SINGULAR VECTORS AND -DIRICHLET NUMBERS over function field
Abstract.
We show that the only -Dirichlet numbers in a function field over a finite field are rational functions, unlike -Dirichlet numbers in . We also prove that there are uncountably many totally irrational singular vectors with large uniform exponent in quadratic surfaces over a positive characteristic field.
1. Introduction
1.1. -Dirichlet numbers
Following [17], we define -Dirichlet vectors in and we denote the set of those vectors as . For the definitions of norms in , readers are referred to §1.4.
Definition 1.1.
Let be a function. A vector is said to be -Dirichlet if for all sufficiently large there exists , satisfying the following system
(1.1) | ||||
Let . If is -Dirichlet for every , then is called singular vector.
In recent years, -Dirichlet vectors were studied in [17, 16, 15]. Even in the classical setting not much is known.
Diophantine approximation in function field has been a topic of interest since the work of Artin, [3], which developed the theory of continued fraction, and followed by Mahler’s work in [20], which studied geometry of numbers in function field. For recent developments, we refer readers the survey [19], and to [9, 6, 18, 12, 13, 4, 1, 2] for a necessarily incomplete set of references. There are many interesting similarities and contrasts between the theory of Diophantine approximation over the real numbers and in function field over finite fields. The main theorems in this paper show both of these features.
In [16], for a non-increasing function with non-decreasing, it was shown that in has zero-one law for Lebesgue measure depending on divergence or convergence of certain series involving . Surprisingly, the same is not true over function field as we prove the following.
Theorem 1.1.
Let be non-increasing. If for sufficiently large , then .
1.2. Plenty of singular vectors
The second part of this paper deals with singular vectors in submanifolds of function fields. Note that, if belongs to a rational affine hyperplane, then it must be singular. These are the most trivial singular vectors. In fact, the converse is also true when (ref.[8]). So, we have the following definition to find vectors those can not be singular in a trivial manner.
Definition 1.2.
We call a vector totally irrational vector if it is not inside a rational affine hyperplane of .
For , in [11] Khintchine showed the existence of infinitely many totally irrational singular vectors in . Moreover, Kleinbock, Moshchevitin and Weiss in [14] showed that for real analytic submanifolds (of dimension greater than ) which are not contained inside a rational affine subspace, there are uncountably many totally irrational singular vectors. In this paper, we prove analogous result for certain submanifolds in .
As a special case of our Theorem 1.3, we prove the following theorem.
Theorem 1.2.
Suppose . Let be an open subset of . We consider of the following two types:
-
•
, where and each is a degree polynomial.
-
•
, where .
Suppose that is not contained inside any affine rational hyperplane, then there exist uncountably many totally irrational singular vectors in .
The main challenge comes from the lack of understanding about intersections of a surface and an affine subspace in the function field setting. Another difficulty comes due to total disconnectedness of function field. For real submanifolds, intersection of a connected analytic surface and an affine subspace is well understood due to [5], §2. Both of these facts were used in [14] in a crucial manner. The proof in [14] relies on understanding how ‘semianalytic’ sets can spilt into connected analytic sets. This becomes difficult in , as the notion of semianalyticity is not well defined due to the lack of order and the space is totally disconnected. That is why we had to tackle case by case and we prove the theorem for a class of submanifolds which is smaller than the class of submanifolds that was taken in [14].
1.3. On uniform exponent
One can define , as follows, which quantifies singularity of a vector.
(1.2) |
Dirichlet’s Theorem (ref. [7]) gives that for all . Our next theorem verifies that for a certain analytic submanifolds in , there are plenty of totally irrational vectors whose exponents are infinity.
Theorem 1.3.
Let . Let be an open subset of . We consider of the following two types:
-
•
, where and each is a degree polynomial.
-
•
where .
Suppose that is not contained inside any rational affine hyperplane, then there exist uncountably many totally irrational in such that .
Remark 1.
-
(1)
In Lemma 3.6, we show that the above theorem is true for some higher dimensional submanifolds.
- (2)
1.4. Norms and topology
In this section and in the following sections, we will use (resp. ) to denote norm in (resp. ), unless otherwise mentioned. Let be a prime and , where and consider the finite field . We consider the integral domain , the set of polynomials with coefficients in . Then we consider the function field . We define a norm on as follow:
for all nonzero . Clearly is a nontrivial, non-archimedian and discrete absolute value in . The completion field of with respect to this absolute value is , i.e. the field of Laurent series over . We will denote the absolute value of by the same notation , is given as follows. Let ,
This clearly extends the absolute value of and moreover, the extension remains non-archimedian and discrete. In the above, we call to be the degree of , . It is obvious that is discrete in . For any , throughout is assumed to be equipped with the supremum norm which is defined as for all , and with the topology induced by this norm. Clearly is discrete in . Since the topology on considered here is the usual product topology on , it follows that is locally compact as is locally compact. Note this construction is similar to . Let be the Haar measure on which takes the value 1 on the closed unit ball .
2. -Dirichlet numbers in function field
2.1. Continued fraction over function field
Suppose where , we call as the integer part of and as the fractional part of . Note that if , and otherwise we have . Also, note that . This observation leads us to construct continued fraction expansion of . An expression of the form where is called a simple continued fraction; see § in [22]. An expression of the form , where is called a finite continued fraction. An element of , can be represented as an unique finite continued fraction. An can be represented as a simple continued fraction in the form of and we call the numbers the convergents of . Note that is increasing as . The relation between two consequitive covergents is given by the following equation:
Hence we have,
In fact, by Equation in [22] we have
(2.1) |
where is a convergent of . We recall the following definition and theorem of best approximation [21], §1.2.
Definition 2.1.
We say a rational is the best approximation to some if for all such that we have .
Theorem 2.1.
Let and let be its convergents. Let with be two relatively prime polynomials. Then is a best approximation to if and only if it is a convergent to .
We want to recall the following Lemma from [16] which was stated for real numbers. The verbatim proof will give the following lemma for function field. The proof uses the fact that convergents are best approximations, which we have by Theorem 2.1. In what follows are convergents of .
Lemma 2.1.
Let be non-increasing. Then is -Dirichlet if and only if for sufficiently large .
2.2. Proof of Theorem 1.1
3. Too many vectors with high uniform exponent
In this section we study totally irrational singular vectors in submanifolds of . In dimension , Theorem in [8] implies that the set of numbers in that are singular is .
In order to state the main theorem of this section, we need to define the irrationality measure function as follows. We follow the definition in [14].
Definition 3.1.
We define to be a proper function if the set is finite for any . For any arbitrary and any , we define the irrationality measure function .
We can now state one of the main theorems of this section.
Theorem 3.1.
Let . Let be an open subset of . We consider of the following two types:
-
•
, where each is a degree polynomial.
-
•
where .
Suppose that is not contained inside any rational affine hyperplane. Then for any proper function and for any non-increasing function , there exist uncountably many totally irrationals such that for all large enough .
As an application of the previous Theorem we get Theorem 1.3.
Let us recall Theorem from [14], which was proved for locally closed subsets of . The same proof verbatim will work for locally closed subsets in . It is noteworthy that the proof follows Khintchine’s argument in [11]. We define , where , and is a primitive vector in .
Theorem 3.2.
Let be a nonempty locally closed subset. Let and be disjoint collections of distinct closed subsets of , each of which is contained in a rational affine hyperplane in , and for each let be a rational affine hyperplane containing , assume the following hold:
-
(a)
-
(b)
For each and each ,
-
(c)
For each , and for any finite subsets of indices , with , we have
-
(d)
is dense in .
Then for arbitrary proper function and for any non-increasing function , there exist uncountably many totally irrationals such that for all large enough .
We will call the property (a), (b), (c), and (d) defined above as “property A”. Let us recall the following Theorem in [10] which we are going to use throughout the rest of this section.
Theorem 3.3.
Let be an arbitrary field and assume that for some , every in is in satisfying and further , in which is the Jacobian. Then there exists a unique with every in satisfying and further .
3.1. When each is a degree polynomial and is of any positive characteristic
Let us consider
where is an open subset of , and
with and not being zero simutaneously for . Let us take to be a rational affine hyperplane in and we assume that is not contained inside . We can define by the linear equation , where is primitive. Note that is given by the solutions to the equation;
(3.1) |
where
We see that is a polynomial of degree less than or equal to . Now note that
(3.2) |
and
(3.3) |
If , then by Theorem 3.3 we get a neighborhood of , where is a -analytic function of . If then locally we can write as a -analytic function of . Hence in order to find out all possible such that there is no neighborhood of that is analytic curve in , we consider the linear system
(3.4) |
The corresponding coefficient matrix of the system is
First note that if are zero, then is an analytic curve as the equation of would be . Therefore one of must be nonzero, and without loss of generality we assume that . Next let us denote,
for . With this setting, we have the following lemma.
Lemma 3.1.
If , then there are at most finitely many points of such that their neighbourhood is not an -analytic curve in .
Proof.
Note that . Hence by the hypothesis, we know that the system has only one solution. This completes the proof.
∎
From the proof above we know that the key is to solve the following equations;
(3.5) | ||||
(3.6) |
(3.7) |
Lemma 3.2.
If
(3.8) |
and , then there are at most finitely many points of that its neighbourhood is not an -analytic curve in .
Proof.
Suppose that there exists no point that satisfies System (3.4), then conclusion of the lemma holds trivially.
Now suppose that there exists a point that satisfies System (3.4). Using Equation (3.8) we have,
(3.9) |
For any ,
(3.10) | ||||
For any point that satisfies System (3.4),
(3.11) | ||||
Suppose that . We know that Equation (3.5) would never be satisfied for any points satisfying the System (3.4). Therefore there is no point in such that its neighborhood is not an -analytic curve.
Now suppose that
(3.12) |
Let and be . Then clearly any point in satisfies (3.6), (3.7) and using (3.12) one can see that any point in also satisfies (3.5). Hence we have . Note here Equations (3.6) and (3.7) are essentially the same.
Now for any point in , we have
The last equality holds because as . Since is already an analytic curve, this completes the proof of the lemma.
∎
Lemma 3.3.
If and , then there are at most finitely many points of that its neighbourhood is not an -analytic curve in .
Proof.
If and there exists one point whose neighborhood is not an -analytic curve, then being nonempty implies , because . This contradicts the standing assumption that is not contained inside .
If , then . Since , and cannot be zero simultaneously, assume without loss of generality that and .
Suppose that there exists no point that satisfies System (3.4), then conclusion of the lemma holds. So let us assume that there exists a point that satisfies System (3.4), then Equation (3.7) gives us . By Equation (3.5), we have
At most two can satisfy the above equation, say they are and respectively. Then
and
are both analytic curves and they are inside .
In addition, since
we know that any point on must have an -coordinate equal to or , which means that it is in or . In other words, . Thus the proof is complete for .
If , then and . If either of the two is nonzero then the conclusion of the lemma holds. Otherwise, for every we have
Since , and cannot be zero simultaneously, assume without loss of generality that . Suppose is not square. Now if we have two points and in , then
The above gives a contradiction to the assumption that is not a square. Hence in this case there could be at max one point in .
Now let us assume for some .
Any point in satisfies the equation .
This is equivalent to , where .
Suppose that is a point in , which implies .
Suppose that
is in . For any , let . Then the corresponding can be given as . Therefore , and this shows that gives an analytic curve.
∎
Proposition 3.1.
Let , where is an open set of , and each is a degree polynomial and be an affine rational hyperplane in . Suppose that is not contained inside then there are at most finitely many points of such that its neighbourhood is not an -analytic curve in .
By the above proposition we have that , where is a finite set of points which do not have an -analytic curve as neighborhood, and is an -analytic curve which is a neighborhood of in . Also note that is open and closed. Since is a second countable space, we know that there exists a countable subcovering , i.e. .
Theorem 3.4.
Let be as in the previous proposition. There exist , , as mentioned in Theorem 3.2, that satisfy property A.
Proof.
Let be the set of affine rational hyperplanes normal to one of -axis or -axis in . By possibly replacing with a smaller restriction on and , we can ensure that for any , is normal to and -axis. Now let us define , which are closed subsets and curves of .
Next we define . For any affine rational hyperplane that has a nonempty intersection with , by proposition 3.1, we have that is union of , excluding finitely many points. Let . Now we want to verify that these collections satisfy four hypotheses of Theorem 3.2.
Property (a) of Theorem 3.2 follows directly from how we defined these sets.
First let us consider those , where , . Now let us consider , where . Since , and for some , is dense in , property (b) follows.
Let and are as in hypothesis (c) of Theorem 3.2. Let
where . Note that each with is either empty or consists of only a single point.
For any , by the equation given above, . We can write is a subset of , and therefore, is a subset of . Hence only finitely many solution is possible and has no interior. The same proof will work when where is normal to -axis.
To verify property (d) of Theorem 3.2 it is enough to observe that looks like
or
where . Clearly they form a dense set in . ∎
3.2. A special case in higher degree
Proposition 3.2.
Let . Let , where and is an open subset in . There are at most finitely many points of such that its neighbourhood is not an -analytic curve in .
Proof.
Without loss of generality let us assume that is nonempty, and that there exists at least one point of whose neighbourhood is not an -analytic curve in . This implies , and we can also assume after normalization. This implies that . Suppose also without loss of generality that , , and . The intersection is given by where
Since is nonempty, must have a -th root, say it is . Therefore, we have that is given by
If has one point which has no neighbourhood that is an -analytic curve, we have . Hence we can write as,
Since is nonempty must have a -th root, say it is . Since , where
we have that is defined by .
We use an induction to derive the desired results. For any satisfying , , and , assume that is given by , where
for some . We have and . By System (3.4), we derive that . Now since
and by our assumption that is nonempty, we know that must have a -th root, say it is , i.e., . Define
Then
This implies that the intersection is given by the equation .
Repeating the steps above we see that the intersection is given by . Also from the induction above, we see that
This tells us that is completely given by the curve
where
∎
The exact same proof as Theorem 3.4 with suitable changes gives the following theorem.
Theorem 3.5.
Let , where , and not all s or s are being zero. Then there exists , , that satisfies property A.
We now have everything we need to prove the main theorem of this section:
3.3. Surface to higher dimensional manifold
The following theorem is somewhat an analogue to Lemma in [14].
Theorem 3.6.
Suppose that for and is the open and closed ball in of radius . Let be a -dimensional submanifold of which is the image of under an immersion . Suppose that is nonplanar, where is the Lebesgue measure in . Then there exists such that the surface is not contained inside an rational affine hyperplane, where , .
Proof.
We will prove by contradiction. Suppose for every there exists a rational affine subspace such that By Baire category theorem there is one such that contains an open ball inside . This contradicts the fact that is nonplanar. ∎
Remark 2.
-
(1)
In the above theorem we don’t need to consider the manifold to be analytic but we need a stronger assumption that is nonplanar as compared to the manifold being not inside an affine hyperplane.
-
(2)
The above theorem shows it is enough to prove Theorem 1.3 for surfaces.
3.4. Product of perfect sets
Let us recall that a subset of is called perfect if it is compact and has no isolated points. If , we can define . The proof of the next proposition will be exactly the same as the proof of Theorem 1.6 in [14].
Proposition 3.3.
Let and let be perfect subsets of such that is dense in and is dense in . Let . Then there exist a collection of , , of that satisfy property A.
Theorem 3.7.
Let and let be perfect subsets of such that is dense in and is dense in . Let . Then there exist uncountably many totally irrational singular vectors in .
Acknowledgements
We thank Ralf Spatzier and Anish Ghosh for many helpful discussions and several helpful remarks which have improved the paper. We also thank Subhajit Jana for several helpful suggestions about the presentation of this paper. The second named author thanks University of Michigan for providing help as this project started as a Research Experience for Undergraduates project in summer 2021.
References
- [1] B. Adamczewski and Y. Bugeaud. On the Littlewood conjecture in fields of power series. In Probability and number theory—Kanazawa 2005, volume 49 of Adv. Stud. Pure Math., pages 1–20. Math. Soc. Japan, Tokyo, 2007.
- [2] F. Adiceam, E. Nesharim, and F. Lunnon. On the t-adic Littlewood conjecture. Duke Mathematical Journal, 170(10):2371 – 2419, 2021.
- [3] E. Artin. Quadratische Körper im Gebiete der höheren Kongruenzen. I. Math. Z., 19(1):153–206, 1924.
- [4] J. Bell and Y. Bugeaud. Mahler’s and Koksma’s classifications in fields of power series. Nagoya Math. J., 246:355–371, 2022.
- [5] E. Bierstone and P. D. Milman. Semianalytic and subanalytic sets. Inst. Hautes Études Sci. Publ. Math., 67:5–42, 1988.
- [6] A. Ganguly. Khintchine’s theorem for affine hyperplanes in positive charateristic. J. Number Theory, 238:17–36, 2022.
- [7] A. Ganguly and A. Ghosh. Dirichlet’s theorem in function fields. Canadian Journal of Mathematics, 69(3):532–547, 2017.
- [8] A. Ganguly and A. Ghosh. Remarks on Diophantine approximation in function fields. Math. Scand., 124(1):5–14, 2019.
- [9] A. Ghosh. Metric Diophantine approximation over a local field of positive characteristic. J. Number Theory, 124(2):454–469, 2007.
- [10] J. ichi Igusa. An Introduction to the Theory of Local Zeta Functions, volume 14 of AMS/IP Studies in Advanced Mathematics. American Mathematical Society, Providence, 2000.
- [11] A. Khintchine. Uber eine klasse linearer diophantische approximationen. Rendiconti Circ. Mat. Soc. Palermo, 50:170–195, 1926.
- [12] D. H. Kim and H. Nakada. Metric inhomogeneous Diophantine approximation on the field of formal Laurent series. Acta Arith., 150(2):129–142, 2011.
- [13] T. Kim, S. Lim, and F. Paulin. On hausdorff dimension in inhomogeneous diophantine approximation over global function fields. https://arxiv.org/abs/2112.04144, 2021.
- [14] D. Kleinbock, N. Moshchevitin, and B. Weiss. Singular vectors on manifolds and fractals. Israel J. Math., 245(2):589–613, 2021.
- [15] D. Kleinbock, A. Strömbergsson, and S. Yu. A measure estimate in geometry of numbers and improvements to Dirichlet’s theorem. Proceedings of London Mathematical Society, (in press), arxiv.org/abs/2108.04638.
- [16] D. Kleinbock and N. Wadleigh. A zero-one law for improvements to Dirichlet’s Theorem. Proc. Amer. Math. Soc., 146(5):1833–1844, 2018.
- [17] D. Kleinbock and N. Wadleigh. An inhomogeneous Dirichlet theorem via shrinking targets. Compos. Math., 155(7):1402–1423, 2019.
- [18] S. Kristensen. On well-approximable matrices over a field of formal series. Math. Proc. Cambridge Philos. Soc., 135(2):255–268, 2003.
- [19] A. Lasjaunias. A survey of Diophantine approximation in fields of power series. Monatsh. Math., 130(3):211–229, 2000.
- [20] K. Mahler. An analogue to Minkowski’s geometry of numbers in a field of series. Ann. of Math. (2), 42:488–522, 1941.
- [21] F. Malagoli. Continued fractions in function fields: polynomial analogues of mcmullen’s and zaremba’s conjectures. https://arxiv.org/abs/1704.02640, 2017.
- [22] W. Schmidt. On continued fractions and diophantine approximation in power series fields. Acta arithmetica, 95(2):139–166, 2000.