On a Special Metric in Cyclotomic Fields
Abstract.
Let be an odd prime, and let be a primitive th root of unity. In this paper, we introduce a metric on the cyclotomic field . 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 behave in a highly uniform way under this metric. More specifically, we prove that for a certain hypercube in centered at the origin, almost all pairs of points in the cube are almost equi-distanced from each other, when and are large enough. When suitably normalized, this distance is exactly .
Key words and phrases:
Cyclotomic fields, Lattice points2020 Mathematics Subject Classification:
Primary 11R18; Secondary 11B99, 11P211. 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 be an odd prime and let be a primitive -th root of unity. The extension of generated by in the field of complex numbers is the -th cyclotomic field . We shall denote by the trace map of the number field (the precise definition of will be reviewed in Section 2).
For , we denote by the vector in whose th component is , for . In this paper, we define , the distance between and in , as the Euclidean distance between the vectors and in . We shall show that is a metric on , where positive-definiteness is the only nontrivial property. Note that is canonically defined and is independent of the choice of .
We aim to investigate this metric from several perspectives. In Section 3, we show that has certain nice properties that are related to the algebraic and number-theoretic structure of . For instance, the metric is invariant under the action of the Galois group . In turn, this gives us an analogy of Krasner’s lemma within the context of cyclotomic fields equipped with the metric .
In Section 4, we derive an explicit formula for the metric in terms of the coordinates under the canonical basis of .
In the rest of the paper, we build on the ideas of [1] and [2] to study the metric from a statistical point of view. More specifically, for a positive integer , we denote by the symmetric box of cyclotomic lattice points:
which lies in the ring of integers . In Section 5, we normalize the metric so that the diameter of is exactly 1 in the sense of metric spaces, i.e., the points furthest apart in are at a distance of exactly from each other. This gives us a scaled distance, denoted by , which serves as a unitary means of comparing the spacing of points in different hypercubes , as and vary. Our main theorem states that points in are almost equi-distanced from each other in the following sense.
Theorem 1.1.
For any , there exists an absolute and effectively computable constant such that if , then
Theorem 1.1 reveals a surprising uniformity in the spacing of points among the high-dimensional lattice points in . 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 .
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 be an odd prime, and let be a primitive th root of unity, say . Let be the th cyclotomic field. It is well known that the Galois group is isomorphic to the group , which is cyclic of order .
We denote by the ring of integers of , that is, the integral closure of in . It is well known that rings of integers have integral bases, and in this case, an integral basis of is given by . Therefore,
Many key properties of number fields can be studied via the trace map. Since cyclotomic fields are always Galois over , the trace map has a simple definition in this case, which is
(1) |
It can be proved that for all . Furthermore, if , then .
Finally, for complex-valued functions and , we write or to indicate that there exists an absolute an effectively computable constant such that for all inputs.
2.2. Definition of the metric
We now formally define the metric mentioned in the introduction. The metric is, in fact, induced by a norm on as a -vector space. The norm is defined as follows:
Definition 2.1.
For any , we define
where is the vector whose th component is and denotes the usual Euclidean norm on .
Definition 2.2.
For , we define their distance to be .
Theorem 2.3.
The function defined as in Definition 2.1 is a norm on .
Proof.
We verify the three conditions of a norm. The triangle inequality follows immediately from the usual triangle inequality in Euclidean spaces .
For any and , we need to prove that . This follows from the -linearity of trace, since it implies that .
It remains to prove positive-definiteness. Clearly . Suppose . Then is the zero vector. Hence,
for all . Suppose . Then we may write
Therefore,
Taking the trace of both sides, and using the fact that trace is -linear, we have
which is a contradiction. Hence, . ∎
It follows that the function is indeed a metric on . 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 on 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 is invariant under the action of the Galois group . In other words, for any and , we have
Proof.
It suffices to show that the norm in Definition 2.1 is invariant under , i.e., for all and . Suppose , where . Then we have
where the third equality follows from the fact that is invariant under . Since must be coprime to , it follows that is a permutation of . Hence, , as required. ∎
3.2. An analogue of Krasner’s lemma
As a consequence of Proposition 3.1, we now prove that the metric has another surprising property, with which we will draw an analogy between the following Krasner’s lemma.
Theorem 3.2 (Krasner’s lemma).
Let be a complete field with respect to a nonarchimedean valuation and let be an algebraic closure of . Let be separable over and let be the conjugates of over . Suppose that for we have
where denotes the unique extension of the valuation to . Then .
Proof.
See Lemma 8.1.6 in [7] ∎
We now prove the following analogous result.
Theorem 3.3.
Let , where is a primitive th root of unity. Let be an element of and let be the conjugates of over , with . Suppose that for we have
where is the metric in Definition 2.2. Then .
Proof.
Let . By Galois theory, if and only if . Since is a cyclic group of order , the preceding condition is equivalent to dividing , which is then equivalent to dividing . By the tower law, this is equivalent to dividing .
As in the statement, let be the Galois conjugates of over , with . Similarly, let be the Galois conjugates of over , with . Since is Galois, we have and . Therefore, we need to prove that divides under the hypothesis that for all .
Let
Then . For any element , denote by the open ball centered at with radius under the metric . Observe that if are two distinct Galois conjugates of , say and , where , then
It follows that any two distinct conjugates and are at a distance of at least from each other. In particular, the open balls are pairwise disjoint.
We claim that for every there exists an such that . In other words, the balls contain all conjugates of . Indeed, if , then
by Proposition 3.1. Hence, .
Furthermore, we claim that for each ball contains the same number of conjugates of . Indeed, suppose . Then by Proposition 3.1 again, we have
Therefore, if and only if . Since is a bijection, this proves that and contain the same number of Galois conjugates of . This number is nonzero because . Since the balls are disjoint, we conclude that divides , as desired. ∎
Remark 3.4.
The following example illustrates that the constant is optimal, in the sense that any larger constant would make the statement false. Consider , , and . Then only has one Galois conjugate other than itself, namely . A straightforward computation shows that
Therefore,
but is not contained in .
As a simple consequence, we deduce the following corollary, which is reminiscent of the primitive element theorem in field theory.
Corollary 3.5.
Let . Define . Then for all sufficiently large .
Proof.
Clearly , so it suffices to prove the reverse inclusion. Note that
Thus, when is sufficiently large, we would have for all . Theorem 3.3 implies that . In particular, , and so , 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 .
4. Computing the metric in coordinates
In this section, we aim to derive an explicit formula of the metric in terms of the coordinates of under the integral basis . We first note that , and . Therefore, if
then
and for , we have
Therefore,
Hence, we have arrived at the following convenient formula, which we shall frequently use in the later sections:
Lemma 4.1.
Suppose . Then
(2) |
where the Euclidean norm of , i.e., .
Also note that by the Cauchy-Schwarz inequality,
so we conclude that
In other words, the norm of is always larger than or equal to the Euclidean norm of .
5. The normalized distance
Let be the hypercube
Then contains points in total. In this section, we introduce a normalized distance on . In order to do so, we shall need to compute the diameter of the hypercube ). This is done in the following lemma.
Lemma 5.1.
The diameter of , i.e., the maximum distance between two points in , is exactly
which is achieved by the following pairs of points
Proof.
It suffices to maximize equation (2) for , where . Note that
It is easy to see that choosing such and would simultaneously maximize the Euclidean norm and minimize the trace term , because in this case . Therefore, the maximum distance must be achieved by this pair. It follows from Lemma 4.1 that
as required. ∎
Definition 5.2.
For , we define the normalized distance of and in the cube by
If we normalize the metric in this way, then the diameter of the hypercube 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 , as and vary.
6. Almost all points in are almost equi-distanced
In this section we show that, in an appropriate sense, almost all points in 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 , and let be a positive integer. We define the th moment of distances between points in to be the following averaged sum:
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 and , consider the sum of powers
Then we have:
Proof.
When is odd, the sum is zero because . When 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 is given by
Proof.
By Lemma 4.1, we have
We break this sum into two pieces by linearity. The first piece equals
(3) |
The second piece equals
We now simplify the second piece. If , then the terms and are independent, in which case the sum is zero because
(4) |
If , then , in which case the sum becomes
which is exactly the same as (3), up to a difference in the coefficient. It follows that
Again, we break the above sum by linearity, and noting that
(5) |
and
(6) |
where the value of is computed in Lemma 6.2. Therefore, we obtain
and the result follows from dividing the above quantity by . ∎
We will argue that almost all pairs of points are almost away from each other, where
(7) |
is exactly the main term appearing in the expression in Lemma 6.3. To this end, we shall need to compute the fourth moment .
6.2. Computation of the fourth moment
The following lemma will be used several times in the evaluation of , so we prove it here explicitly.
Lemma 6.4.
We have
(8) | |||
Proof.
Lemma 6.5.
The fourth moment of distances between points in is given by
Proof.
By Lemma 4.1, we need to compute
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) |
Observe that this has been calculated in Lemma 6.4
We now evaluate the second piece of the sum, which is
(12) |
Omitting the coefficient , (12) equals
because when , 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) | ||||
Depending on the relations between and , the above sum can be split into pieces that correspond to the set of all integer partitions of 4. For example, if , then the partition is ; if , then the partition is . 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
Hence, only the partitions and result in nonzero summands. Therefore, (13) equals (omitting the coefficient )
which can be further simplified to
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 yields the result. ∎
Remark 6.6.
We shall never appeal to the first explicit formula of in Lemma 6.5. Rather, the second asymptotic estimate of 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 , which is formally defined by
where is defined by (7). 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
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 , we chose to normalize it by in Theorem 6.8. This choice makes the computations much cleaner, and it will not at all affect the end result since and are asymptotic as .
Theorem 6.8.
For any and any positive positive integer ,
(14) |
where the implied constant is absolute and effectively computable.
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.