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

SINGULAR VECTORS AND ψ\psi-DIRICHLET NUMBERS over function field

Shreyasi Datta and Yewei Xu Department of Mathematics, University of Michigan, Ann Arbor, MI 48109-1043 [email protected] [email protected]
Abstract.

We show that the only ψ\psi-Dirichlet numbers in a function field over a finite field are rational functions, unlike ψ\psi-Dirichlet numbers in \mathbb{R}. 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. ψ\psi-Dirichlet numbers

Following [17], we define ψ\psi-Dirichlet vectors in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n} and we denote the set of those vectors as D(ψ)D(\psi). For the definitions of norms in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}, readers are referred to §1.4.

Definition 1.1.

Let ψ:[t0,+)+\psi\ :[t_{0},+\infty)\to\mathbb{R_{+}} be a function. A vector 𝐱=(x1,,xn)𝔽q((T1))n\mathbf{x}=(x_{1},\cdots,x_{n})\in\mathbb{F}_{q}((T^{-1}))^{n} is said to be ψ\psi-Dirichlet if for all sufficiently large Q>0Q>0 there exists 𝟎𝐪𝔽q[T]n\mathbf{0}\neq\mathbf{q}\in\mathbb{F}_{q}[T]^{n}, q0𝔽q[T]q_{0}\in\mathbb{F}_{q}[T] satisfying the following system

(1.1) |𝐪𝐱+q0|<ψ(Q),\displaystyle|\mathbf{q}\cdot\mathbf{x}+q_{0}|<\psi(Q),
𝐪Q.\displaystyle\|\mathbf{q}\|\leq Q.

Let ψc(Q)=cQn\psi_{c}(Q)=\frac{c}{Q^{n}}. If 𝐱𝔽q((T1))n\mathbf{x}\in\mathbb{F}_{q}((T^{-1}))^{n} is ψc\psi_{c}-Dirichlet for every c>0c>0, then 𝐱\mathbf{x} is called singular vector. In recent years, ψ\psi-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 ψ(t)<1t\psi(t)<\frac{1}{t} with ttψ(t)t\to t\psi(t) non-decreasing, it was shown that D(ψ)D(\psi) in \mathbb{R} has zero-one law for Lebesgue measure depending on divergence or convergence of certain series involving ψ\psi. Surprisingly, the same is not true over function field as we prove the following.

Theorem 1.1.

Let ψ:[t0,+)+\psi\ :[t_{0},+\infty)\to\mathbb{R}_{+} be non-increasing. If ψ(t)<1t\psi(t)<\frac{1}{t} for sufficiently large tt, then D(ψ)=𝔽q(T)D(\psi)=\mathbb{F}_{q}(T).

The above theorem shows that analogue of the main theorem in [16] over function field becomes drastically different than the real case. The main tool in proving Theorem 1.1 is the use of continued fraction expansion.

1.2. Plenty of singular vectors

The second part of this paper deals with singular vectors in submanifolds of function fields. Note that, if (x1,,xn)𝔽q((T1))n(x_{1},\cdots,x_{n})\in\mathbb{F}_{q}((T^{-1}))^{n} 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 n=1n=1 (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 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}.

For n>1n>1, in [11] Khintchine showed the existence of infinitely many totally irrational singular vectors in n\mathbb{R}^{n}. Moreover, Kleinbock, Moshchevitin and Weiss in [14] showed that for real analytic submanifolds (of dimension greater than 22) 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 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}.

As a special case of our Theorem 1.3, we prove the following theorem.

Theorem 1.2.

Suppose char(𝔽q((T1)))=p<\text{char}(\mathbb{F}_{q}((T^{-1})))=p<\infty. Let UU be an open subset of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})). We consider SS of the following two types:

  • S={(x,y,p3(x,y),,pn(x,y))x,yU}𝔽q((T1))nS=\{(x,y,p_{3}(x,y),\cdots,p_{n}(x,y))\ \mid\ x,y\in U\}\subset\mathbb{F}_{q}((T^{-1}))^{n}, where and each pi(x,y)p_{i}(x,y) is a degree 22 polynomial.

  • S={(x,y,p(x,y))|x,yU}𝔽q((T1))3S=\{(x,y,p(x,y))~{}|~{}x,y\in U\}\subset\mathbb{F}_{q}((T^{-1}))^{3}, where p(x,y)=i=0maipixpi+j=0nbjpjypjp(x,y)=\sum_{i=0}^{m}a_{i}^{p^{i}}x^{p^{i}}+\sum_{j=0}^{n}b_{j}^{p^{j}}y^{p^{j}}.

Suppose that SS is not contained inside any affine rational hyperplane, then there exist uncountably many totally irrational singular vectors in SS.

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 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}, as the notion of semianalyticity is not well defined due to the lack of order and the space 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) 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 ω^()\hat{\omega}(\cdot), as follows, which quantifies singularity of a vector.

(1.2) ω^(𝐲):=sup{ω| for all large enough Q>0,(q0,𝐪)𝔽q[T]n+1{0} s.t.𝐪𝐲+q01Qω,𝐪Q}\hat{\omega}(\mathbf{y}):=\sup\left\{\omega~{}\left|~{}\text{ for all large enough }Q>0,\exists~{}(q_{0},\mathbf{q})\in\mathbb{F}_{q}[T]^{n+1}\setminus\{0\}\text{ s.t.}\begin{aligned} &\|\mathbf{q}\cdot\mathbf{y}+q_{0}\|\leq\frac{1}{Q^{\omega}},\\ &\|\mathbf{q}\|\leq Q\end{aligned}\right.\right\}

Dirichlet’s Theorem (ref. [7]) gives that ω^(𝐲)n\hat{\omega}(\mathbf{y})\geq n for all 𝐲𝔽q((T1))n\mathbf{y}\in\mathbb{F}_{q}((T^{-1}))^{n}. Our next theorem verifies that for a certain analytic submanifolds in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}, there are plenty of totally irrational vectors whose exponents ω^()\hat{\omega}(\cdot) are infinity.

Theorem 1.3.

Let char(𝔽q((T1)))=p<\text{char}(\mathbb{F}_{q}((T^{-1})))=p<\infty. Let UU be an open subset of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})). We consider SS of the following two types:

  • S={(x,y,p3(x,y),,pn(x,y))x,yU}𝔽q((T1))nS=\{(x,y,p_{3}(x,y),\cdots,p_{n}(x,y))\ \mid\ x,y\in U\}\subset\mathbb{F}_{q}((T^{-1}))^{n}, where and each pi(x,y)p_{i}(x,y) is a degree 22 polynomial.

  • S={(x,y,p(x,y))|x,yU}𝔽q((T1))3S=\{(x,y,p(x,y))~{}|~{}x,y\in U\}\subset\mathbb{F}_{q}((T^{-1}))^{3} where p(x,y)=i=0maipixpi+j=0nbjpjypjp(x,y)=\sum_{i=0}^{m}a_{i}^{p^{i}}x^{p^{i}}+\sum_{j=0}^{n}b_{j}^{p^{j}}y^{p^{j}}.

Suppose that SS is not contained inside any rational affine hyperplane, then there exist uncountably many totally irrational 𝐲\mathbf{y} in SS such that ω^(𝐲)=\hat{\omega}(\mathbf{y})=\infty.

Remark 1.
  1. (1)

    In Lemma 3.6, we show that the above theorem is true for some higher dimensional submanifolds.

  2. (2)

    Theorem 2.42.4 in [8] shows that only Dirichlet improvable numbers in function field are rational functions. Our Theorem 1.1 generalizes the above mentioned result showing that even ψ\psi-Dirichlet numbers are also only rational functions. We note that the technique of [8] is different than ours.

1.4. Norms and topology

In this section and in the following sections, we will use |||\cdot| (resp. \|\cdot\| ) to denote norm in 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) (resp. 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}), unless otherwise mentioned. Let pp be a prime and q:=prq:=p^{r}, where rr\in\mathbb{N} and consider the finite field 𝔽q\mathbb{F}_{q}. We consider the integral domain 𝔽q[T]\mathbb{F}_{q}[T], the set of polynomials with coefficients in 𝔽q\mathbb{F}_{q}. Then we consider the function field 𝔽q(T)\mathbb{F}_{q}(T). We define a norm |||\cdot| on 𝔽q(T)\mathbb{F}_{q}(T) as follow:

|0|:=0;|PQ|:=edegPdegQ|0|:=0;\left|\frac{P}{Q}\right|:=e^{\mathrm{degP}-\mathrm{degQ}}

for all nonzero P,Q𝔽q[T]P,Q\in\mathbb{F}_{q}[T] . Clearly |||\cdot| is a nontrivial, non-archimedian and discrete absolute value in 𝔽q(T)\mathbb{F}_{q}(T). The completion field of 𝔽q(T)\mathbb{F}_{q}(T) with respect to this absolute value is 𝔽q((T1))\mathbb{F}_{q}((T^{-1})), i.e. the field of Laurent series over 𝔽q\mathbb{F}_{q}. We will denote the absolute value of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) by the same notation |||\cdot|, is given as follows. Let a𝔽q((T1))a\in\mathbb{F}_{q}((T^{-1})),

|a|:={0 if a=0,ek0 if a=kk0akTk,k0,ak𝔽q and ak00.|a|:=\left\{\begin{aligned} &0\text{ if }a=0,\\ &e^{k_{0}}\text{ if }a=\sum_{k\leq k_{0}}a_{k}T^{k},k_{0}\in\mathbb{Z},a_{k}\in\mathbb{F}_{q}\text{ and }a_{k_{0}}\neq 0.\end{aligned}\right.

This clearly extends the absolute value |||\cdot| of 𝔽q(T)𝔽q((T1))\mathbb{F}_{q}(T)\to\mathbb{F}_{q}((T^{-1})) and moreover, the extension remains non-archimedian and discrete. In the above, we call k0k_{0} to be the degree of aa, dega\mathrm{deg}~{}a. It is obvious that 𝔽q[T]\mathbb{F}_{q}[T] is discrete in 𝔽q((T1))\mathbb{F}_{q}((T^{-1})). For any nn\in\mathbb{N}, throughout 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n} is assumed to be equipped with the supremum norm which is defined as 𝐱:=max1in|xi|\|\mathbf{x}\|:=\max_{1\leq i\leq n}|x_{i}| for all 𝐱=(x1,x2,,xn)𝔽q((T1))n\mathbf{x}=(x_{1},x_{2},...,x_{n})\in\mathbb{F}_{q}((T^{-1}))^{n} , and with the topology induced by this norm. Clearly 𝔽q[T]n\mathbb{F}_{q}[T]^{n} is discrete in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}. Since the topology on 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n} considered here is the usual product topology on 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}, it follows that 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n} is locally compact as 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) is locally compact. Note this construction 𝔽q[T]𝔽q(T)𝔽q((T1))\mathbb{F}_{q}[T]\subset\mathbb{F}_{q}(T)\subset\mathbb{F}_{q}((T^{-1})) is similar to \mathbb{Z}\subset\mathbb{Q}\subset\mathbb{R}. Let λ\lambda be the Haar measure on 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n} which takes the value 1 on the closed unit ball 𝐱=1\|\mathbf{x}\|=1.

2. ψ\psi-Dirichlet numbers in function field

2.1. Continued fraction over function field

Suppose a=kk0akTk𝔽q((T1))a=\sum_{k\leq k_{0}}a_{k}T^{k}\in\mathbb{F}_{q}((T^{-1})) where ak00a_{k_{0}}\neq 0, we call [a]:=kk00akTk[a]:=\sum_{k\leq k_{0}}^{0}a_{k}T^{k} as the integer part of aa and a=k<0akTk\langle a\rangle=\sum_{k<0}a_{k}T^{k} as the fractional part of aa. Note that |[a]|=ek01|[a]|=e^{k_{0}}\geq 1 if ak00a_{k_{0}}\neq 0, and otherwise we have [a]=0[a]=0. Also, note that |a|1|\langle a\rangle|\leq 1. This observation leads us to construct continued fraction expansion of aa. An expression of the form a0+1a1+1a2+a_{0}+\frac{1}{a_{1}+\frac{1}{a_{2}+\dots}} where a0,a1,a2𝔽q[T]a_{0},a_{1},a_{2}\in\mathbb{F}_{q}[T] is called a simple continued fraction; see §11 in [22]. An expression of the form pnqn=a0+1a1+1a2++1an\frac{p_{n}}{q_{n}}=a_{0}+\frac{1}{a_{1}+\frac{1}{a_{2}+{\dots}_{+\frac{1}{a_{n}}}}}, where pn,qn,a0,a1,,an𝔽q[T]p_{n},q_{n},a_{0},a_{1},\dots,a_{n}\in\mathbb{F}_{q}[T] is called a finite continued fraction. An element of 𝔽q(T)\mathbb{F}_{q}(T), can be represented as an unique finite continued fraction. An α𝔽q((T1))𝔽q[T]\alpha\in\mathbb{F}_{q}((T^{-1}))\setminus\mathbb{F}_{q}[T] can be represented as a simple continued fraction in the form of [a0,a1,a2,][a_{0},a_{1},a_{2},\dots] and we call the numbers pnqn=[a0,a1,,an]\frac{p_{n}}{q_{n}}=[a_{0},a_{1},\dots,a_{n}] the convergents of α\alpha. Note that |qn||q_{n}| is increasing as nn\to\infty. The relation between two consequitive covergents is given by the following equation:

piqi+1pi+1qi=(1)i+1 for i,i2.p_{i}q_{i+1}-p_{i+1}q_{i}=(-1)^{i+1}\text{ for }i\in\mathbb{Z},i\geq-2.

Hence we have,

|pn+1qn+1pnqn|=|pn+1qnpnqn+1qnqn+1|=|±1qnqn+1|=1|qn||qn+1|1|qn|2.\left|\frac{p_{n+1}}{q_{n+1}}-\frac{p_{n}}{q_{n}}\right|=\left|\frac{p_{n+1}q_{n}-p_{n}q_{n+1}}{q_{n}q_{n+1}}\right|=\left|\frac{\pm 1}{q_{n}q_{n+1}}\right|=\frac{1}{|q_{n}|\cdot|q_{n+1}|}\leq\frac{1}{|q_{n}|^{2}}.

In fact, by Equation 1.121.12 in [22] we have

(2.1) |αpnqn|=|1qnqn+1|,\left|\alpha-\frac{p_{n}}{q_{n}}\right|=\left|\frac{1}{q_{n}q_{n+1}}\right|,

where pnqn\frac{p_{n}}{q_{n}} is a convergent of α𝔽q((T1))𝔽q(T)\alpha\in\mathbb{F}_{q}((T^{-1}))\setminus\mathbb{F}_{q}(T). We recall the following definition and theorem of best approximation [21], §1.2.

Definition 2.1.

We say a rational ab\frac{a}{b} is the best approximation to some α𝔽q((T1))\alpha\in\mathbb{F}_{q}((T^{-1})) if for all cd\frac{c}{d} such that |d||b||d|\leq|b| we have |bαa||dαc||b\alpha-a|\leq|d\alpha-c|.

Theorem 2.1.

Let α𝔽q((T1))\alpha\in\mathbb{F}_{q}((T^{-1})) and let (pnqn)n(\frac{p_{n}}{q_{n}})_{n} be its convergents. Let p,q𝔽q[T]p,q\in\mathbb{F}_{q}[T] with q0q\neq 0 be two relatively prime polynomials. Then pq\frac{p}{q} is a best approximation to α\alpha if and only if it is a convergent to α\alpha.

We want to recall the following Lemma 2.12.1 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 pnqn\frac{p_{n}}{q_{n}} are convergents of xx.

Lemma 2.1.

Let ψ:[t0,+)+\psi\ :[t_{0},+\infty)\to\mathbb{R}_{+} be non-increasing. Then x𝔽q((T1))𝔽q(T)x\in\mathbb{F}_{q}((T^{-1}))\setminus\mathbb{F}_{q}(T) is ψ\psi-Dirichlet if and only if |qn1x|<ψ(|qn|)|\langle q_{n-1}x\rangle|<\psi(|q_{n}|) for sufficiently large nn.

2.2. Proof of Theorem 1.1

It is easy to see that 𝔽q(T)D(ψ)\mathbb{F}_{q}(T)\subset D(\psi). We want to show that D(ψ)𝔽q(T)D(\psi)\subset\mathbb{F}_{q}(T). By Equation (2.1) for x𝔽q((T1))𝔽q(T)x\in\mathbb{F}_{q}((T^{-1}))\setminus\mathbb{F}_{q}(T) we have |qn1x|=1|qn|,|\langle q_{n-1}x\rangle|=\frac{1}{|q_{n}|}, n\forall~{}n. Since ψ(t)<1t\psi(t)<\frac{1}{t} for all large enough tt, by Lemma 2.1 we conclude that there is no x𝔽q((T1))𝔽q(T)x\in\mathbb{F}_{q}((T^{-1}))\setminus\mathbb{F}_{q}(T) such that xx is ψ\psi-Dirichlet.

3. Too many vectors with high uniform exponent

In this section we study totally irrational singular vectors in submanifolds of 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}. In dimension n=1n=1, Theorem 2.42.4 in [8] implies that the set of numbers yy in 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) that are singular is 𝔽q(T)\mathbb{F}_{q}(T).

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 Φ:𝔽q[T]n{0}+\Phi:\mathbb{F}_{q}[T]^{n}\setminus\{0\}\to\mathbb{R}_{+} to be a proper function if the set {𝐪𝔽q[T]n{0}:Φ(𝐪)C}\{\mathbf{q}\in\mathbb{F}_{q}[T]^{n}\setminus\{0\}\ :\Phi(\mathbf{q})\leq C\} is finite for any C>0C>0. For any arbitrary Φ\Phi and any 𝐲𝔽q((T1))n\mathbf{y}\in\mathbb{F}_{q}((T^{-1}))^{n}, we define the irrationality measure function ψΦ,𝐲(t):=min(q0,𝐪)𝔽q[T]×𝔽q[T]n{𝟎},Φ(𝐪)t|𝐪𝐲+q0|\psi_{\Phi,\mathbf{y}}(t):=\min_{(q_{0},\mathbf{q})\in\mathbb{F}_{q}[T]\times\mathbb{F}_{q}[T]^{n}\setminus\{\mathbf{0}\},\Phi(\mathbf{q})\leq t}|\mathbf{q}\cdot\mathbf{y}+q_{0}|.

We can now state one of the main theorems of this section.

Theorem 3.1.

Let char(𝔽q((T1)))=p<\text{char}(\mathbb{F}_{q}((T^{-1})))=p<\infty. Let UU be an open subset of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})). We consider SS of the following two types:

  • S={(x,y,p3(x,y),,pn(x,y))x,yU}𝔽q((T1))nS=\{(x,y,p_{3}(x,y),\cdots,p_{n}(x,y))\ \mid\ x,y\in U\}\subset\mathbb{F}_{q}((T^{-1}))^{n}, where each pi(x,y)p_{i}(x,y) is a degree 22 polynomial.

  • S={(x,y,p(x,y))|x,yU}𝔽q((T1))3S=\{(x,y,p(x,y))~{}|~{}x,y\in U\}\subset\mathbb{F}_{q}((T^{-1}))^{3} where p(x,y)=i=0naipixpi+j=0nbjpjypjp(x,y)=\sum_{i=0}^{n}a_{i}^{p^{i}}x^{p^{i}}+\sum_{j=0}^{n}b_{j}^{p^{j}}y^{p^{j}}.

Suppose that SS is not contained inside any rational affine hyperplane. Then for any proper function Φ:𝔽q[T]n{0}+\Phi:\mathbb{F}_{q}[T]^{n}\setminus\{0\}\to\mathbb{R}_{+} and for any non-increasing function ϕ:++\phi\ :\mathbb{R}_{+}\to\mathbb{R}_{+}, there exist uncountably many totally irrationals 𝐲S\mathbf{y}\in S such that ψΦ,𝐲(t)ϕ(t)\psi_{\Phi,\mathbf{y}}(t)\leq\phi(t) for all large enough tt.

As an application of the previous Theorem we get Theorem 1.3.

Proof of Theorem 1.3.

By taking Φ(𝐪)=𝐪\Phi(\mathbf{q})=\|\mathbf{q}\|, Theorem 1.3 follows from Theorem 3.1. ∎

Let us recall Theorem 1.11.1 from [14], which was proved for locally closed subsets of n\mathbb{R}^{n}. The same proof verbatim will work for locally closed subsets in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}. It is noteworthy that the proof follows Khintchine’s argument in [11]. We define |A|:=maxi=1n+1|ai||A|:=\max_{i=1}^{n+1}|a_{i}|, where A:a1x1++anxn=an+1A:a_{1}x_{1}+\cdots+a_{n}x_{n}=a_{n+1}, and (a1,,an+1)(a_{1},\cdots,a_{n+1}) is a primitive vector in 𝔽q[T]n+1\mathbb{F}_{q}[T]^{n+1}.

Theorem 3.2.

Let S𝔽q((T1))nS\subset\mathbb{F}_{q}((T^{-1}))^{n} be a nonempty locally closed subset. Let {L1,L2,}\{L_{1},L_{2},\dots\} and {L1,L2,}\{L_{1}^{\prime},L_{2}^{\prime},\dots\} be disjoint collections of distinct closed subsets of SS, each of which is contained in a rational affine hyperplane in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}, and for each ii let AiA_{i} be a rational affine hyperplane containing LiL_{i}, assume the following hold:

  1. (a)
    iLijLj={xS:x is contained in a rational affine hyperplane};\bigcup_{i}L_{i}\cup\bigcup_{j}L_{j}^{\prime}=\{x\in S\ :x\text{ is contained in a rational affine hyperplane}\};
  2. (b)

    For each ii and each α>0\alpha>0,

    Li=|Aj|>αLiLj¯;L_{i}=\overline{\bigcup_{|A_{j}|>\alpha}L_{i}\cap L_{j}};
  3. (c)

    For each ii, and for any finite subsets of indices FF, FF^{\prime} with iFi\not\in F, we have

    Li=Li(kFLkkFLk)¯;L_{i}=\overline{L_{i}-(\bigcup_{k\in F}L_{k}\cup\bigcup_{k^{\prime}\in F^{\prime}}L_{k^{\prime}}^{\prime})};
  4. (d)

    iLi\bigcup_{i}L_{i} is dense in SS.

Then for arbitrary Φ:𝔽q[T]n{0}+\Phi:\mathbb{F}_{q}[T]^{n}\setminus\{0\}\to\mathbb{R}_{+} proper function and for any non-increasing function ϕ:++\phi\ :\mathbb{R}_{+}\to\mathbb{R}_{+}, there exist uncountably many totally irrationals 𝐲S\mathbf{y}\in S such that ψΦ,𝐲(t)ϕ(t)\psi_{\Phi,\mathbf{y}}(t)\leq\phi(t) for all large enough tt.

We will call the property (a), (b), (c), and (d) defined above as “property A”. Let us recall the following Theorem 2.1.12.1.1 in [10] which we are going to use throughout the rest of this section.

Theorem 3.3.

Let KK be an arbitrary field and assume that for some mm, nn every Fi(x,y)F_{i}(x,y) in F(x,y)=(F1(x,y),,Fm(x,y))F(x,y)=(F_{1}(x,y),\dots,F_{m}(x,y)) is in K[[X,Y]]=K[[x1,,xn,y1,,ym]]K[[X,Y]]=K[[x_{1},\dots,x_{n},y_{1},\dots,y_{m}]] satisfying Fi(0,0)=0F_{i}(0,0)=0 and further (F1,,Fm)(y1,,ym)(0,0)0\frac{\partial(F_{1},\dots,F_{m})}{\partial(y_{1},\dots,y_{m})}\mid_{(0,0)}\neq 0, in which (F1,,Fm)(y1,,ym)\frac{\partial(F_{1},\dots,F_{m})}{\partial(y_{1},\dots,y_{m})} is the Jacobian. Then there exists a unique f(x)=(f1(x),,fm(x))f(x)=(f_{1}(x),\dots,f_{m}(x)) with every fi(x)f_{i}(x) in K[[x]]=K[[x1,,xm]]K[[x]]=K[[x_{1},\dots,x_{m}]] satisfying fi(0)=0f_{i}(0)=0 and further F(x,f(x))=0F(x,f(x))=0.

3.1. When each pi(x,y)p_{i}(x,y) is a degree 22 polynomial and 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) is of any positive characteristic

Let us consider

S={(x,y,p3(x,y),,pn(x,y))x,yU},S=\{(x,y,p_{3}(x,y),\cdots,p_{n}(x,y))\ \mid\ x,y\in U\},

where UU is an open subset of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})), and

pi(x,y)=bi,1x2+bi,2xy+bi,3y2+bi,4x+bi,5y+bi,6p_{i}(x,y)=b_{i,1}x^{2}+b_{i,2}xy+b_{i,3}y^{2}+b_{i,4}x+b_{i,5}y+b_{i,6}

with b1,i,bi,2,,bi,6𝔽q((T1))b_{1,i},b_{i,2},\dots,b_{i,6}\in\mathbb{F}_{q}((T^{-1})) and bi,1,bi,2,bi,3b_{i,1},b_{i,2},b_{i,3} not being zero simutaneously for i=3,,ni=3,\cdots,n. Let us take AA to be a rational affine hyperplane in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n} and we assume that SS is not contained inside AA. We can define AA by the linear equation a1x1+a2x2++anxn=an+1a_{1}x_{1}+a_{2}x_{2}+\cdots+a_{n}x_{n}=a_{n+1}, where (a1,a2,,an+1)𝔽q[T]n+1(a_{1},a_{2},\cdots,a_{n+1})\in\mathbb{F}_{q}[T]^{n+1} is primitive. Note that SAS\cap A is given by the solutions to the equation;

(3.1) f(x,y):=0,f(x,y):=0,

where

f(x,y)=i=3naipi(x,y)+a1x+a2yan+1.f(x,y)=\sum_{i=3}^{n}a_{i}p_{i}(x,y)+a_{1}x+a_{2}y-a_{n+1}.

We see that ff is a polynomial of degree less than or equal to 22. Now note that

(3.2) fx=i=3n(2aibi,1x+aibi,2y)+(a1+i=3naibi,4)\frac{\partial f}{\partial x}=\sum_{i=3}^{n}(2a_{i}b_{i,1}x+a_{i}b_{i,2}y)+(a_{1}+\sum_{i=3}^{n}a_{i}b_{i,4})

and

(3.3) fy=i=3n(aibi,2x+2aibi,3y)+(a2+i=3naibi,5).\frac{\partial f}{\partial y}=\sum_{i=3}^{n}(a_{i}b_{i,2}x+2a_{i}b_{i,3}y)+(a_{2}+\sum_{i=3}^{n}a_{i}b_{i,5}).

If fx(x0,y0)0\frac{\partial f}{\partial x}(x_{0},y_{0})\neq 0, then by Theorem 3.3 we get a neighborhood of (x0,y0)(x_{0},y_{0}), where yy is a 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic function of xx. If fy(x0,y0)0\frac{\partial f}{\partial y}(x_{0},y_{0})\neq 0 then locally we can write xx as a 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic function of yy. Hence in order to find out all possible (x0,y0)(x_{0},y_{0}) such that there is no neighborhood of (x0,y0,p3(x0,y0),,pn(x0,y0))SA(x_{0},y_{0},p_{3}(x_{0},y_{0}),\cdots,p_{n}(x_{0},y_{0}))\in S\cap A that is analytic curve in SAS\cap A, we consider the linear system

(3.4) {fx=0;fy=0.\begin{cases}\frac{\partial f}{\partial x}=0;\\ \frac{\partial f}{\partial y}=0.\end{cases}

The corresponding coefficient matrix MMat2×2(𝔽q((T1)))M\in\text{Mat}_{2\times 2}(\mathbb{F}_{q}((T^{-1}))) of the system is

[i=3n2aibi,1i=3naibi,2i=3naibi,2i=3n2aibi,3].\begin{bmatrix}\sum_{i=3}^{n}2a_{i}b_{i,1}&\sum_{i=3}^{n}a_{i}b_{i,2}\\ \sum_{i=3}^{n}a_{i}b_{i,2}&\sum_{i=3}^{n}2a_{i}b_{i,3}\end{bmatrix}.

First note that if a3,,ana_{3},\cdots,a_{n} are zero, then SAS\cap A is an analytic curve as the equation of AA would be a1x+a2y=an+1a_{1}x+a_{2}y=a_{n+1}. Therefore one of a3,,ana_{3},\cdots,a_{n} must be nonzero, and without loss of generality we assume that a30a_{3}\neq 0. Next let us denote,

bk=i=3naibi,kb_{k}=\sum_{i=3}^{n}a_{i}b_{i,k}

for k=1,,6k=1,\cdots,6. With this setting, we have the following lemma.

Lemma 3.1.

If b224b1b3b_{2}^{2}\neq 4b_{1}b_{3}, then there are at most finitely many points of ASA\cap S such that their neighbourhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}.

Proof.

Note that det(M)=4b1b3b22\det(M)=4{b_{1}}{b_{3}}-b_{2}^{2}. 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) f(x,y)\displaystyle f(x,y) =i=3naipi(x,y)+a1x+a2yan+1\displaystyle=\sum_{i=3}^{n}a_{i}p_{i}(x,y)+a_{1}x+a_{2}y-a_{n+1}
=b1x2+b2xy+b3y2+(a1+b4)x+(a2+b5)y+(b6an+1)=0,\displaystyle=b_{1}x^{2}+b_{2}xy+b_{3}y^{2}+(a_{1}+b_{4})x+(a_{2}+b_{5})y+(b_{6}-a_{n+1})=0,
(3.6) fx(x,y)=2b1x+b2y+(a1+b4)=0,\frac{\partial f}{\partial x}(x,y)=2{b_{1}}x+{b_{2}}y+(a_{1}+{b_{4}})=0,
(3.7) fy(x,y)=b2x+2b3y+(a2+b5)=0.\frac{\partial f}{\partial y}(x,y)=b_{2}x+2b_{3}y+(a_{2}+b_{5})=0.
Lemma 3.2.

If

(3.8) b22=4b1b3b_{2}^{2}=4b_{1}b_{3}

and b20b_{2}\neq 0, then there are at most finitely many points of ASA\cap S that its neighbourhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}.

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 (x0,y0)(x_{0},y_{0}) that satisfies System (3.4). Using Equation (3.8) we have,

(3.9) b2(a2+b5)=(3.7)b22x02b3(b2y0)=(3.6)b22x0+2b3(2b1x0+(a1+b4))=(3.8)2b3(a1+b4).b_{2}(a_{2}+b_{5})\stackrel{{\scriptstyle\eqref{crucial3}}}{{=}}-b_{2}^{2}x_{0}-2b_{3}(b_{2}y_{0})\stackrel{{\scriptstyle\eqref{crucial2}}}{{=}}-b_{2}^{2}x_{0}+2b_{3}(2b_{1}x_{0}+(a_{1}+b_{4}))\stackrel{{\scriptstyle\eqref{w0}}}{{=}}2b_{3}(a_{1}+b_{4}).

For any x,yUx,y\in U,

(3.10) b22f(x,y)\displaystyle b_{2}^{2}f(x,y) =(3.5)b22(b1x2+b2xy+b3y2)+b22((a1+b4)x+(a2+b5)y)+b22(b6an+1)\displaystyle\stackrel{{\scriptstyle\eqref{crucial1}}}{{=}}b_{2}^{2}(b_{1}x^{2}+b_{2}xy+b_{3}y^{2})+b_{2}^{2}((a_{1}+b_{4})x+(a_{2}+b_{5})y)+b_{2}^{2}(b_{6}-a_{n+1})
=(3.8)b3(2b1x+b2y)2+b22((a1+b4)x+(a2+b5)y)+b22(b6an+1)\displaystyle\stackrel{{\scriptstyle\eqref{w0}}}{{=}}b_{3}(2b_{1}x+b_{2}y)^{2}+b_{2}^{2}((a_{1}+b_{4})x+(a_{2}+b_{5})y)+b_{2}^{2}(b_{6}-a_{n+1})
=(3.9)b3(2b1x+b2y)2+b2(a1+b4)(b2x+2b3y)+b22(b6an+1).\displaystyle\stackrel{{\scriptstyle\eqref{w1}}}{{=}}b_{3}(2b_{1}x+b_{2}y)^{2}+b_{2}(a_{1}+b_{4})(b_{2}x+2b_{3}y)+b_{2}^{2}(b_{6}-a_{n+1}).

For any point that satisfies System (3.4),

(3.11) b3(2b1x+b2y)2+b2(a1+b4)(b2x+2b3y)+b22(b6an+1)\displaystyle b_{3}(2b_{1}x+b_{2}y)^{2}+b_{2}(a_{1}+b_{4})(b_{2}x+2b_{3}y)+b_{2}^{2}(b_{6}-a_{n+1})
=(3.6),(3.7)b3(a1+b4)2b2(a1+b4)(a2+b5)+b22(b6an+1)\displaystyle\stackrel{{\scriptstyle\eqref{crucial2},\eqref{crucial3}}}{{=}}b_{3}(a_{1}+b_{4})^{2}-b_{2}(a_{1}+b_{4})(a_{2}+b_{5})+b_{2}^{2}(b_{6}-a_{n+1})
=(3.9)b22(b6an+1)b3(a1+b4)2.\displaystyle\stackrel{{\scriptstyle\eqref{w1}}}{{=}}b_{2}^{2}(b_{6}-a_{n+1})-b_{3}(a_{1}+b_{4})^{2}.

Suppose that b22(b6an+1)b3(a1+b4)2b_{2}^{2}(b_{6}-a_{n+1})\neq b_{3}(a_{1}+b_{4})^{2}. We know that Equation (3.5) would never be satisfied for any points satisfying the System (3.4). Therefore there is no point in SAS\cap A such that its neighborhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve.

Now suppose that

(3.12) b22(b6an+1)=b3(a1+b4)2.b_{2}^{2}(b_{6}-a_{n+1})=b_{3}(a_{1}+b_{4})^{2}.

Let g(x):=2b1x+(a1+b4)b2g(x):=-\frac{2b_{1}x+(a_{1}+b_{4})}{b_{2}} and γ\gamma be {(x,g(x),p3(x,g(x)),,pn(x,g(x)))xU}\{(x,g(x),p_{3}(x,g(x)),\cdots,p_{n}(x,g(x)))\ \mid\ x\in U\}. Then clearly any point in γ\gamma satisfies (3.6), (3.7) and using (3.12) one can see that any point in γ\gamma also satisfies (3.5). Hence we have γSA\gamma\subseteq S\cap A. Note here Equations (3.6) and (3.7) are essentially the same.

Now for any point in SAS\cap A, we have

f(x,y)=0\displaystyle f(x,y)=0
(3.10)b3(2b1x+b2y)2+b2(a1+b4)(b2x+2b3y)+b22(b6an+1)=0\displaystyle~{}\stackrel{{\scriptstyle\eqref{w4}}}{{\implies}}b_{3}(2b_{1}x+b_{2}y)^{2}+b_{2}(a_{1}+b_{4})(b_{2}x+2b_{3}y)+b_{2}^{2}(b_{6}-a_{n+1})=0
(3.12),(3.8)b3(2b1x+b2y+a1+b4)2=0\displaystyle\stackrel{{\scriptstyle\eqref{w3},\eqref{w0}}}{{\implies}}b_{3}(2b_{1}x+b_{2}y+a_{1}+b_{4})^{2}=0
y=2b1x+(a1+b4)b2.\displaystyle~{}\implies y=-\frac{2b_{1}x+(a_{1}+b_{4})}{b_{2}}.

The last equality holds because b30b_{3}\neq 0 as b20b_{2}\neq 0. Since γ\gamma is already an analytic curve, this completes the proof of the lemma.

Lemma 3.3.

If b22=4b1b3b_{2}^{2}=4b_{1}b_{3} and b2=0b_{2}=0, then there are at most finitely many points of ASA\cap S that its neighbourhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}.

Proof.

If b1=0,b2=0,b3=0b_{1}=0,b_{2}=0,b_{3}=0 and there exists one point whose neighborhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve, then SAS\cap A being nonempty implies SA=AS\cap A=A, because f(x,y)=b6an+1=0f(x,y)=b_{6}-a_{n+1}=0. This contradicts the standing assumption that SS is not contained inside AA.
If char(𝔽q((T1)))2\text{char}(\mathbb{F}_{q}((T^{-1})))\neq 2, then b1b3=0b_{1}b_{3}=0. Since b1b_{1}, b2b_{2} and b3b_{3} cannot be zero simultaneously, assume without loss of generality that b10b_{1}\neq 0 and b3=0b_{3}=0.

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 (x0,y0)(x_{0},y_{0}) that satisfies System (3.4), then Equation (3.7) gives us a2+b5=0a_{2}+b_{5}=0. By Equation (3.5), we have

b1x2+(a1+b4)x+(b6an+1)=0.\displaystyle b_{1}x^{2}+(a_{1}+b_{4})x+(b_{6}-a_{n+1})=0.

At most two xx can satisfy the above equation, say they are x1x_{1} and x2x_{2} respectively. Then

γ1={(x1,y,p3(x1,y),,pn(x1,y))yU}\gamma_{1}=\{(x_{1},y,p_{3}(x_{1},y),\cdots,p_{n}(x_{1},y))\ \mid\ y\in U\}

and

γ2={(x2,y,p3(x2,y),,pn(x2,y))yU}\gamma_{2}=\{(x_{2},y,p_{3}(x_{2},y),\cdots,p_{n}(x_{2},y))\ \mid\ y\in U\}

are both analytic curves and they are inside SAS\cap A.

In addition, since

f(x,y)=b1x2+(a1+b4)x+(b6an+1),f(x,y)=b_{1}x^{2}+(a_{1}+b_{4})x+(b_{6}-a_{n+1}),

we know that any point on SAS\cap A must have an xx-coordinate equal to x1x_{1} or x2x_{2}, which means that it is in γ1\gamma_{1} or γ2\gamma_{2}. In other words, SA=γ1γ2S\cap A=\gamma_{1}\sqcup\gamma_{2}. Thus the proof is complete for char(𝔽q((T1)))2\mathrm{char}(\mathbb{F}_{q}((T^{-1})))\neq 2.

If char(𝔽q((T1)))=2\text{char}(\mathbb{F}_{q}((T^{-1})))=2, then fx=a1+b4\frac{\partial f}{\partial x}=a_{1}+b_{4} and fy=a2+b5\frac{\partial f}{\partial y}=a_{2}+b_{5}. If either of the two is nonzero then the conclusion of the lemma holds. Otherwise, for every x,yx,y we have

f(x,y)=b1x2+b3y2+(b6an+1).f(x,y)=b_{1}x^{2}+b_{3}y^{2}+(b_{6}-a_{n+1}).

Since b1b_{1}, b2b_{2} and b3b_{3} cannot be zero simultaneously, assume without loss of generality that b30b_{3}\neq 0. Suppose b1b3\frac{b_{1}}{b_{3}} is not square. Now if we have two points (x0,y0,p3(x0,y0),,pn(x0,y0))(x_{0},y_{0},p_{3}(x_{0},y_{0}),\cdots,p_{n}(x_{0},y_{0})) and (x1,y1,p3(x1,y1),,pn(x1,y1))(x_{1},y_{1},p_{3}(x_{1},y_{1}),\cdots,p_{n}(x_{1},y_{1})) in SAS\cap A, then

b1x02+b3y02=b1x12+b3y12\displaystyle\ \ \ \ \ \ \ b_{1}x_{0}^{2}+b_{3}y_{0}^{2}=b_{1}x_{1}^{2}+b_{3}y_{1}^{2}
b1(x1x0)2=b3(y1y0)2.\displaystyle\implies b_{1}(x_{1}-x_{0})^{2}=b_{3}(y_{1}-y_{0})^{2}.

The above gives a contradiction to the assumption that b1b3\frac{b_{1}}{b_{3}} is not a square. Hence in this case there could be at max one point in SAS\cap A.
Now let us assume b1b3=α2\frac{b_{1}}{b_{3}}=\alpha^{2} for some α𝔽q((T1))\alpha\in\mathbb{F}_{q}((T^{-1})). Any point in SAS\cap A satisfies the equation b1x2+b3y2=(an+1b6)b_{1}x^{2}+b_{3}y^{2}=(a_{n+1}-b_{6}). This is equivalent to α2x2+y2=β\alpha^{2}x^{2}+y^{2}=\beta, where β=an+1b6b3\beta=\frac{a_{n+1}-b_{6}}{b_{3}}. Suppose that (x0,y0,p3(x0,y0),,pn(x0,y0))(x_{0},y_{0},p_{3}(x_{0},y_{0}),\cdots,p_{n}(x_{0},y_{0})) is a point in SAS\cap A, which implies α2x02+y02=β\alpha^{2}x_{0}^{2}+y_{0}^{2}=\beta. Suppose that

(x(t),y(t),p3(x(t),y(t)),,pn(x(t),y(t)))(x(t),y(t),p_{3}(x(t),y(t)),\cdots,p_{n}(x(t),y(t)))

is in SAS\cap A. For any t𝔽q((T1))t\in\mathbb{F}_{q}((T^{-1})), let x(t)=x0+tx(t)=x_{0}+t. Then the corresponding y(t)y(t) can be given as (y(t))2=βα2(x(t))2=(βα2x02)α2t2=y02+α2t2(y(t))^{2}=\beta-\alpha^{2}(x(t))^{2}=(\beta-\alpha^{2}x_{0}^{2})-\alpha^{2}t^{2}=y_{0}^{2}+\alpha^{2}t^{2}. Therefore y(t)=y0+αty(t)=y_{0}+\alpha t, and this shows that SAS\cap A gives an analytic curve.

Combining Lemma 3.1, Lemma 3.2 and Lemma 3.3 we get the following proposition.

Proposition 3.1.

Let S={(x,y,p3(x,y),,pn(x,y))|x,yU}S=\{(x,y,p_{3}(x,y),\cdots,p_{n}(x,y))~{}|~{}x,y\in U\}, where UU is an open set of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})), and each pi(x,y)p_{i}(x,y) is a degree 22 polynomial and AA be an affine rational hyperplane in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}. Suppose that SS is not contained inside AA then there are at most finitely many points of SAS\cap A such that its neighbourhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}.

By the above proposition we have that SAJ=𝐳SAγ(𝐳)S\cap A\setminus J=\cup_{\mathbf{z}\in S\cap A}\gamma(\mathbf{z}), where JJ is a finite set of points which do not have an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve as neighborhood, and γ(𝐳)\gamma(\mathbf{z}) is an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve which is a neighborhood of 𝐳\mathbf{z} in SAS\cap A. Also note that γ(𝐳)\gamma(\mathbf{z}) is open and closed. Since SAJS\cap A\setminus J is a second countable space, we know that there exists a countable subcovering γj\gamma_{j}, i.e. SAJ=iγiS\cap A\setminus J=\cup_{i}\gamma_{i}.

Theorem 3.4.

Let SS be as in the previous proposition. There exist {Li}\{L_{i}\}, {Lj}\{L_{j}^{\prime}\}, {Aj}\{A_{j}\} as mentioned in Theorem 3.2, that satisfy property A.

Proof.

Let {Ai}\{A_{i}\} be the set of affine rational hyperplanes normal to one of xx-axis or yy-axis in 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n}. By possibly replacing SS with a smaller restriction on xx and yy, we can ensure that for any ζS\zeta\in S, TζST_{\zeta}S is normal to xx and yy-axis. Now let us define Li=SAiL_{i}=S\cap A_{i}, which are closed subsets and curves of SS.

Next we define {Lj}\{L_{j}^{\prime}\}. For any affine rational hyperplane AA that has a nonempty intersection with SS, by proposition 3.1, we have that SAS\cap A is union of γj\gamma_{j}, excluding finitely many points. Let {Lj}={γj:i,γjLi}\{L_{j}^{\prime}\}=\{\gamma_{j}:\forall i,\gamma_{j}\not\subset L_{i}\}. 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 Li=SAiL_{i}=S\cap A_{i}, where Ai:x=aA_{i}:x=a, a𝔽q(T)a\in\mathbb{F}_{q}(T). Now let us consider Ajk:y=bTkA_{j_{k}}:y=\frac{b}{T^{k}}, where b𝔽q[T]b\in\mathbb{F}_{q}[T]. Since Li={(a,y,p3(a,y),,pn(a,y)|yU}L_{i}=\{(a,y,p_{3}(a,y),\cdots,p_{n}(a,y)|y\in U\}, and {bTk|km}\{\frac{b}{T^{k}}|k\geq m\} for some mm\in\mathbb{N}, is dense in UU, property (b) follows.
Let FF and FF^{\prime} are as in hypothesis (c) of Theorem 3.2. Let

Li={(a,y,p3(a,y),pn(a,y))|x,yU},L_{i}=\{(a,y,p_{3}(a,y)\cdots,p_{n}(a,y))~{}|~{}x,y\in U\},

where a𝔽q(T)a\in\mathbb{F}_{q}(T). Note that each LiLkL_{i}\cap L_{k} with kFk\in F is either empty or consists of only a single point. For any kFk^{\prime}\in F^{\prime}, by the equation given above, LiLk=LiγkL_{i}\cap L_{k^{\prime}}^{\prime}=L_{i}\cap\gamma_{k^{\prime}}. We can write γk\gamma_{k^{\prime}} is a subset of a1x+a2y+i=3naipi(x,y)an+1=0a_{1}x+a_{2}y+\sum_{i=3}^{n}a_{i}p_{i}(x,y)-a_{n+1}=0, and therefore, LiγkL_{i}\cap\gamma_{k^{\prime}} is a subset of a1a+a2y+i=3naipi(a,y)an+1=0a_{1}a+a_{2}y+\sum_{i=3}^{n}a_{i}p_{i}(a,y)-a_{n+1}=0. Hence only finitely many solution is possible and LiLkL_{i}\cap L^{\prime}_{k^{\prime}} has no interior. The same proof will work when Li=AiSL_{i}=A_{i}\cap S where AiA_{i} is normal to yy-axis.
To verify property (d) of Theorem 3.2 it is enough to observe that SAiS\cap A_{i} looks like

(x,a,p3(x,y),,pn(x,y)|x,yU}(x,a,p_{3}(x,y),\cdots,p_{n}(x,y)~{}|~{}x,y\in U\}

or

(b,y,p3(x,y),,pn(x,y)|x,yU},(b,y,p_{3}(x,y),\cdots,p_{n}(x,y)~{}|~{}x,y\in U\},

where a,b𝔽q(T)a,b\in\mathbb{F}_{q}(T). Clearly they form a dense set in SS. ∎

3.2. A special case in higher degree p(x,y)p(x,y)

Proposition 3.2.

Let char(𝔽q((T1)))=p<+\text{char}(\mathbb{F}_{q}((T^{-1})))=p<+\infty. Let S={(x,y,p(x,y))|x,yU}S=\{(x,y,p(x,y))~{}|~{}x,y\in U\}, where p(x,y)=i=0mbipixpi+j=0ncjpjypjp(x,y)=\sum_{i=0}^{m}b_{i}^{p^{i}}x^{p^{i}}+\sum_{j=0}^{n}c_{j}^{p^{j}}y^{p^{j}} and UU is an open subset in 𝔽q((T1))\mathbb{F}_{q}((T^{-1})). There are at most finitely many points of SAS\cap A such that its neighbourhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve in 𝔽q((T1))3\mathbb{F}_{q}((T^{-1}))^{3}.

Proof.

Without loss of generality let us assume that SAS\cap A is nonempty, and that there exists at least one point of SAS\cap A whose neighbourhood is not an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve in 𝔽q((T1))3\mathbb{F}_{q}((T^{-1}))^{3}. This implies a30a_{3}\neq 0, and we can also assume a3=1a_{3}=1 after normalization. This implies that a1+b0=a2+c0=0a_{1}+b_{0}=a_{2}+c_{0}=0. Suppose also without loss of generality that mnm\geq n, bm0b_{m}\neq 0, and cn0c_{n}\neq 0. The intersection SAS\cap A is given by f(x,y)=0,f(x,y)=0, where

f(x,y)\displaystyle f(x,y) =i=1mbipixpi+j=1ncjpjypja4\displaystyle=\sum_{i=1}^{m}b_{i}^{p^{i}}x^{p^{i}}+\sum_{j=1}^{n}c_{j}^{p^{j}}y^{p^{j}}-a_{4}
=(i=1mbipixpi1+j=1ncjpjypj1)pa4\displaystyle=\left(\sum_{i=1}^{m}b_{i}^{p^{i}}x^{p^{i-1}}+\sum_{j=1}^{n}c_{j}^{p^{j}}y^{p^{j-1}}\right)^{p}-a_{4}

Since SAS\cap A is nonempty, a4a_{4} must have a pp-th root, say it is a4,0p=a4a_{4,0}^{p}=a_{4}. Therefore, we have that SAS\cap A is given by

f0(x,y):=i=1mbipixpi1+j=1ncjpjypj1a4,0.f_{0}(x,y):=\sum_{i=1}^{m}b_{i}^{p^{i}}x^{p^{i-1}}+\sum_{j=1}^{n}c_{j}^{p^{j}}y^{p^{j-1}}-a_{4,0}.

If SAS\cap A has one point which has no neighbourhood that is an 𝔽q((T1))\mathbb{F}_{q}((T^{-1}))-analytic curve, we have b1=c1=0b_{1}=c_{1}=0. Hence we can write f0(x,y)=0f_{0}(x,y)=0 as,

f0(x,y)=(i=2mbipixpi2+j=2ncjpjypj2)pa4,0.f_{0}(x,y)=\left(\sum_{i=2}^{m}b_{i}^{p^{i}}x^{p^{i-2}}+\sum_{j=2}^{n}c_{j}^{p^{j}}y^{p^{j-2}}\right)^{p}-a_{4,0}.

Since SAS\cap A is nonempty a4,0a_{4,0} must have a pp-th root, say it is a4,1p=a4,0a_{4,1}^{p}=a_{4,0}. Since f0(x,y)=(f1(x,y))pf_{0}(x,y)=(f_{1}(x,y))^{p}, where

f1(x,y):=i=2mbipixpi2+j=2ncjpjypj2a4,1,f_{1}(x,y):=\sum_{i=2}^{m}b_{i}^{p^{i}}x^{p^{i-2}}+\sum_{j=2}^{n}c_{j}^{p^{j}}y^{p^{j-2}}-a_{4,1},

we have that SAS\cap A is defined by f1(x,y)=0f_{1}(x,y)=0.

We use an induction to derive the desired results. For any kk\in\mathbb{N} satisfying 1kn11\leq k\leq n-1, b1==bk1=0b_{1}=\dots=b_{k-1}=0, and c1==ck1=0c_{1}=\dots=c_{k-1}=0, assume that SAS\cap A is given by fk(x,y)=0f_{k}(x,y)=0, where

fk(x,y)=i=k+1mbipixpik1+j=k+1ncjpjypjk1a4,k\displaystyle f_{k}(x,y)=\sum_{i=k+1}^{m}b_{i}^{p^{i}}x^{p^{i-k-1}}+\sum_{j=k+1}^{n}c_{j}^{p^{j}}y^{p^{j-k-1}}-a_{4,k}

for some a4,k𝔽q((T1))a_{4,k}\in\mathbb{F}_{q}((T^{-1})). We have fkx=bk+1\frac{\partial f_{k}}{\partial x}=b_{k+1} and fky=ck+1\frac{\partial f_{k}}{\partial y}=c_{k+1}. By System (3.4), we derive that bk+1=ck+1=0b_{k+1}=c_{k+1}=0. Now since

fk(x,y)\displaystyle f_{k}(x,y) =i=k+2mbipixpik1+j=k+2ncjpjypjk1a4,k\displaystyle=\sum_{i=k+2}^{m}b_{i}^{p^{i}}x^{p^{i-k-1}}+\sum_{j=k+2}^{n}c_{j}^{p^{j}}y^{p^{j-k-1}}-a_{4,k}
=(i=k+2mbipixpik2+j=k+2ncjpjypjk2)pa4,k,\displaystyle=\left(\sum_{i=k+2}^{m}b_{i}^{p^{i}}x^{p^{i-k-2}}+\sum_{j=k+2}^{n}c_{j}^{p^{j}}y^{p^{j-k-2}}\right)^{p}-a_{4,k},

and by our assumption that SAS\cap A is nonempty, we know that a4,ka_{4,k} must have a pp-th root, say it is a4,(k+1)𝔽q((T1))a_{4,(k+1)}\in\mathbb{F}_{q}((T^{-1})), i.e., a4,(k+1)p=a4,ka_{4,(k+1)}^{p}=a_{4,k}. Define

fk+1(x,y):=i=k+2mbipixpik2+j=k+2ncjpjypjk2a4,(k+1).f_{k+1}(x,y):=\sum_{i=k+2}^{m}b_{i}^{p^{i}}x^{p^{i-k-2}}+\sum_{j=k+2}^{n}c_{j}^{p^{j}}y^{p^{j-k-2}}-a_{4,(k+1)}.

Then

fk(x,y)=(fk+1(x,y))p.f_{k}(x,y)=(f_{k+1}(x,y))^{p}.

This implies that the intersection SAS\cap A is given by the equation fk+1(x,y)=0f_{k+1}(x,y)=0.

Repeating the steps above we see that the intersection SAS\cap A is given by fn2(x,y)=0f_{n-2}(x,y)=0. Also from the induction above, we see that

fn2(x,y)\displaystyle f_{n-2}(x,y) =cnpnya4,n2+i=nmbipixpin+1.\displaystyle=c^{p^{n}}_{n}y-a_{4,n-2}+\sum_{i=n}^{m}b_{i}^{p^{i}}x^{p^{i-n+1}}.

This tells us that SAS\cap A is completely given by the curve

{(x,g(x),p(x,g(x)))xU},\{(x,g(x),p(x,g(x)))\ \mid\ x\in U\},

where

g(x):=1cnpn(a4,n2i=nmbipixpin+1), which is analytic.g(x):=\frac{1}{c_{n}^{p^{n}}}\left(a_{4,n-2}-\sum_{i=n}^{m}b_{i}^{p^{i}}x^{p^{i-n+1}}\right),\text{ which is analytic.}

The exact same proof as Theorem 3.4 with suitable changes gives the following theorem.

Theorem 3.5.

Let S={(x,y,i=0mbipixpi+j=0ncjpjypj)}S=\{(x,y,\sum_{i=0}^{m}b_{i}^{p^{i}}x^{p^{i}}+\sum_{j=0}^{n}c_{j}^{p^{j}}y^{p^{j}})\}, where x,y,bi,cj𝔽q((T1))x,y,b_{i},c_{j}\in\mathbb{F}_{q}((T^{-1})), and not all bib_{i}s or cjc_{j}s are being zero. Then there exists {Li}\{L_{i}\}, {Lj}\{L_{j}^{\prime}\}, {Aj}\{A_{j}\} that satisfies property A.

We now have everything we need to prove the main theorem of this section:

Proof of Theorem 3.1.

Theorem 3.4 and Theorem 3.5 guarantee that the hypotheses of Theorem 3.2 are met by surfaces considered in Theorem 3.1. Therefore by Theorem 3.2, Theorem 3.1 follows. ∎

3.3. Surface to higher dimensional manifold

The following theorem is somewhat an analogue to Lemma 3.53.5 in [14].

Theorem 3.6.

Suppose that for k3k\geq 3 and B(0,1)B(0,1) is the open and closed ball in 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) of radius 11. Let MM be a kk-dimensional submanifold of 𝔽q((T1))n\mathbb{F}_{q}((T^{-1}))^{n} which is the image of B(0,1)kB(0,1)^{k} under an immersion 𝐟:B(0,1)k𝔽q((T1))n\mathbf{f}\ :B(0,1)^{k}\to\mathbb{F}_{q}((T^{-1}))^{n}. Suppose that (𝐟,λk)(\mathbf{f},\lambda_{k}) is nonplanar, where λk\lambda_{k} is the Lebesgue measure in B(0,1)kB(0,1)^{k}. Then there exists 𝐲B(0,1)d2\mathbf{y}\in B(0,1)^{d-2} such that the surface M𝐲:=𝐟𝐲((B(0,1)2)M_{\mathbf{y}}:=\mathbf{f}_{\mathbf{y}}((B(0,1)^{2}) is not contained inside an rational affine hyperplane, where 𝐟𝐲:B(0,1)2𝔽q((T1))n\mathbf{f}_{\mathbf{y}}\ :B(0,1)^{2}\to\mathbb{F}_{q}((T^{-1}))^{n}, 𝐟𝐲(x1,x2)=𝐟(x1,x2,𝐲)\mathbf{f}_{\mathbf{y}}(x_{1},x_{2})=\mathbf{f}(x_{1},x_{2},\mathbf{y}).

Proof.

We will prove by contradiction. Suppose for every 𝐲B(0,1)k2\mathbf{y}\in B(0,1)^{k-2} there exists a rational affine subspace A𝐲A_{\mathbf{y}} such that B(0,1)2×{𝐲}𝐟1(A𝐲M)B(0,1)k=A is a rational affine subspace𝐟1(AM).B(0,1)^{2}\times\{\mathbf{y}\}\subset\mathbf{f}^{-1}(A_{\mathbf{y}}\cap M)\implies B(0,1)^{k}=\cup_{A\text{ is a rational affine subspace}}\mathbf{f}^{-1}(A\cap M). By Baire category theorem there is one AA such that 𝐟1(AM)\mathbf{f}^{-1}(A\cap M) contains an open ball inside B(0,1)kB(0,1)^{k}. This contradicts the fact that (𝐟,λk)(\mathbf{f},\lambda_{k}) is nonplanar. ∎

Remark 2.
  1. (1)

    In the above theorem we don’t need to consider the manifold to be analytic but we need a stronger assumption that (𝐟,λ)(\mathbf{f},\lambda) is nonplanar as compared to the manifold being not inside an affine hyperplane.

  2. (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 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) is called perfect if it is compact and has no isolated points. If L={𝐲𝔽q((T1))n|i=1naiyi=a0}L=\{\mathbf{y}\in\mathbb{F}_{q}((T^{-1}))^{n}~{}|~{}\sum_{i=1}^{n}a_{i}y_{i}=a_{0}\}, we can define |L|:=maxi=1n|ai||L|:=\max_{i=1}^{n}|a_{i}|. The proof of the next proposition will be exactly the same as the proof of Theorem 1.6 in [14].

Proposition 3.3.

Let n2n\geq 2 and let S1,,SnS_{1},\dots,S_{n} be perfect subsets of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) such that 𝔽q(T)S1\mathbb{F}_{q}(T)\cap S_{1} is dense in S1S_{1} and 𝔽q(T)S2\mathbb{F}_{q}(T)\cap S_{2} is dense in S2S_{2}. Let S=j=1nSjS=\prod_{j=1}^{n}S_{j}. Then there exist a collection of {Li}\{L_{i}\}, {Lj}\{L_{j}^{\prime}\}, {Ai}\{A_{i}\} of SS that satisfy property A.

Thus we have the following theorem combining Theorem 3.2 and Proposition 3.3.

Theorem 3.7.

Let n2n\geq 2 and let S1,,SnS_{1},\dots,S_{n} be perfect subsets of 𝔽q((T1))\mathbb{F}_{q}((T^{-1})) such that (𝔽q(T)S1)(\mathbb{F}_{q}(T)\cap S_{1}) is dense in S1S_{1} and (𝔽q(T)S2)(\mathbb{F}_{q}(T)\cap S_{2}) is dense in S2S_{2}. Let S=j=1nSjS=\prod_{j=1}^{n}S_{j}. Then there exist uncountably many totally irrational singular vectors in SS.

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.