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

Real Mahler Series

Andean E. Medjedovic
Abstract

Let α\alpha and β\beta be positive real numbers. Let F(x)K[[xΓ]]F(x)\in K[[x^{\Gamma}]] be a Hahn series. We prove that if F(x)F(x) is both α\alpha-Mahler and β\beta-Mahler then it must be a rational function, F(x)K(x)F(x)\in K(x), assuming some non-degeneracy conditions on α\alpha and β\beta.

1 Introduction

During the 1920’s Mahler was able to show that the function

F(x)=n=0x2n{}F(x)=\sum_{n=0}^{\infty}x^{2^{n}} (1)

takes on transcendental values for every algebraic number 0<x<10<x<1 [11]. Since then his work has been generalized and the Mahler Method is now a staple of transcendence theory [13]. Further work by Cobham has highlighted connections to automata theory [7] [3]. Cobham was able to prove the following.

Theorem 1.1 (Cobham).

Suppose kk and ll are multiplicatively independent integers. Then if a sequence of natural numbers is both kk-automatic and ll-automatic, it is eventually periodic.

The proof is infamous for its difficulty despite using only elementary techniques. The correspondence between kk-automatic sequences and what are known as Mahler functions has been widely known for at least 15 years [8]. Cobham’s theorem naturally leads to the analogous question for Mahler functions, which we will now introduce. The astute reader will immediately verify that the series in equation 1 satisfies the below functional equation:

F(x2)=F(x)x.F(x^{2})=F(x)-x. (2)

We say that particular F(x)F(x) is 22-Mahler of degree 11. Generalizing this notion, suppose a series satisfies

Pd(x)F(xkd)++P1(x)F(xk)+P0(x)F(x)=A(x).P_{d}(x)F(x^{k^{d}})+\ldots+P_{1}(x)F\left(x^{k}\right)+P_{0}(x)F\left(x\right)=A(x). (3)

We will then say that FF is kk-Mahler of degree dd. Here PiP_{i} and AA are polynomials. An equivalent formulation would be to say that

{1,F(x),F(xk),F(xk2),}\{1,F(x),F(x^{k}),F(x^{k^{2}}),\ldots\} (4)

forms a dd-dimensional vector space over K[x]K[x].

Let α>0\alpha>0 and β>0\beta>0 be real numbers. The reader will recall that if log(α)log(β)\frac{\log(\alpha)}{\log(\beta)}\not\in\mathbb{Q} then α\alpha and β\beta are said to be multiplicatively independent. Loxton and Van der Poorten conjectured the following counterpart of Cobham’s theorem for Mahler functions

Conjecture 1.2 (Loxton and Van der Poorten).

Let kk and ll be multiplicatively independent positive integers. Let FK[[x]]F\in K[[x]] be a power series over a number field KK that is both kk- and ll-Mahler. Then FF is rational.

Notice that the converse of the theorem holds trivially, rational functions are α\alpha-Mahler for any α\alpha. In 2013 this conjecture was resolved by Bell and Adamczewski who were the first to prove the more general theorem, over any field of characteristic 0 [1] .

Theorem 1.3 (Bell and Adamczewski).

Let kk and ll be multiplicatively independent positive integers. Let FK[[x]]F\in K[[x]] be a power series over a field of characteristic 0, KK, that is both kk- and ll-Mahler. Then FF is rational.

The theorem was also proven by Schafke and Singer through entirely different methods in 2017 [15]. In this paper we will extend the result beyond positive integers. The main results are as follows.

Theorem 1.4.

Let α>0\alpha>0 and β>0\beta>0 be two multiplicatively independent elements of \mathbb{Q}. Suppose F(x)F(x) is a Hahn series over \mathbb{R} that is both α\alpha- and β\beta-Mahler. Then F(x)F(x) is rational.

Theorem 1.5.

Let α>0\alpha>0 and β>0\beta>0 be two algebraically independent elements of \mathbb{R}. Suppose F(x)F(x) is a Hahn series over \mathbb{R} that is both α\alpha- and β\beta-Mahler. Then F(x)F(x) is rational.

2 A proof sketch

Here we outline the structure of the argument giving a summary of each section.

  1. 1.

    In the next section we will define and summarize some basic facts about Hahn series. We then go on to prove a few quick lemmas that allow us to reduce the form of the Mahler functional equations to something more manageable, at least in the rational case. Later on in the proof we will use some more reductions of a similar nature that will be proven as we need them.

  2. 2.

    Afterward, we will restrict our vision for a moment, and look at only Laurent series instead of Hahn series. The goal of this section is to prove a nonexistence result of α\alpha-Mahler Laurent series when α\alpha is not an integer. This result, while somewhat interesting in its own right, will allow us to make further claims when dealing with the more general case of Hahn series.

  3. 3.

    We will analyze the supports of Hahn series under the Mahler functional equations to gain a broad understanding of how they interact. This will then prove invaluable in making further reductions to a more tractable problem.

  4. 4.

    Finally, we will tackle the problem and prove theorem 1.4 under the assumption that α\alpha and β\beta are rational. In the case that α\alpha and β\beta are integers we appeal to the Adamczewski-Bell theorem. In the other case we aim to reduce the Hahn series by putting together all the results found in the previous sections.

  5. 5.

    Some additional details are needed when one of α\alpha or β\beta is irrational. We delineate these cases and prove theorem 1.5 for them. This completes the paper.

3 Preliminaries

In this section we introduce some definitions and notation. We will also take the time to prove some elementary facts and reductions that will be helpful to us later on. Let KK be a number field with K[x],K(x),K[[x]],K[x],K(x),K[[x]], and K[[x]][x1]K[[x]][x^{-1}] denoting polynomials, rational functions, power series and Laurent series over KK. We will use (p,q)(p,q) as a shorthand for the greatest common denominator of integers pp and qq.

3.1 Hahn series

Let Γ\Gamma be an ordered group. Consider the set of formal expressions, F,F, so that

F(x)=iΓfixiF(x)=\sum_{i\in\Gamma}f_{i}x^{i} (5)

with fiKf_{i}\in K and the additional constraint that the support of FF is well-ordered. We will use P(F)P(F) to designate this set of non-zero powers appearing in FF,

P(F):={iΓ:fi0}.P(F):=\{i\in\Gamma:f_{i}\neq 0\}. (6)

Such a formal expression is known as a Hahn series or sometimes as a Hahn-Mal’cev-Neumann series. Of course, one can add Hahn series together in the natural way. We require P(F)P(F) to be well-ordered to ensure that multiplication of Hahn series is well defined. If F=iΓfixiF=\sum_{i\in\Gamma}f_{i}x^{i} and G=iΓgixiG=\sum_{i\in\Gamma}g_{i}x^{i} are two Hahn series then

F(x)G(x)=iΓ(j1+j2=ifj1gj2)xi.F(x)G(x)=\sum_{i\in\Gamma}\left(\sum_{j_{1}+j_{2}=i}f_{j_{1}}g_{j_{2}}\right)x^{i}. (7)

Since P(F)P(F) and P(G)P(G) are well-ordered the inner sum has finitely many non-zero terms and is therefore well-defined. The field of all Hahn series over KK with group Γ\Gamma will be indicated by K[[xΓ]]K[[x^{\Gamma}]]. This field has the following valuation which we will sometimes make use of

v(F):=miniP(F)i.v(F):=\min_{i\in P(F)}i. (8)

Unless otherwise stated, we default to Γ=\Gamma=\mathbb{R}.

3.2 Reductions in the Rational case

Suppose for a moment that α=p1q\alpha=\frac{p_{1}}{q} and β=p2q\beta=\frac{p_{2}}{q} are rationals, and we do not necessarily have (p1,q)=(p2,q)=1\left(p_{1},q\right)=\left(p_{2},q\right)=1. Let FF be both α\alpha- and β\beta-Mahler as a Hahn series. We show in this subsection that it is possible to obtain from F(x)F(x), some other Hahn series with desirable properties.

Pd(x)F(xαd)++P1(x)F(xα)+P0(x)F(x)=A(x)\displaystyle{}P_{d}(x)F(x^{\alpha^{d}})+\ldots+P_{1}(x)F\left(x^{\alpha}\right)+P_{0}(x)F\left(x\right)=A(x) (9)
Qd(x)F(xβd)++Q1(x)F(xβ)+Q0(x)F(x)=B(x)\displaystyle Q_{d}(x)F(x^{\beta^{d}})+\ldots+Q_{1}(x)F\left(x^{\beta}\right)+Q_{0}(x)F\left(x\right)=B(x) (10)
Lemma 3.1.

We can re-write the functional equations 9 as α\alpha- and β\beta-Mahler with A(x)A(x) and B(x)B(x) = 0.

Proof.

Apply the operator xxαqx\mapsto x^{\alpha q} to the first line of equation 9. Multiply both sides by A(x)A(x). Take the original equation and multiply it by A(xαq)A(x^{\alpha q}). These operations take polynomials to polynomials and if we let F(xq)=G(x)F(x^{q})=G(x) then we are left with 22 Mahler equations for GG with equal right hand sides, namely, A(xαq)A(x)A(x^{\alpha q})A(x). Take the difference between the two equations and we are left with a homogeneous α\alpha-Mahler equation for GG. Repeat the argument for the second equation to force B(x)=0B(x)=0. ∎

Lemma 3.2.

If FF satisfies equation 9 then there is a Mahler function, GG where we can assume P0(x)0P_{0}(x)\neq 0.

Proof.

Suppose P0(x)=P1(x)==Pi1(x)=0P_{0}(x)=P_{1}(x)=\ldots=P_{i-1}(x)=0 and Pi(x)0P_{i}(x)\neq 0. Define G(x)G(x) to be the Hahn series with F(x)=G(x1αi)F(x)=G(x^{\frac{1}{\alpha^{i}}}). Then G(x)G(x) satisfies an α\alpha-Mahler equation with non-zero initial term. ∎

Lemma 3.3.

If FF satisfies equation 9 we can assume α>1\alpha>1 and β>1\beta>1.

Proof.

Let α=pq<1\alpha=\frac{p}{q}<1 written in lowest common form. Let G(x)=F(xpd)G(x)=F(x^{p^{d}}). Apply xx1αdx\mapsto x^{\frac{1}{\alpha^{d}}} to the first equation in 9. This give a 1α\frac{1}{\alpha}-Mahler equation for FF over the coefficient ring K[x1αd]K[x^{\frac{1}{\alpha^{d}}}]. Taking this equation under the operator xxpdx\mapsto x^{p^{d}} takes FF to GG while taking K[x1αd]K[x^{\frac{1}{\alpha^{d}}}] to K[x]K[x]. Since 1α>1\frac{1}{\alpha}>1, this will suffice. The proof for β\beta is the exact same applied to the second. We can guarantee the two series GG will be equal by raising to the greatest common denominator of the numerators of α\alpha and β\beta, G(x)G(xa)G(x)\mapsto G(x^{a}), in each case. ∎

4 Non-integer Mahler power series

Let α=pq\alpha=\frac{p}{q} be a non-integer positive rational with (p,q)=1\left(p,q\right)=1. Require that α1n\alpha\neq\frac{1}{n} for an integers nn. Then we will show here that FF is a α\alpha-Mahler Laurent series if and only if FF is rational. Notice that the backward direction is trivial. The restriction on α\alpha is quite necessary. If nn, a natural number, is multiplicatively dependent to 1n\frac{1}{n} we can construct 1nk\frac{1}{n^{k}}-Mahler functions from nn-Mahler functions. We give one such example and leave the rest to the reader.

Let F(x)=n=2x2nF(x)=\sum_{n=2}^{\infty}x^{2^{n}}. Then F(x)F(x) is 22-Mahler and satisfies

F(x2)=F(x)x4.F(x^{2})=F(x)-x^{4}. (11)

Normalize so that we get a homogeneous functional equation

F(x4)(1+x4)F(x2)+F(x)=0F(x^{4})-(1+x^{4})F(x^{2})+F(x)=0 (12)

And applying the xx12x\mapsto x^{\frac{1}{2}} operator twice:

F(x)(1+x)F(x12)+xF(x14)=0F(x)-(1+x)F(x^{\frac{1}{2}})+xF(x^{\frac{1}{4}})=0 (13)

yields a 12\frac{1}{2}-Mahler functional equation.

Moving on to a proof, we say a Laurent series, FF, in K[[x]][x1]K[[x]][x^{-1}] is α\alpha-Mahler of degree dd if there exist polynomials P0(x),,Pd(x)P_{0}(x),\ldots,P_{d}(x) in K[x]K[x] so that

Pd(x)F(xαd)+Pd1(x)F(xαd1)++P0(x)F(x)=0.{}P_{d}(x)F(x^{{\alpha}^{d}})+P_{d-1}(x)F(x^{\alpha^{d-1}})+\ldots+P_{0}(x)F(x)=0. (14)

under these conditions we must have

Theorem 4.1.

F(x)F(x) is in K(x)K(x). That is, FF is rational.

The approach is to first reduce to the degree 11 case. From there we argue using an infinite product formulation of the problem.

Lemma 4.2.

We must have F(x)=G(xqd)F(x)=G(x^{q^{d}}) for some Laurent series G(x)G(x).

Proof.

Consider coefficients of terms with exponents of form a+k(pq)da+k\left(\frac{p}{q}\right)^{d} in 14 for a,ka,k\in\mathbb{N} with qkq\nmid k. They have to be 0 from the right hand side and on the left hand side they only come from the Pd(x)F(xαd)P_{d}(x)F(x^{\alpha^{d}}) term. Thus F(x)=F0(xq)F(x)=F_{0}(x^{q}) for some Laurent series F0F_{0}. Now substitute F0(xq)F_{0}(x^{q}) for F(x)F(x) in 14 and repeat the argument mutatis mutandis. ∎

Corollary 4.3.

F(xαi)F(x^{\alpha^{i}}) are Laurent series as well for i=0,1,di=0,1\ldots,d.

We use +γ\mathbb{Z}+\gamma to denote the set {z+γ}\{z+\gamma\} for zz\in\mathbb{Z}.

Lemma 4.4.

Let F(x)F(x) be a Laurent series and let S(x),T(x)S(x),T(x) be Hahn series. Let

S(x)=T(x)F(x).S(x)=T(x)F(x).

If [S(x)]+γ{}_{\mathbb{Z}+\gamma}\left[S(x)\right] denotes the sum of terms composing S(x)S(x) with exponents in +γ\mathbb{Z}+\gamma then

[S(x)]+γ=+γ[T(x)]F(x).{}_{\mathbb{Z}+\gamma}\left[S(x)\right]=_{\mathbb{Z}+\gamma}\left[T(x)\right]F(x).
Proof.

Since F(x)F(x) is a Laurent series, the exponents of terms in F(x)F(x) are in \mathbb{Z}. So each term of form xγx^{\gamma} maps to a sum of terms of form xz+γx^{z+\gamma} after multiplication by F(x)F(x). ∎

We will later use the more abstract version of this notation. If GG is a Hahn series we will take [G]A{}_{A}[G] to mean the Hahn series containing only terms with powers contained in AA. An immediate consequence of this definition is P([G]A)=AP(G)P({}_{A}[G])=A\cap P(G).

We are ready to prove the central result of this section and let α>1\alpha>1. Write F(x)=xv(F)1G(x)F(x)=x^{v(F)-1}G(x), then G(xl)G(x^{l}) is also α\alpha-Mahler for some ll that divides v(Qi)v(Q_{i}) for polynomials in the functional equation defining If we know

Qd(x)G(xαd)+Qd1(x)G(xαd1)++Q0(x)G(x)=0Q_{d}(x)G(x^{{\alpha}^{d}})+Q_{d-1}(x)G(x^{\alpha^{d-1}})+\ldots+Q_{0}(x)G(x)=0 (15)

then we know the powers of xx across the left hand side must appear in 22 of the dd terms and cancel to 0. In particular, the minimal terms in each of the dd summands must satisfy this constraint so we know

αnl+b0=αml+b1\alpha^{n}l+b_{0}=\alpha^{m}l+b_{1}

for some integers b0b_{0} and b1b_{1} and naturals nn and mm. If α\alpha is a rational that is not an integer then

αnαm\alpha^{n}-\alpha^{m}\not\in\mathbb{Z}

while b1b0l\frac{b_{1}-b_{0}}{l} is. Contradiction. ∎

5 Reduction to linear subsets

Let SS be a well ordered subset of \mathbb{R} under <<. Consider the formal Hahn series FKS[[x]]F\in K^{S}[[x]]. Suppose FF is a homogeneous α\alpha-Mahler function as well, with α\alpha\in\mathbb{R} for some α>1\alpha>1\in\mathbb{R}. Again, by Lemma 3.3, α>1\alpha>1 is assumed to be the case.

Let Pd(x)0P_{d}(x)\neq 0 and assume FF satisfies

i=0dPi(x)F(xαi)=0\sum_{i=0}^{d}P_{i}(x)F(x^{{\alpha}^{i}})=0 (16)

where F(x)=sSfsxsF(x)=\sum_{s\in S}f_{s}x^{s}. We can assume fsf_{s} is never 0 by taking a subset of SS if need be. SS must be countable, and use the well-ordering principle to index S={s0,s1,}S=\{s_{0},s_{1},\ldots\}.

The point of this section is to establish this theorem.

Theorem 5.1.

Let FF be an α\alpha-Mahler Hahn series. Then, using the notation instituted earlier, we can rewrite FF as a sum of other α\alpha-Mahler functions corresponding to each equivalence class

F(x)=sP(F)^[F(x)]T(s)F(x)=\sum_{s\in\widehat{P(F)}}{}_{T(s)}[F(x)] (17)

where T(s):=αs+[α][α1]T(s):=\alpha^{\mathbb{Z}}s+\mathbb{Z}[\alpha][\alpha^{-1}].

We use this theorem later on in the proof of our main theorem to reduce to the case where the set of powers of xx is a subset of [α][α1]\mathbb{Z}[\alpha][\alpha^{-1}].

Proof.

For now assume FF satisfies equation and write FF as

F(x)=i=1dRi(x)F(xαi){}F(x)=\sum_{i=1}^{d}R_{i}(x)F(x^{\alpha^{i}}) (18)

for some rational functions Ri(x)R_{i}(x). Notice the exponents of terms of the right hand side all of form

αis+c\alpha^{i}s+c (19)

where cc\in\mathbb{Z} and sP(F)s\in P(F). Every exponent occurring on the left must also occur on the right. In this case the left is P(F)P(F) while the exponents occurring on the right will be a union of the sets αiP(F)+c\alpha^{i}P(F)+c as ii and cc vary. Then

P(F)n=1αnP(F)+[α][α1]{}P(F)\subset\bigcup_{n=1}^{\infty}\alpha^{n}P(F)+\mathbb{Z}[\alpha][\alpha^{-1}] (20)

where A+BA+B is the set a+ba+b over all aAa\in A and bBb\in B. Let xx and yy be elements of \mathbb{R} and we will define the equivalence relation \sim to mean there is some p[α][α1]p\in\mathbb{Z}[\alpha][\alpha^{-1}] and integer mm for which

αmx+p=y.\alpha^{m}x+p=y. (21)

The reader can check that this is indeed an equivalence relation. The upshot of this equivalence is that we can pass it through equation 20. Let P(F)^=P(F)/\widehat{P(F)}=P(F)/\sim so that we now have

P(F)n=1αnP(F)^+[α][α1].P(F)\subset\bigcup_{n=1}\alpha^{n}\widehat{P(F)}+\mathbb{Z}[\alpha][\alpha^{-1}]. (22)

Notice that each equivalence class αs+[α][α1]\alpha^{\mathbb{Z}}s+\mathbb{Z}[\alpha][\alpha^{-1}] is closed under multiplication by α\alpha and addition of integers. Therefore, we can rewrite FF as a sum over a α\alpha-Mahler function corresponding to each equivalence class

F(x)=sP(F)^[F(x)]T(s)F(x)=\sum_{s\in\widehat{P(F)}}{}_{T(s)}[F(x)] (23)

where T(s):=αs+[α][α1]T(s):=\alpha^{\mathbb{Z}}s+\mathbb{Z}[\alpha][\alpha^{-1}]. In which case [F]T(s){}_{T(s)}[F] is α\alpha-Mahler with the same polynomial scalars, as required. Note that each term in the sum is α\alpha-Mahler with the same functional equation.

We shall soon see that we can sharpen the restriction P(F)[α][α1]+αsP(F)\subset\mathbb{Z}[\alpha][\alpha^{-1}]+\alpha^{\mathbb{Z}}s further still, under some assumptions. We will however, wait and do this spread out over a few cases. For now, we point out another restriction, namely

Lemma 5.2.

Let FF be an α\alpha-Mahler Hahn series. P(F)T(s)=P(F)\cap T(s)=\emptyset for all but finitely many ss\in\mathbb{R}. In other words, P(F)^\widehat{P(F)} is a finite set.

Proof.

Suppose FF satisfies

i=0dPi(x)F(xαi)=0.\sum_{i=0}^{d}P_{i}(x)F(x^{\alpha^{i}})=0. (24)

Consider the valuation of each term in the sum, v(Pi(x)F(xαi))=ci+αif0v(P_{i}(x)F(x^{\alpha^{i}}))=c_{i}+\alpha^{i}f_{0} where v(Pi)=civ(P_{i})=c_{i} and v(F)=f0v(F)=f_{0}. For the right hand side to be 0, terms on the left hand side must all cancel out. In particular, we must have solutions to

cj+αjf0=ci+αif0c_{j}+\alpha^{j}f_{0}=c_{i}+\alpha^{i}f_{0} (25)

for some ij{0,,d}i\neq j\in\{0,\ldots,d\}. Here i,ji,j are allowed to vary while the ci,cjc_{i},c_{j} and α\alpha are fixed by the functional equation. Of course, each equation is linear so there are only finitely many f0f_{0} that are possible as ii and jj vary. From this we conclude T(s)P(F)=T(s)\cap P(F)=\emptyset for all but finitely many ss\in\mathbb{R}. ∎

We end this section we a small observation of what occurs in the case where one of the [F]T(s){}_{T(s)}[F] is rational, which will be useful to us.

Lemma 5.3.

Suppose F(x)F(x) is an α\alpha-Mahler Hahn series with α\alpha rational. There is some integer ll for which P(F(xl))[α,α1]P(F(x^{l}))\subset\mathbb{Z}[\alpha,\alpha^{-1}]. Trivially, F(xl)F(x^{l}) is an α\alpha-Mahler Hahn series as well.

Proof.

Let ss\in\mathbb{R} be arbitrary. There are only finitely many ss for which P(F)T(s)P(F)\cap T(s)\neq\emptyset by Lemma 5.2. Note that \sim is an equivalence on all of \mathbb{R}, so any element of P(F)P(F) is contained in a member of it. That is, the union of all T(s)T(s) cover P(F)P(F). By Lemma 6.2 we may further assume ss\in\mathbb{Q}. We have finitely many rationals ss for which the intersection with the support is non-trivial, let ll be a multiple of the least common multiple of all such ss. Then l×T(s)[α,α1]l\times T(s)\subset\mathbb{Z}[\alpha,\alpha^{-1}]. From this it follows P(F(xl))[α,α1]P(F(x^{l}))\subset\mathbb{Z}[\alpha,\alpha^{-1}]. ∎

6 Doubly rational Mahler Hahn series

As before let F(x)K[[x]]F(x)\in K^{\mathbb{R}}[[x]] be a Hahn series over \mathbb{R}. Let α=p1q\alpha=\frac{p_{1}}{q} and β=p2q\beta=\frac{p_{2}}{q} be multiplicatively independent rationals (although not necessarily written in lowest common form). In this section we aim to prove theorem 1.4

Theorem.

If F(x)F(x) is both α\alpha and β\beta-Mahler then F(x)F(x) is rational.

We begin by strengthening the previous Lemma. Our goal is to show we can restrict to where P(F)[α,α1]P(F)\subset\mathbb{Z}[\alpha,\alpha^{-1}], and similarly for β\beta. Decompose FF across equivalence classes as in theorem 5.1. By focusing on each term of the decomposition individually we can restrict to where P(F)T(s)P(F)\subset T(s) for some properly chosen ss by Lemma 5.1. That is, suppose P(F)αs+[α,α1]P(F)\subset\alpha^{\mathbb{Z}}s+\mathbb{Z}[\alpha,\alpha^{-1}] for some s≁0s\not\sim 0. We show that ss must be rational. In the case that ss is rational, we can take F(x)F(x) to a large enough power by the operator xxlx\mapsto x^{l}.

Lemma 6.1.

Let FF be α\alpha-Mahler. Then the functions FF and [F]T(s){}_{T(s)}[F] satisfy the same α\alpha-Mahler function.

Proof.

The proof is trivial, begin with

n=0dPd(x)F(xαi)=0\sum_{n=0}^{d}P_{d}(x)F(x^{\alpha^{i}})=0 (26)

and apply the []T(s){}_{T(s)}[\cdot] to the equation. Since +T(s)=T(s)\mathbb{Z}+T(s)=T(s) it follows that the operator is linear over K[x]K[x]. ∎

Lemma 6.2.

Suppose ss is irrational. Then [F]T(s)=0{}_{T(s)}[F]=0.

Proof.

Suppose ss is irrational. We will show that this forces F=0F=0. Notice that if ss is irrational then the only solutions to

αms+p1=αns+p2\alpha^{m}s+p_{1}=\alpha^{n}s+p_{2} (27)

where m,nm,n are integers and p1,p2[α,α1]p_{1},p_{2}\in\mathbb{Z}[\alpha,\alpha^{-1}] is when m=nm=n and p1=p2p_{1}=p_{2}. We aim to contradict the well-ordered property of P(F)P(F). We know FF must satisfy

P0(x)F(x)=n=1dPd(x)F(xαi).P_{0}(x)F(x)=\sum_{n=1}^{d}P_{d}(x)F(x^{\alpha^{i}}). (28)

We can write every element of fT(s)f\in T(s) uniquely as αns+p\alpha^{n}s+p with nn\in\mathbb{Z} and p[α,α1]p\in\mathbb{Z}[\alpha,\alpha^{-1}]. We call the pair (n,deg(p))(n,\deg(p)) the degree of ff. The degree of p[α,α1]p\in\mathbb{Z}[\alpha,\alpha^{-1}] is the largest power of α\alpha. Suppose fP(F)f\in P(F). So a term of form xfx^{f} appears in the left hand side of the above equation. Of course, it must also come from a term on the right hand side so

f=c+αif0f=c+\alpha^{i}f_{0} (29)

where cc is a term coming from rational scaling terms PiP_{i} and f0f_{0} is another element of P(F)T(s)P(F)\subset T(s). Now notice the degree of f0f_{0} is either (ni,deg(p)i)(n-i,\deg(p)-i) or (ni,i)(n-i,-i) depending on whether the constant cc comes into play. Apply the same argument to f0f_{0} inductively to get a sequence of fjf_{j} and notice the degree lexicographically decreases every iteration. The sequence of polynomials which give the second argument of the degree either converges to 0 or becomes arbitrarily close to cαi\frac{c}{\alpha^{i}} over all possible c,ic,i, of which there are finitely many. In the latter case choose a subsequence of the fjf_{j} so that the polynomials converge to a fixed c0αi0\frac{c_{0}}{\alpha^{i_{0}}}, in which case fjf_{j} must also converge to the same value.

Recall that P(F)P(F) must be well-ordered by the definition of Hahn series. By the convergence of the fjf_{j}, P(F)P(F) is dense somewhere. The reader will notice that this is a contradiction. This completes the proof. We move on to the case where ss is rational.

Lemma 6.3.

Suppose [F]αs+[α,α1]0{}_{\alpha^{\mathbb{Z}}s+\mathbb{Z}[\alpha,\alpha^{-1}]}[F]\neq 0 and FF is α\alpha-Mahler. Suppose furthermore [F]βs+[β,β1]0{}_{\beta^{\mathbb{Z}}s^{\prime}+\mathbb{Z}[\beta,\beta^{-1}]}[F]\neq 0 and FF is β\beta-Mahler. Then there is an integer nn for which [F(xn)][α,α1]0{}_{\mathbb{Z}[\alpha,\alpha^{-1}]}[F(x^{n})]\neq 0 and [F(xn)][β,β1]0{}_{\mathbb{Z}[\beta,\beta^{-1}]}[F(x^{n})]\neq 0.

Proof.

We have seen that if ss is irrational, then [F]T(s)=0{}_{T(s)}[F]=0. Suppose s=kls=\frac{k}{l} is rational. Apply xxlx\mapsto x^{l} to FF to get a function, GG, with P(G)[α,α1]P(G)\subset\mathbb{Z}[\alpha,\alpha^{-1}]. Do a similar operation to FF for β\beta. Suppose we decompose FF with [F]T(s)0{}_{T(s^{\prime})}[F]\neq 0, and T(s)=βs+[β,β1]T(s^{\prime})=\beta^{\mathbb{Z}}s^{\prime}+\mathbb{Z}[\beta,\beta^{-1}] where s=kls^{\prime}=\frac{k^{\prime}}{l^{\prime}}. Taking xxllx\mapsto x^{l\cdot l^{\prime}} gives P(F)[α,α1]P(F)\subset\mathbb{Z}[\alpha,\alpha^{-1}] and P(F)[β,β1]P(F)\subset\mathbb{Z}[\beta,\beta^{-1}]. Then we can proceed to the below proof.

The case where both α\alpha and β\beta are integers is the Adamczewski-Bell theorem [1]. Therefore we can assume at least one of α\alpha or β\beta is not an integer.

We quickly deal with the case that β\beta is an integer. Assume that β\beta is an integer. By the theorem we just proved, 5.1 ,we have that P(F)P(F)\subset\mathbb{Z} in which case α\alpha being a non-integer rational and F(x)F(x) being α\alpha-Mahler contradicts the theorem established two sections ago, theorem 4.1 unless F(x)F(x) is rational.

We begin the main thrust with a Lemma.

Lemma 6.4.

Fix an N>0N>0. There is an integer ll so that, if F(x)F(x) is α\alpha and β\beta-Mahler then F(xl)F(x^{l}) is αnβm\alpha^{n}\beta^{m}-Mahler for all integers m,nm,n with |m|<N|m|<N and |n|<N|n|<N.

Proof.

The idea is the same as in proposition 8.18.1 in Adamczewski-Bell . We must be more careful to take the powers appearing in the polynomials into account. Suppose F(x)F(x) is α\alpha-Mahler of degree d1d_{1} and β\beta-Mahler of degree d2d_{2}. Consider the functional equations defining F(x)F(x):

i=0d1Pi(x)F(xαi)=0{}\sum_{i=0}^{d_{1}}P_{i}(x)F(x^{\alpha^{i}})=0 (30)
i=0d1Qi(x)F(xβi)=0.{}\sum_{i=0}^{d_{1}}Q_{i}(x)F(x^{\beta^{i}})=0. (31)

Applying the xxαix\mapsto x^{\alpha^{i}} to equation gives a linear dependence for F(xαd1+i)F(x^{\alpha^{d_{1}+i}}) in terms of lower order terms over K[xαi]K[x^{\alpha^{i}}]. Using the formula for the lower order terms gives F(xαd1+1)F(x^{\alpha^{d_{1}+1}}) in the span

F(xαd1+1)SpanK[xαi]{F(x),,F(xαd11)}.{}F(x^{\alpha^{d_{1}+1}})\in\operatorname{Span}_{K[x^{\alpha^{i}}]}\{F(x),\ldots,F(x^{\alpha^{d_{1}-1}})\}. (32)

Of course, the same fact holds for β\beta after making the necessary changes. In case, we are interested in F(xαm)F(x^{\alpha^{m}}) for m<0m<0 apply xxαmx\mapsto x^{\alpha^{m}}. This gives a linear dependence of for F(xαm)F(x^{\alpha^{m}}) but in terms of higher degree terms over the coefficient field K[xαm]K[x^{\alpha^{m}}]. Inductively, we conclude the negative version of equation 32:

F(xαm)SpanK[xαm]{F(x),,F(xαd11)}.F(x^{\alpha^{m}})\in\operatorname{Span}_{K[x^{\alpha^{m}}]}\{F(x),\ldots,F(x^{\alpha^{d_{1}-1}})\}. (33)

The same holds for β\beta. Let

Vl=SpanK[x1l]{F(xαiβj)}0id110jd21.V_{l}=\operatorname{Span}_{K[x^{\frac{1}{l}}]}\{F(x^{\alpha^{i}\beta^{j}})\}_{\begin{subarray}{c}0\leq i\leq d_{1}-1\\ 0\leq j\leq d_{2}-1\end{subarray}}.

Notice that if ll is sufficiently large then by the above logic we know F(xαn0βm0)F(x^{\alpha^{n_{0}}\beta^{m_{0}}}) is in VlV_{l} for all |n0|<d1d2N|n_{0}|<d_{1}d_{2}N and |m0|<d1d2N|m_{0}|<d_{1}d_{2}N. Since the dimension of VlV_{l} is at most d1d2d_{1}d_{2} we know and F(xαjnβjm)F(x^{\alpha^{jn}\beta^{jm}}) is in VlV_{l} for all j=0,d1d2j=0,\ldots d_{1}d_{2}.

From this we conclude that F(x)F(x) is αnβm\alpha^{n}\beta^{m}-Mahler of degree at most d1d2d_{1}d_{2}, over the coefficient ring K[x1l]K[x^{\frac{1}{l}}]. Finally, applying xxlx\mapsto x^{l} gives the desired conclusion. ∎

So we consider instead G(x)=F(xl)G(x)=F(x^{l}). We haven’t yet specified NN, and ll depends on it. We will leave it unspecified for now. Let pp be a prime and vpv_{p}, the pp-adic valuation. We need one more quick lemma before we embark on the final stretch of the rational case.

Lemma 6.5.

There is some natural number NN for which

vp(αnβm)=0v_{p}(\alpha^{n}\beta^{m})=0 (34)

has a solution in nn and mm, for any prime pp, with |n||n| and |m||m| bounded by NN.

Proof.

The proof is trivial; expand vp(αnβm)=nvp(α)+mvp(β)v_{p}(\alpha^{n}\beta^{m})=nv_{p}(\alpha)+mv_{p}(\beta) and notice we can take n=kvp(β),m=kvp(α)n=kv_{p}(\beta),m=-kv_{p}(\alpha) for an integer kk larger than, say d1d2d_{1}d_{2}. Since only finitely many primes appear in the expansions of α\alpha or β\beta, a finite NN will indeed exist. ∎

Moving on to the proof of theorem 1.4.

Lemma 6.6.

Let NN be a natural number. Let F(x)F(x) an arbitrary α\alpha-Mahler Hahn series with α\alpha rational. We can choose an integer ll for which the support of G(x)=F(xl)G(x)=F(x^{l}) is

P(G)|n|,|m|N[αnβm,(αnβm)1].P(G)\subset\bigcap_{|n|,|m|\leq N}\mathbb{Z}[\alpha^{n}\beta^{m},\left(\alpha^{n}\beta^{m}\right)^{-1}]. (35)
Proof.

By Lemma 6.4 there is an integer l0l_{0} for which F(xl0)F(x^{l_{0}}) is αnβm\alpha^{n}\beta^{m}-Mahler for all integers m,nm,n with |m|<N|m|<N and |n|<N|n|<N. By Lemma 5.3, for each of the MM different αnβm\alpha^{n}\beta^{m}-Mahler functional equations we obtained (nn and mm vary) there is another integer lil_{i} (ii goes from 11 to MM) for which P(F(xl0li))[αnβm,αnβm]P(F(x^{l_{0}l_{i}}))\subset\mathbb{Z}[\alpha^{n}\beta^{m},\alpha^{-n}\beta^{-m}]. Let

l=i=0Mli.l=\prod_{i=0}^{M}l_{i}.

Then if G(x)=F(xl)G(x)=F(x^{l}) we have

P(G)|n|,|m|N[αnβm,(αnβm)1]P(G)\subset\bigcap_{|n|,|m|\leq N}\mathbb{Z}[\alpha^{n}\beta^{m},\left(\alpha^{n}\beta^{m}\right)^{-1}]

as required.

Using the above Lemma we can complete the proof in a few lines. Take any prime, pp. By construction, there is an element of the intersection, [αnβm]\mathbb{Z}[\alpha^{n^{\prime}}\beta^{m^{\prime}}] which does not contain 1p\frac{1}{p}. From this we conclude

P(G).P(G)\subset\mathbb{Z}. (36)

But theorem 4.1 forbids this. Contradiction.

7 Doubly irrational Mahler Hahn series

Some care needs to be devoted to the case where one of α\alpha or β\beta is irrational. For this section we assume α\alpha and β\beta are algebraically independent. We demonstrate the analogous theorem to theorem 1.4 in this case, theorem 1.5.

Theorem.

If F(x)F(x) is a Hahn series that is both α\alpha- and β\beta-Mahler then F(x)K(x)F(x)\in K(x).

Proof.

We know FF satisfies both

i=0d1Pi(x)F(xαi)=0\sum_{i=0}^{d_{1}}P_{i}(x)F(x^{\alpha^{i}})=0 (37)

and

i=0d1Qi(x)F(xβi)=0.\sum_{i=0}^{d_{1}}Q_{i}(x)F(x^{\beta^{i}})=0. (38)

Consider the minimal element of P(F)P(F), namely, v(F)v(F). Over all possibilities, there is some ld1l\leq d_{1} and integer aa coming from exponents of xx in PiP_{i} so that the quantity αlv(F)+a\alpha^{l}v(F)+a is minimized. For the left hand side of the above to be 0 the value αlm+a\alpha^{l}m+a must occur in two different terms. That is to say, there is some l1,l2l_{1},l_{2} so that

(αl1αl2)v(F).(\alpha^{l_{1}}-\alpha^{l_{2}})v(F)\in\mathbb{Z}.

Mutatis Mutandis for β\beta in which case

(βk1βk2)v(F).\left(\beta^{k_{1}}-\beta^{k_{2}}\right)v(F)\in\mathbb{Z}.

But then, of course,

(αl1αl2)(βk1βk2)\frac{\left(\alpha^{l_{1}}-\alpha^{l_{2}}\right)}{\left(\beta^{k_{1}}-\beta^{k_{2}}\right)}\in\mathbb{Q}

so α\alpha and β\beta cannot be algebraically independent.

8 Concluding Remarks

To summarize, we have shown that there exist no Laurent series that are α\alpha-Mahler for α\alpha multiplicatively independent from a natural number, other than rational functions FK(x)F\in K(x). We then use this result to prove there exist no doubly Mahler Hahn series, α\alpha- and β\beta-Mahler Hahn series, where α\alpha and β\beta are multiplicatively independent rationals and algebraically independent numbers. A question open for further investigation is whether this result for irrational numbers can be relaxed to α\alpha and β\beta multiplicatively independent instead of algebraically independent:

Conjecture 8.1.

Let α\alpha and β\beta be irrational and multiplicatively independent. If F(x)F(x) is a Hahn series over \mathbb{R} that is both α\alpha- and β\beta-Mahler, then F(x)K(x)F(x)\in K(x).

Acknowledgements. — The author would like the thank Jason Bell for suggesting the problem and discussions surrounding the problem. Without him this paper would not be possible.

References

  • [1] Boris Adamczewski and Jason P. Bell “A problem around Mahler functions”, 2013 arXiv:1303.2019 [math.NT]
  • [2] Boris Adamczewski, Thomas Dreyfus, Charlotte Hardouin and Michael Wibmer “Algebraic independence and linear difference equations”, 2020 arXiv:2010.09266 [math.NT]
  • [3] Jean-Paul Allouche and Jeffrey Shallit “Automatic sequences: theory, applications, generalizations” Cambridge university press, 2003
  • [4] Jason Bell, Frederic Chyzak, Michael Coons and Philippe Dumas “Becker’s conjecture on Mahler functions”, 2018 arXiv:1802.08653 [math.NT]
  • [5] Richard P. Brent, Michael Coons and Wadim Zudilin “Algebraic Independence of Mahler Functions via Radial Asymptotics” In International Mathematics Research Notices Oxford University Press (OUP), 2015, pp. rnv139 DOI: 10.1093/imrn/rnv139
  • [6] Jakub Byszewski and Jakub Konieczny “A density version of Cobham’s theorem”, 2017 arXiv:1710.07261 [math.NT]
  • [7] Alan Cobham “On the base-dependence of sets of numbers recognizable by finite automata” In Mathematical systems theory 3.2 Springer, 1969, pp. 186–192
  • [8] Alan Cobham “On the Hartmanis-Stearns problem for a class of tag machines” In 9th Annual Symposium on Switching and Automata Theory (swat 1968), 1968, pp. 51–60 IEEE
  • [9] Thomas Dreyfus, Charlotte Hardouin and Julien Roques “Hypertranscendence of solutions of Mahler equations” In Journal of the European Mathematical Society 20.9 European Mathematical Society Publishing House, 2018, pp. 2209–2238 DOI: 10.4171/jems/810
  • [10] JH Loxton and AJ Van Der Poorten “Arithmetic properties of automata: regular sequences” In J. reine angew. Math 392.57-69, 1988, pp. 79
  • [11] Kurt Mahler “Arithmetische eigenschaften einer klasse transzendental-transzendenter funktionen” In Mathematische Zeitschrift 32.1 Springer, 1930, pp. 545–585
  • [12] Kumiko Nishioka “Mahler Functions and Transcendence” Springer
  • [13] Federico Pellarin “An introduction to Mahler’s method for transcendence and algebraic independence”, 2011 arXiv:1005.1216 [math.NT]
  • [14] AJ Poorten and JH Loxton “Arithmetic properties of the solutions of a class of functional equations.” In Journal für die reine und angewandte Mathematik 1982.330 De Gruyter, 1982, pp. 159–172
  • [15] Reinhard Schäfke and Michael F. Singer “Mahler equations and rationality”, 2017 arXiv:1605.08830 [math.CA]
  • [16] Ehud Shalit “Elliptic (p,q)-difference modules”, 2020 arXiv:2007.09508 [math.NT]