark that the norm in Definition 2.1 must be distinguished from the usual norm of an algebraic number (say, over a Galois extension), which is defined to be the product of all its Galois conjugates. There also exist several other notions of norms over number fields. For example, one can define the Siegel norm of algebraic numbers (see [4] and [10] for its construction and some interesting properties; for some questions related to Siegel’s trace problem, see [8] and [9]). In this paper, the word “norm” always refers to the norm we just defined, unless stated otherwise.

3. Properties of the metric

The metric dd on K=(ω)K=\mathbb{Q}(\omega) defined as above is the main object we investigate in this paper. To convince the readers that the metric is a natural object worth studying, we shall first prove a number of remarkable facts about this metric, the most important one of which is the invariance under the action of the Galois group. This is the content of the following proposition.

3.1. Invariance under the Galois group action

Proposition 3.1.

The metric dd is invariant under the action of the Galois group G=Gal(K/)G=\operatorname{Gal}(K/\mathbb{Q}). In other words, for any σG\sigma\in G and α,βK\alpha,\beta\in K, we have

d(α,β)=d(σ(α),σ(β)).d(\alpha,\beta)=d(\sigma(\alpha),\sigma(\beta)).
Proof.

It suffices to show that the norm in Definition 2.1 is invariant under GG, i.e., σ(α)=α\|\sigma(\alpha)\|=\|\alpha\| for all αK\alpha\in K and σG\sigma\in G. Suppose σ1(ω)=ωk\sigma^{-1}(\omega)=\omega^{k}, where 1kp11\leq k\leq p-1. Then we have

σ(α)\displaystyle\|\sigma(\alpha)\| =j=1p1(TrK/(σ(α)ωj))2\displaystyle=\sqrt{\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}(\sigma(\alpha)\omega^{j})\right)^{2}}
=j=1p1(TrK/(σ(ασ1(ω)j)))2\displaystyle=\sqrt{\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}\left(\sigma\left(\alpha\cdot\sigma^{-1}(\omega)^{j}\right)\right)\right)^{2}}
=j=1p1(TrK/(ασ1(ω)j))2\displaystyle=\sqrt{\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}\left(\alpha\cdot\sigma^{-1}(\omega)^{j}\right)\right)^{2}}
=j=1p1(TrK/(αωkj))2,\displaystyle=\sqrt{\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}\left(\alpha\omega^{kj}\right)\right)^{2}},

where the third equality follows from the fact that TrK/\operatorname{Tr}_{K/\mathbb{Q}} is invariant under GG. Since kk must be coprime to pp, it follows that {kj:1jp1}\{kj:1\leq j\leq p-1\} is a permutation of {j:1jp1}\{j:1\leq j\leq p-1\}. Hence, σ(α)=α\|\sigma(\alpha)\|=\|\alpha\|, as required. ∎

3.2. An analogue of Krasner’s lemma

As a consequence of Proposition 3.1, we now prove that the metric dd has another surprising property, with which we will draw an analogy between the following Krasner’s lemma.

Theorem 3.2 (Krasner’s lemma).

Let κ\kappa be a complete field with respect to a nonarchimedean valuation and let Ω\Omega be an algebraic closure of κ\kappa. Let αΩ\alpha\in\Omega be separable over κ\kappa and let α=α1,,αn\alpha=\alpha_{1},\ldots,\alpha_{n} be the conjugates of α\alpha over κ\kappa. Suppose that for βΩ\beta\in\Omega we have

|αβ|<|ααi| for i=2,,n,|\alpha-\beta|<\left|\alpha-\alpha_{i}\right|\quad\text{ for }i=2,\ldots,n,

where |||\cdot| denotes the unique extension of the valuation to Ω\Omega. Then κ(α)κ(β)\kappa(\alpha)\subseteq\kappa(\beta).

Proof.

See Lemma 8.1.6 in [7]

We now prove the following analogous result.

Theorem 3.3.

Let K=(ω)K=\mathbb{Q}(\omega), where ω\omega is a primitive ppth root of unity. Let α\alpha be an element of KK and let α1,,αn\alpha_{1},\ldots,\alpha_{n} be the conjugates of α\alpha over KK, with α1=α\alpha_{1}=\alpha. Suppose that for βK\beta\in K we have

d(α,β)<12d(α,αi) for i=2,,n,d(\alpha,\beta)<\frac{1}{2}\,d(\alpha,\alpha_{i})\quad\text{ for }i=2,\ldots,n,

where dd is the metric in Definition 2.2. Then (α)(β)\mathbb{Q}(\alpha)\subseteq\mathbb{Q}(\beta).

Proof.

Let G=Gal(K/)G=\operatorname{Gal}(K/\mathbb{Q}). By Galois theory, (α)(β)\mathbb{Q}(\alpha)\subseteq\mathbb{Q}(\beta) if and only if Gal(K/(β))Gal(K/(α))\operatorname{Gal}(K/\mathbb{Q}(\beta))\subseteq\operatorname{Gal}(K/\mathbb{Q}(\alpha)). Since GG is a cyclic group of order p1p-1, the preceding condition is equivalent to |Gal(K/(β))|\left|\operatorname{Gal}(K/\mathbb{Q}(\beta))\right| dividing |Gal(K/(α))|\left|\operatorname{Gal}(K/\mathbb{Q}(\alpha))\right|, which is then equivalent to [K:(β)][K:\mathbb{Q}(\beta)] dividing [K:(α)][K:\mathbb{Q}(\alpha)]. By the tower law, this is equivalent to [(α):][\mathbb{Q}(\alpha):\mathbb{Q}] dividing [(β):][\mathbb{Q}(\beta):\mathbb{Q}].

As in the statement, let α1,,αn\alpha_{1},\ldots,\alpha_{n} be the Galois conjugates of α\alpha over KK, with α1=α\alpha_{1}=\alpha. Similarly, let β1,,βm\beta_{1},\ldots,\beta_{m} be the Galois conjugates of β\beta over KK, with β1=β\beta_{1}=\beta. Since K/K/\mathbb{Q} is Galois, we have [(α):]=n[\mathbb{Q}(\alpha):\mathbb{Q}]=n and [(β):]=m[\mathbb{Q}(\beta):\mathbb{Q}]=m. Therefore, we need to prove that nn divides mm under the hypothesis that d(α,β)<12d(α,αi)d(\alpha,\beta)<\frac{1}{2}d(\alpha,\alpha_{i}) for all i=2,,ni=2,\ldots,n.

Let

r=12min2ind(α,αi).r=\frac{1}{2}\min_{2\leq i\leq n}d(\alpha,\alpha_{i}).

Then d(α,β)<rd(\alpha,\beta)<r. For any element xKx\in K, denote by B(x,r)B(x,r) the open ball centered at xx with radius rr under the metric dd. Observe that if αi,αj\alpha_{i},\alpha_{j} are two distinct Galois conjugates of α\alpha, say αi=σ(α)\alpha_{i}=\sigma(\alpha) and αj=τ(α)\alpha_{j}=\tau(\alpha), where σ,τGal(K/(α))\sigma,\tau\in\operatorname{Gal}(K/\mathbb{Q}(\alpha)), then

d(αi,αj)=d(σ(α),τ(α))=d(α,σ1τ(α))2r.d(\alpha_{i},\alpha_{j})=d(\sigma(\alpha),\tau(\alpha))=d(\alpha,\sigma^{-1}\tau(\alpha))\geq 2r.

It follows that any two distinct conjugates αi\alpha_{i} and αj\alpha_{j} are at a distance of at least 2r2r from each other. In particular, the open balls {B(αi,r):1in}\left\{B(\alpha_{i},r):1\leq i\leq n\right\} are pairwise disjoint.

We claim that for every βj\beta_{j} there exists an αi\alpha_{i} such that βjB(αi,r)\beta_{j}\in B(\alpha_{i},r). In other words, the balls contain all conjugates of β\beta. Indeed, if βj=σ(β)\beta_{j}=\sigma(\beta), then

d(σ(α),βj)=d(σ(α),σ(β))=d(α,β)<r,d(\sigma(\alpha),\beta_{j})=d(\sigma(\alpha),\sigma(\beta))=d(\alpha,\beta)<r,

by Proposition 3.1. Hence, βjB(σ(α),r)\beta_{j}\in B(\sigma(\alpha),r).

Furthermore, we claim that for 1in,1\leq i\leq n, each ball B(αi,r)B(\alpha_{i},r) contains the same number of conjugates of β\beta. Indeed, suppose αi=σ(α)\alpha_{i}=\sigma(\alpha). Then by Proposition 3.1 again, we have

d(αi,βj)=d(σ(α),βj)=d(α,σ1(βj)).d(\alpha_{i},\beta_{j})=d(\sigma(\alpha),\beta_{j})=d(\alpha,\sigma^{-1}(\beta_{j})).

Therefore, βjB(αi,r)\beta_{j}\in B(\alpha_{i},r) if and only if σ1(βj)B(α,r)\sigma^{-1}(\beta_{j})\in B(\alpha,r). Since σ\sigma is a bijection, this proves that B(α,r)B(\alpha,r) and B(αi,r)B(\alpha_{i},r) contain the same number of Galois conjugates of β\beta. This number is nonzero because βB(α,r)\beta\in B(\alpha,r). Since the balls are disjoint, we conclude that nn divides mm, as desired. ∎

Remark 3.4.

The following example illustrates that the constant 12\frac{1}{2} is optimal, in the sense that any larger constant would make the statement false. Consider p=3p=3, α=ω\alpha=\omega, and β=12\beta=-\frac{1}{2}. Then α\alpha only has one Galois conjugate other than itself, namely ω2\omega^{2}. A straightforward computation shows that

d(α,β)=32andd(α,ω2)=32.d(\alpha,\beta)=\frac{3}{\sqrt{2}}\quad\text{and}\quad d(\alpha,\omega^{2})=3\sqrt{2}.

Therefore,

d(α,β)=12d(α,ω2),d(\alpha,\beta)=\frac{1}{2}d(\alpha,\omega^{2}),

but (α)\mathbb{Q}(\alpha) is not contained in (β)\mathbb{Q}(\beta).

As a simple consequence, we deduce the following corollary, which is reminiscent of the primitive element theorem in field theory.

Corollary 3.5.

Let α,βK=(ω)\alpha,\beta\in K=\mathbb{Q}(\omega). Define γn=α+βn\gamma_{n}=\alpha+\frac{\beta}{n}. Then (γn)=(α,β)\mathbb{Q}(\gamma_{n})=\mathbb{Q}(\alpha,\beta) for all sufficiently large nn.

Proof.

Clearly (γn)(α,β)\mathbb{Q}(\gamma_{n})\subseteq\mathbb{Q}(\alpha,\beta), so it suffices to prove the reverse inclusion. Note that

d(α,γn)=αγn=βn=βn.d(\alpha,\gamma_{n})=\|\alpha-\gamma_{n}\|=\left\|\frac{\beta}{n}\right\|=\frac{\|\beta\|}{n}.

Thus, when nn is sufficiently large, we would have d(α,γn)<12d(α,σ(α))d(\alpha,\gamma_{n})<\frac{1}{2}d(\alpha,\sigma(\alpha)) for all σG\sigma\in G. Theorem 3.3 implies that (α)(γn)\mathbb{Q}(\alpha)\subseteq\mathbb{Q}(\gamma_{n}). In particular, α(γn)\alpha\in\mathbb{Q}(\gamma_{n}), and so β(γn)\beta\in\mathbb{Q}(\gamma_{n}), as desired. ∎

Not only does Corollary 3.5 prove a special case of the primitive element theorem, but it also provides a simple algorithm to find generators of subextensions of KK.

4. Computing the metric in coordinates

In this section, we aim to derive an explicit formula of the metric dd in terms of the coordinates of αK\alpha\in K under the integral basis {ω,,ωp1}\{\omega,\ldots,\omega^{p-1}\}. We first note that TrK/(1)=p1\operatorname{Tr}_{K/\mathbb{Q}}(1)=p-1, and TrK/(ω)==TrK/(ωp1)=1\operatorname{Tr}_{K/\mathbb{Q}}(\omega)=\ldots=\operatorname{Tr}_{K/\mathbb{Q}}(\omega^{p-1})=-1. Therefore, if

α=a1ω++ap1ωp1,\alpha=a_{1}\omega+\ldots+a_{p-1}\omega^{p-1},

then

TrK/(α)=(a1++ap1),\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)=-(a_{1}+\ldots+a_{p-1}),

and for j=1,,p1j=1,\ldots,p-1, we have

TrK/(αωj)=i=1ipjp1ai+(p1)apj=TrK/(α)+papj.\operatorname{Tr}_{K/\mathbb{Q}}(\alpha\omega^{j})=-\sum_{\begin{subarray}{c}i=1\\ i\neq p-j\end{subarray}}^{p-1}a_{i}+(p-1)a_{p-j}=\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)+pa_{p-j}.

Therefore,

α2\displaystyle\|\alpha\|^{2} =j=1p1TrK/(αωj)2\displaystyle=\sum_{j=1}^{p-1}\operatorname{Tr}_{K/\mathbb{Q}}(\alpha\omega^{j})^{2}
=j=1p1(TrK/(α)+papj)2\displaystyle=\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)+pa_{p-j}\right)^{2}
=j=1p1(TrK/(α)+paj)2\displaystyle=\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)+pa_{j}\right)^{2}
=j=1p1(TrK/(α)2+2pajTrK/(α)+p2aj2)\displaystyle=\sum_{j=1}^{p-1}\left(\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)^{2}+2pa_{j}\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)+p^{2}a_{j}^{2}\right)
=(p1)TrK/(α)2+2pTrK/(α)j=1p1aj+p2j=1p1aj2\displaystyle=(p-1)\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)^{2}+2p\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)\sum_{j=1}^{p-1}a_{j}+p^{2}\sum_{j=1}^{p-1}a_{j}^{2}
=(p1)TrK/(α)22pTrK/(α)2+p2j=1p1aj2\displaystyle=(p-1)\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)^{2}-2p\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)^{2}+p^{2}\sum_{j=1}^{p-1}a_{j}^{2}
=p2j=1p1aj2(p+1)TrK/(α)2.\displaystyle=p^{2}\sum_{j=1}^{p-1}a_{j}^{2}-(p+1)\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)^{2}.

Hence, we have arrived at the following convenient formula, which we shall frequently use in the later sections:

Lemma 4.1.

Suppose α=a1ω++ap1ωp1K\alpha=a_{1}\omega+\ldots+a_{p-1}\omega^{p-1}\in K. Then

(2) α2=p2αE2(p+1)TrK/(α)2,\|\alpha\|^{2}=p^{2}\|\alpha\|_{E}^{2}-(p+1)\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)^{2},

where αE\|\alpha\|_{E} the Euclidean norm of α\alpha, i.e., αE2=i=1p1ai2\|\alpha\|_{E}^{2}=\sum_{i=1}^{p-1}a_{i}^{2}.

Also note that by the Cauchy-Schwarz inequality,

TrK/(α)2=(j=1p1aj)2(p1)j=1p1aj2=(p1)αE2,\operatorname{Tr}_{K/\mathbb{Q}}(\alpha)^{2}=\left(\sum_{j=1}^{p-1}a_{j}\right)^{2}\leq(p-1)\sum_{j=1}^{p-1}a_{j}^{2}=(p-1)\|\alpha\|_{E}^{2},

so we conclude that

α2(p2(p+1)(p1))αE2=αE2.\|\alpha\|^{2}\geq\left(p^{2}-(p+1)(p-1)\right)\|\alpha\|_{E}^{2}=\|\alpha\|_{E}^{2}.

In other words, the norm of α\alpha is always larger than or equal to the Euclidean norm of α\alpha.

5. The normalized distance

Let B(p,N)B(p,N) be the hypercube

B(p,N):={a1ω++ap1ωp1:ai[N,N]}𝒪K.B(p,N):=\left\{a_{1}\omega+\ldots+a_{p-1}\omega^{p-1}:a_{i}\in\mathbb{Z}\cap[-N,N]\right\}\subset\mathcal{O}_{K}.

Then B(p,N)B(p,N) contains (2N+1)p1(2N+1)^{p-1} points in total. In this section, we introduce a normalized distance on B(p,N)B(p,N). In order to do so, we shall need to compute the diameter of the hypercube B(p,NB(p,N). This is done in the following lemma.

Lemma 5.1.

The diameter of B(p,N)B(p,N), i.e., the maximum distance between two points in B(p,N)B(p,N), is exactly

diamB(p,N)=2Npp1,\operatorname{diam}B(p,N)=2Np\sqrt{p-1},

which is achieved by the following pairs of points

α=i=1p1N(ω)i1=NωNω2++Nωp2Nωp1andβ=α.\alpha=\sum_{i=1}^{p-1}N(-\omega)^{i-1}=N\omega-N\omega^{2}+\ldots+N\omega^{p-2}-N\omega^{p-1}\quad\text{and}\quad\beta=-\alpha.
Proof.

It suffices to maximize equation (2) for αβ\alpha-\beta, where α,βB(p,N)\alpha,\beta\in B(p,N). Note that

αβ=2α=2Nω2Nω2+2Nωp22Nωp1.\alpha-\beta=2\alpha=2N\omega-2N\omega^{2}-\ldots+2N\omega^{p-2}-2N\omega^{p-1}.

It is easy to see that choosing such α\alpha and β\beta would simultaneously maximize the Euclidean norm αβE2\|\alpha-\beta\|_{E}^{2} and minimize the trace term (TrK/(αβ))2(\operatorname{Tr}_{K/\mathbb{Q}}(\alpha-\beta))^{2}, because in this case TrK/(αβ)=0\operatorname{Tr}_{K/\mathbb{Q}}(\alpha-\beta)=0. Therefore, the maximum distance must be achieved by this pair. It follows from Lemma 4.1 that

(diamB(p,N))2=αβ2=p2(p1)(2N)2,(\operatorname{diam}B(p,N))^{2}=\|\alpha-\beta\|^{2}=p^{2}(p-1)(2N)^{2},

as required. ∎

Definition 5.2.

For α,βB(p,N)\alpha,\beta\in B(p,N), we define the normalized distance of α\alpha and β\beta in the cube by

𝔡p,N(α,β)=d(α,β)2Npp1.\mathfrak{d}_{p,N}(\alpha,\beta)=\frac{d(\alpha,\beta)}{2Np\sqrt{p-1}}.

If we normalize the metric in this way, then the diameter of the hypercube B(p,N)B(p,N) is exactly 1. This normalized distance is not only more aesthetically appealing but also very useful in comparing the distribution of points in different hypercubes B(p,N)B(p,N), as pp and NN vary.

6. Almost all points in B(p,N)B(p,N) are almost equi-distanced

In this section we show that, in an appropriate sense, almost all points in B(p,N)B(p,N) are “equi-distanced” from each other in the sense of Theorem 1.1. Our proof replies on the explicit calculations of the second and fourth moments of the distances, which we define below.

Definition 6.1.

Fix p,Np,N, and let kk be a positive integer. We define the kkth moment of distances between points in B(p,N)B(p,N) to be the following averaged sum:

Mk(p,N):=1#B(p,N)2αB(p,N)βB(p,N)d(α,β)k.M_{k}(p,N):=\frac{1}{\#B(p,N)^{2}}\sum_{\alpha\in B(p,N)}\sum_{\beta\in B(p,N)}d(\alpha,\beta)^{k}.

6.1. Computation of the second moment

Now, we evaluate the second moment of the distances in the following lemmas.

Lemma 6.2.

For integers r0r\geq 0 and N1N\geq 1, consider the sum of powers

Sr(N):=NaNar.S_{r}(N):=\sum_{-N\leq a\leq N}a^{r}.

Then we have:

Sr(N)=0, if r is odd,S2(N)=13N(N+1)(2N+1),S4(N)=115N(N+1)(2N+1)(3N2+3N1).\displaystyle\begin{split}S_{r}(N)&=0,\quad\text{ if $r$ is odd,}\\ S_{2}(N)&=\mbox{\small$\displaystyle\frac{1}{3}$}N(N+1)(2N+1),\\ S_{4}(N)&=\mbox{\small$\displaystyle\frac{1}{15}$}N(N+1)(2N+1)(3N^{2}+3N-1).\end{split}
Proof.

When rr is odd, the sum is zero because ar+(a)r=0a^{r}+(-a)^{r}=0. When rr is even, this follows from the well-known Faulhaber’s formula of sums of powers (see [3] for example). ∎

Lemma 6.3.

The second moment of distances between points in B(p,N)B(p,N) is given by

M2(p,N)\displaystyle M_{2}(p,N) =23(p32p2+1)N(N+1)\displaystyle=\frac{2}{3}(p^{3}-2p^{2}+1)N(N+1)
=23p3N2+O(p2N2+p3N).\displaystyle=\frac{2}{3}p^{3}N^{2}+O(p^{2}N^{2}+p^{3}N).
Proof.

By Lemma 4.1, we have

αB(p,N)βB(p,N)d(α,β)2\displaystyle\sum_{\alpha\in B(p,N)}\sum_{\beta\in B(p,N)}d(\alpha,\beta)^{2}
=Na1,b1NNap1,bp1N(p2i=1p1(aibi)2(p+1)i=1p1j=1p1(aibi)(ajbj)).\displaystyle=\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\left(p^{2}\sum_{i=1}^{p-1}(a_{i}-b_{i})^{2}-(p+1)\sum_{i=1}^{p-1}\sum_{j=1}^{p-1}(a_{i}-b_{i})(a_{j}-b_{j})\right).

We break this sum into two pieces by linearity. The first piece equals

(3) p2i=1p1Na1,b1NNap1,bp1N(aibi)2.p^{2}\sum_{i=1}^{p-1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}(a_{i}-b_{i})^{2}.

The second piece equals

(p+1)i=1p1j=1p1Na1,b1NNap1,bp1N(aibi)(ajbj).(p+1)\sum_{i=1}^{p-1}\sum_{j=1}^{p-1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}(a_{i}-b_{i})(a_{j}-b_{j}).

We now simplify the second piece. If iji\neq j, then the terms aibia_{i}-b_{i} and ajbja_{j}-b_{j} are independent, in which case the sum is zero because

(4) Nai,biN(aibi)=0.\sum_{-N\leq a_{i},b_{i}\leq N}(a_{i}-b_{i})=0.

If i=ji=j, then (aibi)(ajbj)=(aibi)2(a_{i}-b_{i})(a_{j}-b_{j})=(a_{i}-b_{i})^{2}, in which case the sum becomes

(p+1)i=1p1Na1,b1NNap1,bp1N(aibi)2,(p+1)\sum_{i=1}^{p-1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}(a_{i}-b_{i})^{2},

which is exactly the same as (3), up to a difference in the coefficient. It follows that

αB(p,N)βB(p,N)d(α,β)2\displaystyle\sum_{\alpha\in B(p,N)}\sum_{\beta\in B(p,N)}d(\alpha,\beta)^{2}
=(p2p1)i=1p1Na1,b1NNap1,bp1N(aibi)2\displaystyle=(p^{2}-p-1)\sum_{i=1}^{p-1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}(a_{i}-b_{i})^{2}
=(p2p1)(p1)(2N+1)2p4Na1,b1N(a122a1b1+b12).\displaystyle=(p^{2}-p-1)(p-1)(2N+1)^{2p-4}\sum_{-N\leq a_{1},\leq b_{1}\leq N}(a_{1}^{2}-2a_{1}b_{1}+b_{1}^{2}).

Again, we break the above sum by linearity, and noting that

(5) Na1,b1Na1b1=(Na1Na1)(Nb1Nb1)=0,\sum_{-N\leq a_{1},b_{1}\leq N}a_{1}b_{1}=\bigg{(}\sum_{-N\leq a_{1}\leq N}a_{1}\bigg{)}\bigg{(}\sum_{-N\leq b_{1}\leq N}b_{1}\bigg{)}=0,

and

(6) Na1,b1N(a12+b12)=2(2N+1)NaiNai2=2(2N+1)S2(N),\sum_{-N\leq a_{1},b_{1}\leq N}(a_{1}^{2}+b_{1}^{2})=2(2N+1)\sum_{-N\leq a_{i}\leq N}a_{i}^{2}=2(2N+1)S_{2}(N),

where the value of S2(N)S_{2}(N) is computed in Lemma 6.2. Therefore, we obtain

αB(p,N)βB(p,N)d(α,β)2\displaystyle\sum_{\alpha\in B(p,N)}\sum_{\beta\in B(p,N)}d(\alpha,\beta)^{2} =2(p2p1)(p1)(2N+1)2p313N(N+1)(2N+1)\displaystyle=2(p^{2}-p-1)(p-1)(2N+1)^{2p-3}\cdot\frac{1}{3}N(N+1)(2N+1)
=23(p32p2+1)N(N+1)(2N+1)2p2,\displaystyle=\frac{2}{3}(p^{3}-2p^{2}+1)N(N+1)(2N+1)^{2p-2},

and the result follows from dividing the above quantity by #B(p,N)2=(2N+1)2p2\#B(p,N)^{2}=(2N+1)^{2p-2}. ∎

We will argue that almost all pairs of points (α,β)B(p,N)2(\alpha,\beta)\in B(p,N)^{2} are almost μ\sqrt{\mu} away from each other, where

(7) μ=μ(p,N)=23p3N2\mu=\mu(p,N)=\frac{2}{3}p^{3}N^{2}

is exactly the main term appearing in the expression in Lemma 6.3. To this end, we shall need to compute the fourth moment M4(p,N)M_{4}(p,N).

6.2. Computation of the fourth moment

The following lemma will be used several times in the evaluation of M4(p,N)M_{4}(p,N), so we prove it here explicitly.

Lemma 6.4.

We have

(8) i=1p1j=1p1Na1,b1NNap1,bp1N(aibi)2(ajbj)2\displaystyle\sum^{p-1}_{i=1}\sum^{p-1}_{j=1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\left(a_{i}-b_{i}\right)^{2}\left(a_{j}-b_{j}\right)^{2}
=245N(N+1)(p1)(10N2p+4N2+10Np+4N3).\displaystyle\quad=\frac{2}{45}N(N+1)(p-1)\left(10N^{2}p+4N^{2}+10Np+4N-3\right).
Proof.

We break (8) into two pieces according to whether ii equals jj. The i=ji=j piece equals:

(9) (2N+1)2p4(p1)Na1,b1N(a1b1)4.\displaystyle\left(2N+1\right)^{2p-4}\left(p-1\right)\sum_{-N\leq a_{1},b_{1}\leq N}\left(a_{1}-b_{1}\right)^{4}.

Since (a1b1)4=(a14+b14)+4(a13b+a1b13)+6a12b12(a_{1}-b_{1})^{4}=(a_{1}^{4}+b_{1}^{4})+4(a_{1}^{3}b+a_{1}b_{1}^{3})+6a_{1}^{2}b_{1}^{2}, we can further rewrite the sum in (9) as

2(2N+1)Na1Na14+6(Na1Na12)2,\displaystyle 2(2N+1)\sum_{-N\leq a_{1}\leq N}a^{4}_{1}+6\bigg{(}\sum_{-N\leq a_{1}\leq N}a^{2}_{1}\bigg{)}^{2},

since all term with odd powers vanish. The above quantity can be computed directly using Lemma 6.2.

On the other hand, the iji\neq j piece of (8) equals

(10) (2N+1)2p6(p1)(p2)(Na1,b1N(a1b1)2)2.\left(2N+1\right)^{2p-6}\left(p-1\right)\left(p-2\right)\bigg{(}\sum_{-N\leq a_{1},b_{1}\leq N}\left(a_{1}-b_{1}\right)^{2}\bigg{)}^{2}.

The innermost sum inside the square has been previously calculated in (5) and (6). The result now follows from combining the i=ji=j piece and the iji\neq j piece. We omit the details of the tedious calculation. ∎

Lemma 6.5.

The fourth moment of distances between points in B(p,N)B(p,N) is given by

M4(p,N)\displaystyle M_{4}(p,N) =245N(N+1)(p1)((2N2+2N)(5p58p4+p3+8p221p18)3(p2p1)2)\displaystyle=\frac{2}{45}N(N+1)(p-1)((2N^{2}+2N)\left(5p^{5}-8p^{4}+p^{3}+8p^{2}-21p-18\right)-3(p^{2}-p-1)^{2})
=49p6N4+O(p5N4+p6N3).\displaystyle=\frac{4}{9}p^{6}N^{4}+O(p^{5}N^{4}+p^{6}N^{3}).
Proof.

By Lemma 4.1, we need to compute

αB(p,N)βB(p,N)d(α,β)4\displaystyle\sum_{\alpha\in B(p,N)}\sum_{\beta\in B(p,N)}d(\alpha,\beta)^{4} =Na1,b1NNap1,bp1N(p4(j=1p1(ajbj)2)2\displaystyle=\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\Bigg{(}p^{4}\bigg{(}\sum^{p-1}_{j=1}\left(a_{j}-b_{j}\right)^{2}\bigg{)}^{2}
2p2(p+1)(i=1p1(aibi)2)(j=1p1(ajbj))2+(p+1)2(j=1p1(ajbj))4).\displaystyle-2p^{2}\big{(}p+1\big{)}\bigg{(}\sum^{p-1}_{i=1}\left(a_{i}-b_{i}\right)^{2}\bigg{)}\bigg{(}\sum^{p-1}_{j=1}\left(a_{j}-b_{j}\right)\bigg{)}^{2}+\big{(}p+1\big{)}^{2}\bigg{(}\sum^{p-1}_{j=1}\left(a_{j}-b_{j}\right)\bigg{)}^{4}\Bigg{)}.

By linearity, similar to the proof of Lemma 6.3, we break up the sum above into three pieces. The first piece of the sum equals

(11) p4i=1p1j=1p1Na1,b1NNap1,bp1N(aibi)2(ajbj)2.\displaystyle p^{4}\sum^{p-1}_{i=1}\sum^{p-1}_{j=1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\left(a_{i}-b_{i}\right)^{2}\left(a_{j}-b_{j}\right)^{2}.

Observe that this has been calculated in Lemma 6.4

We now evaluate the second piece of the sum, which is

(12) 2p2(p+1)Na1,b1NNap1,bp1Ni=1p1(aibi)2(i=1p1(aibi))2.\displaystyle 2p^{2}\left(p+1\right)\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\sum^{p-1}_{i=1}\left(a_{i}-b_{i}\right)^{2}\bigg{(}\sum^{p-1}_{i=1}\left(a_{i}-b_{i}\right)\bigg{)}^{2}.

Omitting the coefficient 2p2(p+1)2p^{2}(p+1), (12) equals

i=1p1j=1p1k=1p1Na1,b1NNap1,bp1N(aibi)2(ajbj)(akbk)\displaystyle\sum^{p-1}_{i=1}\sum^{p-1}_{j=1}\sum^{p-1}_{k=1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\left(a_{i}-b_{i}\right)^{2}\left(a_{j}-b_{j}\right)\left(a_{k}-b_{k}\right)
=i=1p1j=1p1Na1,b1NNap1,bp1N(aibi)2(ajbj)2,\displaystyle=\sum^{p-1}_{i=1}\sum^{p-1}_{j=1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\left(a_{i}-b_{i}\right)^{2}\left(a_{j}-b_{j}\right)^{2},

because when jkj\neq k, the sum vanishes as in (4). Hence, (12) is the same as (11), up to a constant multiple, so it can also be calculated using Lemma 6.4.

The third piece of the sum is

(13) (p+1)2Na1,b1NNap1,bp1N(j=1p1(ajbj))4\displaystyle\left(p+1\right)^{2}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}\bigg{(}\sum^{p-1}_{j=1}\left(a_{j}-b_{j}\right)\bigg{)}^{4}
=(p+1)2i=1p1j=1p1k=1p1l=1p1Na1,b1NNap1,bp1N(aibi)(ajbj)(akbk)(albl)\displaystyle=(p+1)^{2}\sum_{i=1}^{p-1}\sum_{j=1}^{p-1}\sum_{k=1}^{p-1}\sum_{l=1}^{p-1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}(a_{i}-b_{i})(a_{j}-b_{j})(a_{k}-b_{k})(a_{l}-b_{l})

Depending on the relations between i,j,k,i,j,k, and ll, the above sum can be split into pieces that correspond to the set of all integer partitions of 4. For example, if i=j=kli=j=k\neq l, then the partition is 4=3+14=3+1; if i=jk=li=j\neq k=l, then the partition is 4=2+24=2+2. Now, observe that if the partition has an odd number in it (which is either 1 or 3 in this case), then the sum must vanish because

Nai,biN(aibi)=Nai,biN(aibi)3=0.\displaystyle\sum_{-N\leq a_{i},b_{i}\leq N}\left(a_{i}-b_{i}\right)=\sum_{-N\leq a_{i},b_{i}\leq N}\left(a_{i}-b_{i}\right)^{3}=0.

Hence, only the partitions 4=44=4 and 4=2+24=2+2 result in nonzero summands. Therefore, (13) equals (omitting the coefficient (p+1)2(p+1)^{2})

i=1p1Na1,b1NNap1,bp1N(aibi)4\displaystyle\sum_{i=1}^{p-1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}(a_{i}-b_{i})^{4}
+(42)2i=1p1j=1jip1Na1,b1NNap1,bp1N(aibi)2(ajbj)2\displaystyle+\frac{\binom{4}{2}}{2}\sum_{i=1}^{p-1}\sum_{\begin{subarray}{c}j=1\\ j\neq i\end{subarray}}^{p-1}\sum_{-N\leq a_{1},b_{1}\leq N}\cdots\sum_{-N\leq a_{p-1},b_{p-1}\leq N}(a_{i}-b_{i})^{2}(a_{j}-b_{j})^{2}

which can be further simplified to

(p1)(2N+1)2p4Na1,b1N(a1b1)4\displaystyle(p-1)(2N+1)^{2p-4}\sum_{-N\leq a_{1},b_{1}\leq N}(a_{1}-b_{1})^{4}
+3(p1)(p2)(2N+1)2p6(Na1,b1N(a1b1)2)2.\displaystyle+3(p-1)(p-2)(2N+1)^{2p-6}\bigg{(}\sum_{-N\leq a_{1},b_{1}\leq N}(a_{1}-b_{1})^{2}\bigg{)}^{2}.

We recognize that these two smaller sums have above been previously calculated in the two subcases of Lemma 6.4 (see equations (9) and (10), respectively). Again, we omit some details of the tedious calculation.

Now, combining these three pieces gives the total sum in the lemma, and dividing the quantity by #B(p,N)2=(2N+1)2p2\#B(p,N)^{2}=(2N+1)^{2p-2} yields the result. ∎

Remark 6.6.

We shall never appeal to the first explicit formula of M4(p,N)M_{4}(p,N) in Lemma 6.5. Rather, the second asymptotic estimate of M4(p,N)M_{4}(p,N) will be much more useful in the following analyses.

6.3. Computation of the second moment about the mean

In this subsection, we apply Lemmas 6.3 and 6.5 to obtain an estimate of the second moment of distances about the mean between points in B(p,N)B(p,N), which is formally defined by

R(p,N):=1#B(p,N)2αB(p,N)βB(p,N)(d(α,β)2μ)2,R(p,N):=\frac{1}{\#B(p,N)^{2}}\sum_{\alpha\in B(p,N)}\sum_{\beta\in B(p,N)}(d(\alpha,\beta)^{2}-\mu)^{2},

where μ\mu is defined by (7). R(p,N)R(p,N) will play a crucial role in the proof of our main theorem, and following lemma establishes an upper bound of this quantity.

Lemma 6.7.

We have

R(p,N)p5N4+p6N3.R(p,N)\ll p^{5}N^{4}+p^{6}N^{3}.
Proof.

Indeed, we have

α,β(d(α,β)2μ)2\displaystyle\sum_{\alpha,\beta}(d(\alpha,\beta)^{2}-\mu)^{2} =α,βd(α,β)42μα,βd(α,β)2+μ2(2N+1)2p2.\displaystyle=\sum_{\alpha,\beta}d(\alpha,\beta)^{4}-2\mu\sum_{\alpha,\beta}d(\alpha,\beta)^{2}+\mu^{2}(2N+1)^{2p-2}.

By Lemmas 6.3 and 6.5, we have

R(p,N)\displaystyle R(p,N) =49p6N4+O(p5N4+p6N3)2(23p3N2)(23p3N2+O(p2N2+p3N))+(23p3N2)2\displaystyle=\frac{4}{9}p^{6}N^{4}+O(p^{5}N^{4}+p^{6}N^{3})-2\cdot\left(\frac{2}{3}p^{3}N^{2}\right)\cdot\left(\frac{2}{3}p^{3}N^{2}+O(p^{2}N^{2}+p^{3}N)\right)+\left(\frac{2}{3}p^{3}N^{2}\right)^{2}
p5N4+p6N3,\displaystyle\ll p^{5}N^{4}+p^{6}N^{3},

where the main terms cancel nicely, leaving us with only the big-O term. ∎

6.4. Proof of Theorem 1.1

Our main result Theorem 1.1 now follows immediately from the following quantitative estimate in Theorem 6.8. Note that, instead of normalizing the distance by a factor of 2Npp12Np\sqrt{p-1}, we chose to normalize it by 2Np3/22Np^{3/2} in Theorem 6.8. This choice makes the computations much cleaner, and it will not at all affect the end result since 2Npp12Np\sqrt{p-1} and 2Np3/22Np^{3/2} are asymptotic as pp\to\infty.

Theorem 6.8.

For any ε>0\varepsilon>0 and any positive positive integer NN,

(14) 1#B(p,N)2#{(α,β)B(p,N)2:|d(α,β)2Np3/216|>ε}1ε2(1p+1N),\frac{1}{\#B(p,N)^{2}}\#\left\{(\alpha,\beta)\in B(p,N)^{2}:\left|\frac{d(\alpha,\beta)}{2Np^{3/2}}-\frac{1}{\sqrt{6}}\right|>\varepsilon\right\}\ll\frac{1}{\varepsilon^{2}}\left(\frac{1}{p}+\frac{1}{N}\right),

where the implied constant is absolute and effectively computable.

Proof.

Multiplying both sides of the required inequality in (14) by 2Np3/22Np^{3/2} gives us

|d(α,β)23Np3/2|>ε2Np3/2,\displaystyle\displaystyle\left|d(\alpha,\beta)-\sqrt{\frac{2}{3}}Np^{3/2}\right|>\varepsilon\cdot 2Np^{3/2},

which may be rewritten as

|d(α,β)μ|>ε6μ.\displaystyle\displaystyle\left|d(\alpha,\beta)-\sqrt{\mu}\right|>\varepsilon\sqrt{6\mu}.

Also note that

|d(α,β)2μ|=|d(α,β)+μ||d(α,β)μ|μ|d(α,β)μ|.\left|d(\alpha,\beta)^{2}-\mu\right|=\left|d(\alpha,\beta)+\sqrt{\mu}\right|\left|d(\alpha,\beta)-\sqrt{\mu}\right|\geq\sqrt{\mu}\left|d(\alpha,\beta)-\sqrt{\mu}\right|.

Therefore, by Lemma 6.7,

p5N4+p6N3\displaystyle p^{5}N^{4}+p^{6}N^{3} 1#B(p,N)2α,β(d(α,β)2μ)2\displaystyle\gg\frac{1}{\#B(p,N)^{2}}\sum_{\alpha,\beta}(d(\alpha,\beta)^{2}-\mu)^{2}
1#B(p,N)2α,β|d(α,β)μ|>ε6μ(d(α,β)2μ)2\displaystyle\gg\frac{1}{\#B(p,N)^{2}}\sum_{\begin{subarray}{c}\alpha,\beta\\ \left|d(\alpha,\beta)-\sqrt{\mu}\right|>\varepsilon\sqrt{6\mu}\end{subarray}}(d(\alpha,\beta)^{2}-\mu)^{2}
#{(α,β)B(p,N)2:|d(α,β)μ|>ε6μ}#B(p,N)2(ε6μμ)2\displaystyle\gg\frac{\#\left\{(\alpha,\beta)\in B(p,N)^{2}:\left|d(\alpha,\beta)-\sqrt{\mu}\right|>\varepsilon\sqrt{6\mu}\right\}}{\#B(p,N)^{2}}(\varepsilon\sqrt{6\mu}\cdot\sqrt{\mu})^{2}
=#{(α,β)B(p,N)2:|d(α,β)/(2Np3/2)1/6|}#B(p,N)26ε2μ2.\displaystyle=\frac{\#\left\{(\alpha,\beta)\in B(p,N)^{2}:\left|d(\alpha,\beta)/(2Np^{3/2})-1/\sqrt{6}\right|\right\}}{\#B(p,N)^{2}}6\varepsilon^{2}\mu^{2}.

Therefore, dividing both sides by 6ε2μ26\varepsilon^{2}\mu^{2}, we have

1#B(p,N)2#{(α,β)B(p,N)2:|d(α,β)2Np3/216|>ε}16ε2p5N4+p6N3μ21ε2(1p+1N),\frac{1}{\#B(p,N)^{2}}\#\left\{(\alpha,\beta)\in B(p,N)^{2}:\left|\frac{d(\alpha,\beta)}{2Np^{3/2}}-\frac{1}{\sqrt{6}}\right|>\varepsilon\right\}\ll\frac{1}{6\varepsilon^{2}}\frac{p^{5}N^{4}+p^{6}N^{3}}{\mu^{2}}\ll\frac{1}{\varepsilon^{2}}\left(\frac{1}{p}+\frac{1}{N}\right),

where the implied constant is absolute. ∎

References

  • [1] J. Anderson, C. Cobeli, and A. Zaharescu. Counterintuitive patterns on angles and distances between lattice points in high dimensional hypercubes. Results Math., 79(2):Paper No. 94, 20, 2024. doi:10.1007/s00025-024-02126-2.
  • [2] J.S. Athreya, C. Cobeli, and A. Zaharescu. Visibility phenomena in hypercubes. Chaos Solitons Fractals, 175:Paper No. 114024, 12, 2023. doi:10.1016/j.chaos.2023.114024.
  • [3] D.E. Knuth. Johann Faulhaber and sums of powers. Math. Comp., 61(203):277–294, 1993. doi:10.2307/2152953.
  • [4] A. Malik, F. Stan, and A. Zaharescu. The Siegel norm, the length function and character values of finite groups. Indag. Math. (N.S.), 25(3):475–486, 2014. doi:10.1016/j.indag.2013.12.001.
  • [5] D.A. Marcus. Number Fields. Universitext. Springer New York, 2018. doi:10.1007/978-3-319-90233-3.
  • [6] J. Neukirch and N. Schappacher. Algebraic Number Theory. Grundlehren der mathematischen Wissenschaften. Springer Berlin Heidelberg, 1999. doi:10.1007/978-3-662-03983-0.
  • [7] J. Neukirch, A. Schmidt, and K. Wingberg. Cohomology of Number Fields. Grundlehren der mathematischen Wissenschaften. Springer Berlin Heidelberg, 2008. doi:10.1007/978-3-540-37889-1.
  • [8] C.L. Siegel. The trace of totally positive and real algebraic integers. Ann. of Math. (2), 46:302–312, 1945. doi:10.2307/1969025.
  • [9] F. Stan and A. Zaharescu. Siegel’s trace problem and character values of finite groups. J. Reine Angew. Math., 637:217–234, 2009. doi:10.1515/CRELLE.2009.097.
  • [10] Florin Stan and Alexandru Zaharescu. The Siegel norm of algebraic numbers. Bull. Math. Soc. Sci. Math. Roumanie (N.S.), 55(103)(1):69–77, 2012. URL: https://www.jstor.org/stable/43679241.
  • [11] L.C. Washington. Introduction to Cyclotomic Fields. Graduate Texts in Mathematics. Springer New York, 1997. doi:10.1007/978-1-4612-1934-7.