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

Transcendental series of reciprocals of Fibonacci and Lucas numbers

Khoa D. Nguyen Khoa D. Nguyen
Department of Mathematics and Statistics
University of Calgary
AB T2N 1N4, Canada
[email protected]
(Date: September 2020)
Abstract.

Let F1=1,F2=1,F_{1}=1,F_{2}=1,\ldots be the Fibonacci sequence. Motivated by the identity k=01F2k=752\displaystyle\sum_{k=0}^{\infty}\frac{1}{F_{2^{k}}}=\frac{7-\sqrt{5}}{2}, Erdös and Graham asked whether k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} is irrational for any sequence of positive integers n1,n2,n_{1},n_{2},\ldots with nk+1nkc>1\frac{n_{k+1}}{n_{k}}\geq c>1. We resolve the transcendence counterpart of their question: as a special case of our main theorem, we have that k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} is transcendental when nk+1nkc>2\frac{n_{k+1}}{n_{k}}\geq c>2. The bound c>2c>2 is best possible thanks to the identity at the beginning. This paper provides a new way to apply the Subspace Theorem to obtain transcendence results and extends previous non-trivial results obtainable by only Mahler’s method for special sequences of the form nk=dk+rn_{k}=d^{k}+r.

Key words and phrases:
Algebraic numbers, transcendental numbers, Subspace Theorem
2010 Mathematics Subject Classification:
Primary: 11J87. Secondary: 11B39.

1. introduction

Let F1=1,F2=1,F_{1}=1,F_{2}=1,\ldots be the Fibonacci sequence and let L1=1L_{1}=1, Ln=Fn1+FnL_{n}=F_{n-1}+F_{n} for n2n\geq 2 be the Lucas sequence. In the chapter “Irrationality and Transcendence” of their book [EG80, p. 64–65], starting from the Millin series:

k=01F2k=752,\sum_{k=0}^{\infty}\frac{1}{F_{2^{k}}}=\frac{7-\sqrt{5}}{2},

Erdös and Graham asked the following:

Question 1.1 (Erdös-Graham, 1980).

Is it true that k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} is irrational for any sequence n1<n2<n_{1}<n_{2}<\ldots with nk+1nkc>1\frac{n_{k+1}}{n_{k}}\geq c>1?

The transcendence counterparts of this and many questions in [EG80, Chapter 7] were implicit throughout the chapter, hence its title. Indeed, this topic inspired intense research activities most of which involved the so called Mahler’s method. As a consequence of our main result (see Theorem 1.3), we resolve the transcendence version of Question 1.1 even when one is allowed to randomly mix the Fibonacci and Lucas numbers:

Theorem 1.2.

Let c>2c>2 and let n1<n2<n_{1}<n_{2}<\ldots be positive integers such that nk+1nkc\displaystyle\frac{n_{k+1}}{n_{k}}\geq c for every kk. Then the number k=11fk\displaystyle\sum_{k=1}^{\infty}\frac{1}{f_{k}} is transcendental where fk{Fnk,Lnk}f_{k}\in\{F_{n_{k}},L_{n_{k}}\} for every kk.

Although Fibonacci and Lucas numbers have been discovered for hundreds of years, some of their basic properties have been established only recently thanks to powerful modern methods, for example [BMS06, Ste13]. The key ingredient of the proof of our main theorems is a new application of the Subspace Theorem in treating transcendence of series in which it is hard to control the denominators of the partial sums. Before providing more details about the method, let us provide a very brief and incomplete survey of known results on irrationality and transcendence of sums like k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}}. In a nutshell, we can divide previous results in two groups.

The first group treats series in which the denominators of the partial sums are very small compared to the reciprocals of the error terms. This includes work of Mignotte on the transcendence of k=11k!F2k\displaystyle\sum_{k=1}^{\infty}\frac{1}{k!F_{2^{k}}} and k=12+(1)kF2k\displaystyle\sum_{k=1}^{\infty}\frac{2+(-1)^{k}}{F_{2^{k}}} [Mig71, Mig77]. In fact, these results predate Erdös-Graham question. As another example, consider s:=k=01F2k+1s:=\displaystyle\sum_{k=0}^{\infty}\frac{1}{F_{2^{k}+1}}. For every sufficiently large integer NN, thanks to divisibility properties of the Fibonacci sequence and the fact that 2k+12^{k}+1 divides 23k+12^{3k}+1, the denominator of k=1N1F2k+1\displaystyle\sum_{k=1}^{N}\frac{1}{F_{2^{k}+1}} is at most k=N/3NF2k+1\displaystyle\prod_{k=\lfloor N/3\rfloor}^{N}F_{2^{k}+1} which is o(F2N+1+1)o(F_{2^{N+1}+1}), hence ss must be irrational. We refer the readers to [Bad93] and the references there for similar results. In fact, if n1<n2<n_{1}<n_{2}<\ldots satisfies nk+1nkc>2\displaystyle\frac{n_{k+1}}{n_{k}}\geq c>2 then k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} is irrational since Fn1FnN=o(FnN+1)F_{n_{1}}\cdots F_{n_{N}}=o(F_{n_{N+1}}) as NN\to\infty. Likewise, if nk+1nkc>3\displaystyle\frac{n_{k+1}}{n_{k}}\geq c>3 then k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} is transcendental by applying Roth’s theorem and the fact that Fn1FnN=O(FnN+1(1/2)ϵ)F_{n_{1}}\cdots F_{n_{N}}=O\left(F_{n_{N+1}}^{(1/2)-\epsilon}\right) for an appropriate ϵ>0\epsilon>0. However, replacing those easy bounds 22 and 33 respectively by smaller numbers for irrationality and transcendence problems appears to be a very difficult task.

The second group constitutes the majority of results in this topic. In 1975, Mahler [Mah75] reproved Mignotte’s result using the method he had invented nearly 50 years earlier (see Nishioka’s notes [Nis96] for an introduction to Mahler’s method). This method is applicable when the sequence nkn_{k} has the special form nk=dk+rn_{k}=d^{k}+r where d,rd,r\in{\mathbb{Z}} with d2d\geq 2. We refer the readers to [BT94, DKT02, KKS09] and the references there for further details. In fact, before this paper, there has not been one result establishing the transcendence of k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} for an arbitrary sequence nkn_{k} with nk+1nkc\displaystyle\frac{n_{k+1}}{n_{k}}\geq c for any single value c<3c<3 (as explained above, c>3c>3 is the easy bound due to an immediate application of Roth’s theorem).

From now on, let α±1\alpha\neq\pm 1 be a real quadratic unit; this means (α){\mathbb{Q}}(\alpha) is real quadratic and α±1\alpha\neq\pm 1 a unit in the ring of algebraic integers. Let σ\sigma be the non-trivial automorphism of (α){\mathbb{Q}}(\alpha) and let β=σ(α)\beta=\sigma(\alpha). Without loss of generality, we assume |β|<1<|α||\beta|<1<|\alpha|. Let HH and hh respectively denote the absolute multiplicative and logarithmic Weil height on ¯\bar{{\mathbb{Q}}}, our main result is the following:

Theorem 1.3.

Let an,bn,cna_{n},b_{n},c_{n} for n1n\geq 1 be sequences of real numbers with the following properties:

  • For every n1n\geq 1, cnc_{n}\in{\mathbb{Q}}, an,bn(α)a_{n},b_{n}\in{\mathbb{Q}}(\alpha), and un:=anαnbnβnu_{n}:=a_{n}\alpha^{n}-b_{n}\beta^{n}\in{\mathbb{Q}}.

  • limnh(an)n=limnh(bn)n=limnh(cn)n=0\displaystyle\lim_{n\to\infty}\frac{h(a_{n})}{n}=\lim_{n\to\infty}\frac{h(b_{n})}{n}=\lim_{n\to\infty}\frac{h(c_{n})}{n}=0.

Let c>2c>2 and let n1<n2<n_{1}<n_{2}<\ldots be positive integers such that nk+1nkc\displaystyle\frac{n_{k+1}}{n_{k}}\geq c, unk0u_{n_{k}}\neq 0, and cnk0c_{n_{k}}\neq 0 for every kk. Then the series k=1cnkunk\displaystyle\sum_{k=1}^{\infty}\frac{c_{n_{k}}}{u_{n_{k}}} is transcendental.

Example 1.4.

Let n1<n2<n_{1}<n_{2}<\ldots be as in Theorem 1.2. Let s=k=11fks=\displaystyle\sum_{k=1}^{\infty}\frac{1}{f_{k}} where fk{Fnk,Lnk}f_{k}\in\{F_{n_{k}},L_{n_{k}}\} for every kk. Note that Fn=αnβn5F_{n}=\displaystyle\frac{\alpha^{n}-\beta^{n}}{\sqrt{5}} and Ln=αn+βnL_{n}=\alpha^{n}+\beta^{n} with α=1+52\alpha=\displaystyle\frac{1+\sqrt{5}}{2} and β=152\beta=\displaystyle\frac{1-\sqrt{5}}{2}. We define cn=1c_{n}=1 for every nn. If n{nk:k1}n\notin\{n_{k}:\ k\geq 1\}, we define an=bn=0a_{n}=b_{n}=0. If n=nkn=n_{k} and fk=Fnkf_{k}=F_{n_{k}}, define an=bn=15a_{n}=b_{n}=\displaystyle\frac{1}{\sqrt{5}}. Finally if n=nkn=n_{k} and fk=Lnkf_{k}=L_{n_{k}}, define an=1a_{n}=1 and bn=1b_{n}=-1. This explains why Theorem 1.2 is a special case of Theorem 1.3.

Example 1.5.

One can consider linear recurrence sequences of rational numbers of the form un=A(n)αn+B(n)βnu_{n}=A(n)\alpha^{n}+B(n)\beta^{n} where A(t),B(t)(α)[t]A(t),B(t)\in{\mathbb{Q}}(\alpha)[t]. Then Theorem 1.3 implies that 1unk\displaystyle\sum\frac{1}{u_{n_{k}}} is transcendental since we may choose cn=1c_{n}=1, an=A(n)a_{n}=A(n), and bn=B(n)b_{n}=-B(n). Note that h(an)=O(logn)h(a_{n})=O(\log n) and h(bn)=O(logn)h(b_{n})=O(\log n) in this case.

There have been two general transcendence results using the Subspace Theorem recently and both involve values of a power series dnzn\sum d_{n}z^{n} at an algebraic number z0z_{0}. One is a result of Adamczewski-Bugeaud [AB07] extending an earlier work of Troi-Zannier [TZ99]. In their work, the coefficients of dnd_{n}’s form an automatic sequence and z0z_{0} is the reciprocal of a Pisot number. The authors rely on the repeating pattern of automatic sequences to apply the Subspace Theorem using linear forms in three variables. The other is a result of Corvaja-Zannier [CZ02] treating the case that dnzn\sum d_{n}z^{n} is lacunary with positive real coefficients and z0(0,1)z_{0}\in(0,1). The problem considered here is different from all the above. While it is true that one can express k=11Fnk\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} as the value of a power series at 1/α1/\alpha with α=1+52\alpha=\displaystyle\frac{1+\sqrt{5}}{2}, neither the coefficients are automatic nor the series is lacunary for an arbitrary choice of nkn_{k} with nk+1/nkc>2n_{k+1}/n_{k}\geq c>2.

The more subtle difference and key reason for the difficulty in settling our current problem are as follows. Let us consider the example s=k=11Fnks=\displaystyle\sum_{k=1}^{\infty}\frac{1}{F_{n_{k}}} and α=1+52\alpha=\displaystyle\frac{1+\sqrt{5}}{2} again. Let sN=k=1N1Fnks_{N}=\displaystyle\sum_{k=1}^{N}\frac{1}{F_{n_{k}}} be the sequence of partial sums so that |ssN1|=O(|α|nN)|s-s_{N-1}|=O(|\alpha|^{-n_{N}}). Now the usual idea is to fix a large integer PP, then truncate each

1FnNP+i=sN,i+O(|α|nN)\frac{1}{F_{n_{N-P+i}}}=s_{N,i}+O(|\alpha|^{-n_{N}})

where sN,is_{N,i} is a finite sum of units for 1iP11\leq i\leq P-1. Then we have:

|ssNPsN,1sN,P1|=O(|α|nN)|s-s_{N-P}-s_{N,1}-\ldots-s_{N,P-1}|=O(|\alpha|^{-n_{N}})

and after assuming that ss is algebraic, one might attempt to apply the Subspace Theorem to this equation for ss, sNPs_{N-P} and the individual units in each sN,is_{N,i}. The difference compared to work of Adamczewski-Bugeaud or Corvaja-Zannier is that while terms in their application of the Subspace Theorem are SS-integers for an appropriate choice of a finite set of places SS in an appropriate number field, here we cannot find such an SS so that the term sNPs_{N-P} above is an SS-integer for infinitely many NN. For this reason, in our situation, when applying the Subspace Theorem we may have the contribution H(sNM)DH(s_{N-M})^{D} where DD is the number of terms. With just the constraint c>2c>2, it is entirely possible for the above contribution to offset the error term O(|α|nN)O(|\alpha|^{-n_{N}}) and one fails to apply the Subspace Theorem. Therefore new ideas are needed to overcome this crucial issue. Moreover, after one applies the Subspace Theorem, it remains a highly nontrivial task to arrive at the desired conclusion from the resulting linear relation. This paper promotes the innovation that one should start with a certain “minimal expression” before applying the Subspace Theorem in order to maximize the benefit of the resulting linear relation. We refer the readers to the discussion right after Proposition 3.6 for more details.

Acknowledgements. The author wishes to thank Professors Yann Bugeaud and Maurice Mignotte for useful comments. The author is partially supported by an NSERC Discovery Grant and a CRC II Research Stipend from the Government of Canada.

2. The Subspace Theorem

The Subspace Theorem is one of the milestones of diophantine geometry in the last 50 years. The first version was obtained by Schmidt [Sch70] and further versions were obtained by Schlickewei and Evertse [Sch92, Eve96, ES02]. This section follows the exposition in the book of Bombieri-Gubler [BG06].

Let M=MM0M_{{\mathbb{Q}}}=M_{{\mathbb{Q}}}^{\infty}\cup M_{{\mathbb{Q}}}^{0} where M0M_{{\mathbb{Q}}}^{0} is the set of pp-adic valuations and MM_{{\mathbb{Q}}}^{\infty} is the singleton consisting of the usual archimedean valuation. More generally, for every number field KK, write MK=MKMK0M_{K}=M_{K}^{\infty}\cup M_{K}^{0} where MKM_{K}^{\infty} is the set of archimedean places and MK0M_{K}^{0} is the set of finite places. Throughout this paper, we fix an embedding of ¯\bar{{\mathbb{Q}}} into {\mathbb{C}} and let |||\cdot| denote the usual absolute value on {\mathbb{C}}. Hence for a number field KK, the set MKM_{K}^{\infty} corresponds to the set of real embeddings and pairs of complex-conjugate embeddings of KK into {\mathbb{C}}. For every wMKw\in M_{K}, let KwK_{w} denote the completion of KK with respect to ww and denote d(w/v)=[Kw:v]d(w/v)=[K_{w}:{\mathbb{Q}}_{v}] where vv is the restriction of ww to {\mathbb{Q}}. Following [BG06, Chapter 1], for every wMKw\in M_{K} restricting to vv on {\mathbb{Q}}, we normalize ||w|\cdot|_{w} as follows:

|x|w=|NKw/v(x)|v1/[K:].|x|_{w}=|\operatorname{N}_{K_{w}/{\mathbb{Q}}_{v}}(x)|_{v}^{1/[K:{\mathbb{Q}}]}.

Let mm\in{\mathbb{N}}, for every vector 𝐮=(u0,,um)Km+1{𝟎}{\mathbf{u}}=(u_{0},\ldots,u_{m})\in K^{m+1}\setminus\{\mathbf{0}\} and wMKw\in M_{K}, let |𝐮|w:=max0im|ui|w|{\mathbf{u}}|_{w}:=\displaystyle\max_{0\leq i\leq m}|u_{i}|_{w}. For Pm(¯)P\in{\mathbb{P}}^{m}(\bar{{\mathbb{Q}}}), let KK be a number field such that PP has a representative 𝐮Km+1{𝟎}{\mathbf{u}}\in K^{m+1}\setminus\{\mathbf{0}\} and define:

H(P)=wMK|𝐮|w.H(P)=\prod_{w\in M_{K}}|{\mathbf{u}}|_{w}.

It is an easy fact that this is independent of the choice of 𝐮{\mathbf{u}} and the number field KK. Then we define h(P)=log(H(P))h(P)=\log(H(P)). For α¯\alpha\in\bar{{\mathbb{Q}}}, write H(α)=H([α:1])H(\alpha)=H([\alpha:1]) and h(α)=log(H(α))h(\alpha)=\log(H(\alpha)). Later on, we will use the classical version of Roth’s theorem [BG06, Chapter 6] to give a weak upper bound on nk+1/nkn_{k+1}/n_{k}:

Theorem 2.1 (Roth’s theorem).

Let κ>2\kappa>2. Let ss be a real algebraic number. Then there are only finitely many rational numbers ss^{\prime} such that

|ss|H(s)κ.|s^{\prime}-s|\leq H(s^{\prime})^{-\kappa}.

Let mm\in{\mathbb{N}}, for every vector 𝐱=(x0,,xm)Km+1{𝟎}{\mathbf{x}}=(x_{0},\ldots,x_{m})\in K^{m+1}\setminus\{\mathbf{0}\}, let 𝐱~\tilde{{\mathbf{x}}} denote the corresponding point in m(K){\mathbb{P}}^{m}(K). For every wMKw\in M_{K}, denote |𝐱|w:=max0im|xi|w|{\mathbf{x}}|_{w}:=\displaystyle\max_{0\leq i\leq m}|x_{i}|_{w}. We have:

Theorem 2.2 (Subspace Theorem).

Let nn\in{\mathbb{N}}, let KK be a number field, and let SMKS\subset M_{K} be finite. For every vSv\in S, let Lv0,,LvnL_{v0},\ldots,L_{vn} be linearly independent linear forms in the variables X0,,XnX_{0},\ldots,X_{n} with KK-algebraic coefficients in KvK_{v}. For every ϵ>0\epsilon>0, the solutions 𝐱Kn+1{𝟎}{\mathbf{x}}\in K^{n+1}\setminus\{\mathbf{0}\} of the inequality:

vSj=0n|Lvj(𝐱)|v|𝐱|vH(𝐱~)n1ϵ\displaystyle\prod_{v\in S}\prod_{j=0}^{n}\displaystyle\frac{|L_{vj}({\mathbf{x}})|_{v}}{|{\mathbf{x}}|_{v}}\leq H(\tilde{{\mathbf{x}}})^{-n-1-\epsilon}

are contained in finitely many hyperplanes of Kn+1K^{n+1}.

3. Preliminary results and preparation for the proof of Theorem 1.3

Throughout this section, we assume the notation in the statement of Theorem 1.3 and put Ak=ankA_{k}=a_{n_{k}}, Bk=bnkB_{k}=b_{n_{k}}, Ck=cnkC_{k}=c_{n_{k}}, Uk=unkU_{k}=u_{n_{k}} for every kk to simplify the notation. From now on, assume that s:=k=1CkUks:=\displaystyle\sum_{k=1}^{\infty}\frac{C_{k}}{U_{k}} is algebraic and let KK be the Galois closure of (α,s){\mathbb{Q}}(\alpha,s). For m1m\geq 1, let sm=k=1mCkUks_{m}=\displaystyle\sum_{k=1}^{m}\frac{C_{k}}{U_{k}} be the sequence of partial sums. We will repeatedly use the following observation: if (tn)n(t_{n})_{n} is a sequence in KK^{*} such that h(tn)/n0h(t_{n})/n\rightarrow 0 as nn\to\infty then |tn|v=eo(n)|t_{n}|_{v}=e^{o(n)} as nn\to\infty for every vMKv\in M_{K}; this means for every ϵ>0\epsilon>0, we have eϵn<|tn|v<eϵne^{-\epsilon n}<|t_{n}|_{v}<e^{\epsilon n} for all sufficiently large nn.

We are given that Ck0C_{k}\neq 0 for every kk. We may assume that AkBk0A_{k}B_{k}\neq 0 for every kk, as follows. Let σ\sigma denote the nontrivial automorphism of (α){\mathbb{Q}}(\alpha). Suppose that Ak=0A_{k}=0 then Uk=BkβnkU_{k}=-B_{k}\beta^{n_{k}}\in{\mathbb{Q}}^{*}. Applying σ\sigma gives Uk=Bkβnk=σ(Bk)/αnkU_{k}=-B_{k}\beta^{n_{k}}=-\sigma(B_{k})/\alpha^{n_{k}}, therefore σ(Bk)/Bk=(α/β)nk\sigma(B_{k})/B_{k}=(\alpha/\beta)^{n_{k}}. Since |σ(Bk)/Bk|=eo(nk)|\sigma(B_{k})/B_{k}|=e^{o(n_{k})}, we conclude that Ak=0A_{k}=0 is possible for only finitely many kk. A similar conclusion holds for Bk=0B_{k}=0 too. By ignoring the first finitely many nkn_{k}’s, we may assume AkBk0A_{k}B_{k}\neq 0 for every kk.

Similarly, from k=m1m2CkUk=Cm1Um1+O(Cm1+1/Um1+1)\displaystyle\sum_{k=m_{1}}^{m_{2}}\frac{C_{k}}{U_{k}}=\frac{C_{m_{1}}}{U_{m_{1}}}+O(C_{m_{1}+1}/U_{m_{1}+1}) for any m1<m2m_{1}<m_{2}\leq\infty, by ignoring the first finitely many nkn_{k}’s, we may assume:

(1) The numbers ss and sks_{k}’s for k=1,2,k=1,2,\ldots are pairwise distinct and non-zero.

We start with several easy estimates:

Lemma 3.1.
  • (i)

    For every positive integer mm, we have:

    (c1)(n1++nm)<nm+1.(c-1)(n_{1}+\ldots+n_{m})<n_{m+1}.
  • (ii)

    For any positive integers m<Nm<N, we have:

    cNm1(c1)(n1++nm)<nN.c^{N-m-1}(c-1)(n_{1}+\ldots+n_{m})<n_{N}.
Proof.

Part (i) follows from

n1++nmnm+1<1c+1c2+.\frac{n_{1}+\ldots+n_{m}}{n_{m+1}}<\frac{1}{c}+\frac{1}{c^{2}}+\ldots.

Part (ii) follows from nNcNm1nm+1n_{N}\geq c^{N-m-1}n_{m+1} and part (i). ∎

Lemma 3.2.
  • (i)

    H(un)=|α|n+o(n)H(u_{n})=|\alpha|^{n+o(n)} and H(cn/un)=|α|n+o(n)H(c_{n}/u_{n})=|\alpha|^{n+o(n)} as nn\to\infty.

  • (ii)

    H(sN)|α|n1++nN+o(nN)H(s_{N})\leq|\alpha|^{n_{1}+\ldots+n_{N}+o(n_{N})} as NN\to\infty.

Proof.

Due to the fact that unu_{n}\in{\mathbb{Q}}, α\alpha and β\beta are units, and our assumption on the AnA_{n}’s and BnB_{n}’s, we have |un|=|α|n+o(n)|u_{n}|=|\alpha|^{n+o(n)} while the non-archimedean contribution is eo(n)e^{o(n)}. This proves the first assertion of part (i), the remaining one follows since H(cn)=|α|o(n)H(c_{n})=|\alpha|^{o(n)} as nn\to\infty.

For part (ii), we use the inequality:

H(sN)Ni=1NH(cni/uni).H(s_{N})\leq N\prod_{i=1}^{N}H(c_{n_{i}}/u_{n_{i}}).

There exists δ1>0\delta_{1}>0 such that H(cni/uni)|α|ni+δ1niH(c_{n_{i}}/u_{n_{i}})\leq|\alpha|^{n_{i}+\delta_{1}n_{i}} for every ii by part (i). Given any ϵ>0\epsilon>0, part (i) also gives that H(uni)|α|ni+ϵniH(u_{n_{i}})\leq|\alpha|^{n_{i}+\epsilon n_{i}} for every sufficiently large ii. Choose a large integer MM so that

δ1(n1++nNM)ϵnN\delta_{1}(n_{1}+\ldots+n_{N-M})\leq\epsilon n_{N}

for every N>MN>M; this is possible thanks to Lemma 3.1. Hence for all sufficiently large NN, we have:

H(sN)N|α|n1++nN+ϵnN+ϵ(nNM+1++nN)|α|n1++nN+4ϵnNH(s_{N})\leq N|\alpha|^{n_{1}+\ldots+n_{N}+\epsilon n_{N}+\epsilon(n_{N-M+1}+\ldots+n_{N})}\leq|\alpha|^{n_{1}+\ldots+n_{N}+4\epsilon n_{N}}

and this finishes the proof. ∎

Corollary 3.3.

ss is irrational.

Proof.

Suppose ss is rational. From:

  • |ssN|=O(CN+1/UN+1)=O(|α|nN+1+o(nN+1))|s-s_{N}|=O(C_{N+1}/U_{N+1})=O(|\alpha|^{n_{N+1}+o(n_{N+1})}),

  • H(sN)=|α|n1++nN+o(nN)H(s_{N})=|\alpha|^{n_{1}+\ldots+n_{N}+o(n_{N})},

  • n1++nN<(c1)nN+1n_{1}+\ldots+n_{N}<(c-1)n_{N+1}, and

  • c>2c>2,

we have that s=sNs=s_{N} for all sufficiently large NN, contradiction. ∎

Proposition 3.4.

There are only finitely many kk such that nk+1nk5\displaystyle\frac{n_{k+1}}{n_{k}}\geq 5.

Proof.

Let ϵ>0\epsilon>0 that will be specified later. Suppose there are infinitely many kk such that nk+1/nk5n_{k+1}/n_{k}\geq 5. For each such kk that is sufficiently large, we have

|ssk|=O(Ck+1/Uk+1)=O(α(1ϵ)nk+1)=O(α5(1ϵ)nk)|s-s_{k}|=O(C_{k+1}/U_{k+1})=O(\alpha^{-(1-\epsilon)n_{k+1}})=O(\alpha^{-5(1-\epsilon)n_{k}})

while H(sk)|α|n1++nk+ϵnkH(s_{k})\leq|\alpha|^{n_{1}+\ldots+n_{k}+\epsilon n_{k}}. Note that

n1++nk<(1+1c1)nk=cc1nk.n_{1}+\ldots+n_{k}<\left(1+\frac{1}{c-1}\right)n_{k}=\frac{c}{c-1}n_{k}.

We now require ϵ\epsilon to satisfy:

(2) 5(1ϵ)nk>(2cc1+3ϵ)nk;5(1-\epsilon)n_{k}>\left(\frac{2c}{c-1}+3\epsilon\right)n_{k};

this is possible since 5>2cc15>\displaystyle\frac{2c}{c-1}. Then Roth’s theorem implies that the sks_{k}’s take a single value for infinitely many such kk. But this contradicts (1). ∎

Remark 3.5.

In Proposition 3.4, the same arguments can be used when we replace 55 by any constant greater than 2cc1\displaystyle\frac{2c}{c-1}. When c>3c>3, we have c>2cc1c>\displaystyle\frac{2c}{c-1} and this explains the transcendence of cnkunk\displaystyle\sum\frac{c_{n_{k}}}{u_{n_{k}}} given the “easy” bound c>3c>3.

Note that αβ=±1\alpha\beta=\pm 1 since they are units. Then we can use the geometric series to express:

(3) CkUk=CkAkαnk(1(Bkβnk)/(Akαnk))=j=0CkAkαnk(Bk(±1)nkAkα2nk)j=j=0(±1)nkjCkBkjAkj+1α(2j+1)nk\displaystyle\begin{split}\frac{C_{k}}{U_{k}}=\frac{C_{k}}{A_{k}\alpha^{n_{k}}(1-(B_{k}\beta^{n_{k}})/(A_{k}\alpha^{n_{k}}))}&=\sum_{j=0}^{\infty}\frac{C_{k}}{A_{k}\alpha^{n_{k}}}\left(\frac{B_{k}(\pm 1)^{n_{k}}}{A_{k}\alpha^{2n_{k}}}\right)^{j}\\ &=\sum_{j=0}^{\infty}\frac{(\pm 1)^{n_{k}j}C_{k}B_{k}^{j}}{A_{k}^{j+1}\alpha^{(2j+1)n_{k}}}\end{split}

which is valid when kk is sufficiently large so that |Bk/Ak|<|(α/β)nk|=|α|2nk.|B_{k}/A_{k}|<|(\alpha/\beta)^{n_{k}}|=|\alpha|^{2n_{k}}.

Let PP be a large positive integers that will be specified later. In the following, NN denotes an arbitrarily large positive integer. In the various OO-notations and oo-notations, the implied constants might depend on the given data and PP but they are independent of NN. We have:

(4) |ssN1|=|α|nN+o(nN).|s-s_{N-1}|=|\alpha|^{-n_{N}+o(n_{N})}.

As mentioned in the Section 1, it is typical in applications of the Subspace Theorem to break sN1s_{N-1} as sNPs_{N-P} and truncate the expression (3) for k=NP+1,,N1k=N-P+1,\ldots,N-1 to maintain the error term αnN+o(nN)\alpha^{-n_{N}+o(n_{N})}.

For 1iP11\leq i\leq P-1, let

DN,i:=nN2nNP+i5Pi2D_{N,i}:=\left\lfloor\frac{n_{N}}{2n_{N-P+i}}\right\rfloor\leq\frac{5^{P-i}}{2}

thanks to Proposition 3.4. The explicit upper bound here is not important: the key fact is that these DN,iD_{N,i}’s can be bounded from above independently of NN.

Proposition 3.6.

For all sufficiently large NN, we have:

|CNP+iUNP+ij=0DN,i(±1)nNP+ijCNP+iBNP+ijANP+ij+1α(2j+1)nNP+i|<|α|nN+o(nN)\left|\frac{C_{N-P+i}}{U_{N-P+i}}-\sum_{j=0}^{D_{N,i}}\frac{(\pm 1)^{n_{N-P+i}j}C_{N-P+i}B_{N-P+i}^{j}}{A_{N-P+i}^{j+1}\alpha^{(2j+1)n_{N-P+i}}}\right|<|\alpha|^{-n_{N}+o(n_{N})}

for i=1,,P1i=1,\ldots,P-1. This means for every ϵ>0\epsilon>0, the LHS is less than |α|nN+ϵnN|\alpha|^{-n_{N}+\epsilon n_{N}} for all sufficiently large NN.

Proof.

Let ϵ>0\epsilon>0. We have

CNP+iUNP+ij=0DN,i(±1)nNP+ijCNP+iBNP+ijANP+ij+1α(2j+1)nNP+i\displaystyle\frac{C_{N-P+i}}{U_{N-P+i}}-\sum_{j=0}^{D_{N,i}}\frac{(\pm 1)^{n_{N-P+i}j}C_{N-P+i}B_{N-P+i}^{j}}{A_{N-P+i}^{j+1}\alpha^{(2j+1)n_{N-P+i}}}
=j=DN,i+1(±1)nNP+ijCNP+iBNP+ijANP+ij+1α(2j+1)nNP+i.\displaystyle=\sum_{j=D_{N,i}+1}^{\infty}\frac{(\pm 1)^{n_{N-P+i}j}C_{N-P+i}B_{N-P+i}^{j}}{A_{N-P+i}^{j+1}\alpha^{(2j+1)n_{N-P+i}}}.

Hence it suffices to require the first term in the RHS:

(±1)nNP+ijCNP+iBNP+ijANP+ij+1α(2j+1)nNP+iwith j=DN,i+1\frac{(\pm 1)^{n_{N-P+i}j}C_{N-P+i}B_{N-P+i}^{j}}{A_{N-P+i}^{j+1}\alpha^{(2j+1)n_{N-P+i}}}\ \text{with $j=D_{N,i}+1$}

to be O(|α|(1ϵ/2)nN)O(|\alpha|^{-(1-\epsilon/2)n_{N}}). This is actually the case, as follows. First, by the definition of DN,iD_{N,i} we have (2DN,i+3)nNP+inN(2D_{N,i}+3)n_{N-P+i}\geq n_{N}. Second |CNP+iBNP+iDN,i+1ANP+iDN,i+2|<|α|(ϵ/2)nN\displaystyle\left|\frac{C_{N-P+i}B_{N-P+i}^{D_{N,i}+1}}{A_{N-P+i}^{D_{N,i}+2}}\right|<|\alpha|^{(\epsilon/2)n_{N}} when NN is sufficiently large since DN,iD_{N,i} is bounded above independently of NN and the assumption on the sequences (Ak)(A_{k}), (Bk)(B_{k}), and (Ck)(C_{k}). ∎

At this point, one may attempt to apply the Subspace Theorem using the inequality:

(5) |ssNPi=1P1j=0DN,i(±1)nNP+ijCNP+iBNP+ijANP+ij+1α(2j+1)nNP+i|<|α|nN+o(nN)\displaystyle\left|s-s_{N-P}-\sum_{i=1}^{P-1}\sum_{j=0}^{D_{N,i}}\frac{(\pm 1)^{n_{N-P+i}j}C_{N-P+i}B_{N-P+i}^{j}}{A_{N-P+i}^{j+1}\alpha^{(2j+1)n_{N-P+i}}}\right|<|\alpha|^{-n_{N}+o(n_{N})}

and linear forms in 2+i=1P1(DN,i+1)2+\displaystyle\sum_{i=1}^{P-1}(D_{N,i}+1) variables for the terms ss, sNPs_{N-P}, and those in the double sum in a similar manner to [CZ04, p. 180–181] or [KMN19, Proposition 3.4]. However, unlike these previous papers, the term sNPs_{N-P} in our situation is not an SS-integer (for infinitely many NN) for any choice of a finite set SMKS\subset M_{K}. Because of this, there is a potential contribution of H(sNP)2+(DN,i+1)\displaystyle H(s_{N-P})^{2+\sum(D_{N,i}+1)} which could completely offset the error term |α|nN+o(nN)|\alpha|^{-n_{N}+o(n_{N})}.

Our new idea is to consider an extra “buffer zone” by specifying another positive integer Q<PQ<P, expressing

(6) CNP+1UNP+1++CNP+QUNP+Q=xNxNwith xN=i=1QUNP+i,\frac{C_{N-P+1}}{U_{N-P+1}}+\ldots+\frac{C_{N-P+Q}}{U_{N-P+Q}}=\frac{x^{\prime}_{N}}{x_{N}}\text{with $x_{N}=\displaystyle\prod_{i=1}^{Q}U_{N-P+i}$},

and rewriting (5) as

|ssNPxNxNi=1PQ1j=0DN,Q+i(±1)nNP+Q+ijCNP+Q+iBNP+Q+ijANP+Q+ij+1α(2j+1)nNP+Q+i|<|α|nN+o(nN).\displaystyle\begin{split}&\left|s-s_{N-P}-\frac{x^{\prime}_{N}}{x_{N}}-\sum_{i=1}^{P-Q-1}\sum_{j=0}^{D_{N,Q+i}}\frac{(\pm 1)^{n_{N-P+Q+i}j}C_{N-P+Q+i}B_{N-P+Q+i}^{j}}{A_{N-P+Q+i}^{j+1}\alpha^{(2j+1)n_{N-P+Q+i}}}\right|\\ &<|\alpha|^{-n_{N}+o(n_{N})}.\end{split}

We then multiply both sides by xNx_{N} to get

(7) |xNsxNsNPxNi=1PQ1j=0DN,Q+ixN(±1)nNP+Q+ijCNP+Q+iBNP+Q+ijANP+Q+ij+1α(2j+1)nNP+Q+i|<|xN||α|nN+o(nN).\displaystyle\begin{split}&\left|x_{N}s-x_{N}s_{N-P}-x^{\prime}_{N}-\sum_{i=1}^{P-Q-1}\sum_{j=0}^{D_{N,Q+i}}x_{N}\frac{(\pm 1)^{n_{N-P+Q+i}j}C_{N-P+Q+i}B_{N-P+Q+i}^{j}}{A_{N-P+Q+i}^{j+1}\alpha^{(2j+1)n_{N-P+Q+i}}}\right|\\ &<|x_{N}||\alpha|^{-n_{N}+o(n_{N})}.\end{split}

After that we expand xN=i=1QUNP+i=i=1Q(ANP+iαnNP+iBNP+iβnNP+i)x_{N}=\displaystyle\prod_{i=1}^{Q}U_{N-P+i}=\displaystyle\prod_{i=1}^{Q}(A_{N-P+i}\alpha^{n_{N-P+i}}-B_{N-P+i}\beta^{n_{N-P+i}}) as a linear combination of 2Q2^{Q} terms, expand xNx^{\prime}_{N} as a linear combination of Q2Q1Q2^{Q-1} terms. Note that each xNsx_{N}s, xNsNPx_{N}s_{N-P}, as well as each individual term in the double sum now consists of 2Q2^{Q} many terms. In a typical application of the Subspace Theorem, one is worse off after performing the above steps. Therefore it is amusing that in our current situation, those steps can help reduce the number of terms significantly while the new error |xN||α|nN+o(nN)|x_{N}||\alpha|^{-n_{N}+o(n_{N})} is not too much larger than the previous |α|nN+o(nN)|\alpha|^{-n_{N}+o(n_{N})}.

Now even if we can apply the Subspace Theorem, there remains one important technical issue to overcome. After expanding xNx^{\prime}_{N} and xNx_{N}, it might happen that certain terms in the LHS of (7) already satisfied a linear relation and the conclusion of the Subspace Theorem trivially illustrates this fact. For instance, in the double sum in the LHS of (5), if there are two different (i1,j1)(i_{1},j_{1}) and (i2,j2)(i_{2},j_{2}) for which (2j1+1)nNP+i1(2j_{1}+1)n_{N-P+i_{1}} and (2j2+1)nNP+i2(2j_{2}+1)n_{N-P+i_{2}} are close (or even equal) to each other then one should “gather” the two terms corresponding to (i1,j1)(i_{1},j_{1}) and (i2,j2)(i_{2},j_{2}) first. So we will also need a way to efficiently “gather similar terms” so that the conclusion of the Subspace Theorem becomes helpful for our purpose. First, we expand xNx_{N} and xNx^{\prime}_{N}:

Lemma 3.7.
  • (i)

    There exists δ2>0\delta_{2}>0 (possibly depending on PP and QQ) such that for all sufficiently large NN, we can express:

    xN=i=12QxN,iαx(N,i)x_{N}=\sum_{i=1}^{2^{Q}}x_{N,i}\alpha^{x(N,i)}

    with the following properties:

    • (a)

      xN,i(α)x_{N,i}\in{\mathbb{Q}}(\alpha)^{*} and x(N,i)x(N,i)\in{\mathbb{Z}} for every ii.

    • (b)

      x(N,1)=i=1QnNP+ix(N,1)=\displaystyle\sum_{i=1}^{Q}n_{N-P+i} and x(N,1)x(N,j)+2nNP+1x(N,1)\geq x(N,j)+2n_{N-P+1} for every j>1j>1.

    • (c)

      |x(N,i)|x(N,1)|x(N,i)|\leq x(N,1) for every ii.

    • (d)

      h(xN,i)/nN0h(x_{N,i})/n_{N}\to 0 as NN\to\infty for every ii.

    • (e)

      |x(N,i)x(N,j)|δ2nN|x(N,i)-x(N,j)|\geq\delta_{2}n_{N} for any 1ij2Q1\leq i\neq j\leq 2^{Q}.

  • (ii)

    For all sufficiently large NN, we can express

    xN=i=1Q2Q1xN,iαx(N,i)x^{\prime}_{N}=\sum_{i=1}^{Q2^{Q-1}}x^{\prime}_{N,i}\alpha^{x^{\prime}(N,i)}

    with the following properties:

    • (a)

      xN,i(α)x^{\prime}_{N,i}\in{\mathbb{Q}}(\alpha)^{*} and x(N,i)x^{\prime}(N,i)\in{\mathbb{Z}} for every ii.

    • (b)

      |x(N,i)|nNP+2++nNP+Q|x^{\prime}(N,i)|\leq n_{N-P+2}+\ldots+n_{N-P+Q} for every ii.

    • (c)

      h(xN,i)/nN0h(x^{\prime}_{N,i})/n_{N}\to 0 as NN\to\infty.

Proof.

For part (i), let 𝒬={1,,Q}\mathscr{Q}=\{1,\ldots,Q\}. For each T𝒬T\subseteq\mathscr{Q}, put

Σ(N,T)=iTnNP+ii𝒬TnNP+i.\Sigma(N,T)=\sum_{i\in T}n_{N-P+i}-\sum_{i\in\mathscr{Q}\setminus T}n_{N-P+i}.

Note that |Σ(N,T)|i=1QnNP+i<2nNP+Q|\Sigma(N,T)|\leq\displaystyle\sum_{i=1}^{Q}n_{N-P+i}<2n_{N-P+Q} where the last inequality follows from Lemma 3.1. We have:

xN\displaystyle x_{N} =j=1Q(ANP+jαnNP+jBNP+jβnNP+j)\displaystyle=\prod_{j=1}^{Q}(A_{N-P+j}\alpha^{n_{N-P+j}}-B_{N-P+j}\beta^{n_{N-P+j}})
=j=1Q(ANP+jαnNP+jBNP+j(±1)nNP+jαnNP+j).\displaystyle=\prod_{j=1}^{Q}(A_{N-P+j}\alpha^{n_{N-P+j}}-B_{N-P+j}(\pm 1)^{n_{N-P+j}}\alpha^{-n_{N-P+j}}).

We fix once and for all a 1-1 correspondence between {1,,2Q}\{1,\ldots,2^{Q}\} and the set of subsets of 𝒬\mathscr{Q} so that 11 corresponds to 𝒬\mathscr{Q}. This allows us to take the x(N,i)x(N,i)’s to be exactly the Σ(N,T)\Sigma(N,T)’s (with x(N,1)=Σ(N,𝒬)x(N,1)=\Sigma(N,\mathscr{Q})) and the xN,ix_{N,i}’s are the corresponding products of terms among the ANP+jA_{N-P+j} and (1)nNP+jBNP+j(-1)^{n_{N-P+j}}B_{N-P+j}; this proves parts (a) and (d). The largest among the Σ(N,T)\Sigma(N,T)’s is i=1QnNP+i\displaystyle\sum_{i=1}^{Q}n_{N-P+i} while the smallest is i=1QnNP+i-\displaystyle\sum_{i=1}^{Q}n_{N-P+i}; this proves part (c). Moreover, the second largest is

nNP+1+nNP+2++nNP+Q-n_{N-P+1}+n_{N-P+2}+\ldots+n_{N-P+Q}

and this proves part (b). It remains to prove part (e).

Consider two different subsets TT and TT^{\prime} of 𝒬\mathscr{Q}. Let jj^{*} be the largest element in TΔTT\Delta T^{\prime}, then we have:

|Σ(N,T)Σ(N,T)|\displaystyle|\Sigma(N,T)-\Sigma(N,T^{\prime})| 2nNP+jj<j2nNP+j\displaystyle\geq 2n_{N-P+j^{*}}-\sum_{j<j^{*}}2n_{N-P+j}
2(c2)c1nNP+j\displaystyle\geq\frac{2(c-2)}{c-1}n_{N-P+j^{*}}
2(c2)(c1)5PjnN\displaystyle\geq\frac{2(c-2)}{(c-1)5^{P-j^{*}}}n_{N}

where the last two inequalities follow from Lemma 3.1 and Proposition 3.4. We can now take δ2=2(c2)5P1(c1)\delta_{2}=\displaystyle\frac{2(c-2)}{5^{P-1}(c-1)}.

The proof of part (ii) is similar by expanding:

xN\displaystyle x^{\prime}_{N} =i=1QCNP+i1jQ,jiUNP+j\displaystyle=\sum_{i=1}^{Q}C_{N-P+i}\prod_{1\leq j\leq Q,\ j\neq i}U_{N-P+j}
=i=1QCNP+i1jQ,ji(ANP+jαnNP+jBNP+j(±α)nNP+j)\displaystyle=\sum_{i=1}^{Q}C_{N-P+i}\prod_{1\leq j\leq Q,\ j\neq i}(A_{N-P+j}\alpha^{n_{N-P+j}}-B_{N-P+j}(\pm\alpha)^{-n_{N-P+j}})

into Q2Q1Q2^{Q-1} many terms. ∎

Then we expand each individual term in the double sum

i=1PQ1j=0DN,Q+ixN(±1)nNP+Q+ijCNP+Q+iBNP+Q+ijANP+Q+ij+1α(2j+1)nNP+Q+i\displaystyle\sum_{i=1}^{P-Q-1}\sum_{j=0}^{D_{N,Q+i}}x_{N}\frac{(\pm 1)^{n_{N-P+Q+i}j}C_{N-P+Q+i}B_{N-P+Q+i}^{j}}{A_{N-P+Q+i}^{j+1}\alpha^{(2j+1)n_{N-P+Q+i}}}

to get:

Lemma 3.8.

Put η=2Qi=1PQ1(DN,Q+i+1)\eta=2^{Q}\displaystyle\sum_{i=1}^{P-Q-1}(D_{N,Q+i}+1). For all sufficiently large NN, we can express:

i=1PQ1j=0DN,Q+ixN(±1)nNP+Q+ijCNP+Q+iBNP+Q+ijANP+Q+ij+1α(2j+1)nNP+Q+i=i=1ηyN,iαy(N,i)\displaystyle\sum_{i=1}^{P-Q-1}\sum_{j=0}^{D_{N,Q+i}}x_{N}\frac{(\pm 1)^{n_{N-P+Q+i}j}C_{N-P+Q+i}B_{N-P+Q+i}^{j}}{A_{N-P+Q+i}^{j+1}\alpha^{(2j+1)n_{N-P+Q+i}}}=\sum_{i=1}^{\eta}y_{N,i}\alpha^{y(N,i)}

with the following properties:

  • (a)

    yN,i(α)y_{N,i}\in{\mathbb{Q}}(\alpha)^{*} and y(N,i)y(N,i)\in{\mathbb{Z}} for every ii.

  • (b)

    h(yN,i)/nN0h(y_{N,i})/n_{N}\to 0 as NN\to\infty.

  • (c)

    y(N,i)x(N,1)nNP+Q+1=nNP+1++nNP+QnNP+Q+1y(N,i)\leq x(N,1)-n_{N-P+Q+1}=n_{N-P+1}+\ldots+n_{N-P+Q}-n_{N-P+Q+1} for every ii.

  • (d)

    y(N,i)>3nNy(N,i)>-3n_{N} for every ii.

Proof.

We use the expression for xNx_{N} in Lemma 3.7 to expand each individual term in the double sum. This proves (a) and (b). The highest exponent of α\alpha in that expression for xNx_{N} is x(N,1)x(N,1) while the highest exponent of α\alpha among the

(±1)nNP+Q+ijCNP+Q+iBNP+Q+ijANP+Q+ij+1α(2j+1)nNP+Q+i\frac{(\pm 1)^{n_{N-P+Q+i}j}C_{N-P+Q+i}B_{N-P+Q+i}^{j}}{A_{N-P+Q+i}^{j+1}\alpha^{(2j+1)n_{N-P+Q+i}}}

for 1iPQ11\leq i\leq P-Q-1 and 0jDN,Q+i0\leq j\leq D_{N,Q+i} is at most nNP+Q+1-n_{N-P+Q+1}, this proves (c). The smallest exponent of α\alpha among those terms is

max{(2j+1)nNP+Q+i: 1iPQ1, 0jDN,Q+i}>2nN-\max\{(2j+1)n_{N-P+Q+i}:\ 1\leq i\leq P-Q-1,\ 0\leq j\leq D_{N,Q+i}\}>-2n_{N}

by using the definition of the DN,Q+iD_{N,Q+i}’s. The smallest exponent of α\alpha in xNx_{N} is nNP+1nNP+Q>nN-n_{N-P+1}-\ldots-n_{N-P+Q}>-n_{N}. This proves (d). ∎

We also need the following:

Lemma 3.9.

For all sufficiently large NN:

  • (i)

    |xN|=|α|nNP+1++nNP+Q+o(nN)\displaystyle|x_{N}|=|\alpha|^{n_{N-P+1}+\ldots+n_{N-P+Q}+o(n_{N})},

  • (ii)

    |xN(ssN1)|=|xNk=NCkUk|=|α|nNP+1++nNP+QnN+o(nN),\displaystyle|x_{N}(s-s_{N-1})|=\left|x_{N}\sum_{k=N}^{\infty}\frac{C_{k}}{U_{k}}\right|=|\alpha|^{n_{N-P+1}+\ldots+n_{N-P+Q}-n_{N}+o(n_{N})},

  • (iii)

    |xN|<|α|nNcPQ1(c1)+o(nN)\displaystyle|x_{N}|<|\alpha|^{\frac{n_{N}}{c^{P-Q-1}(c-1)}+o(n_{N})}, and

  • (iv)

    |xN(ssN1)|<|α|(11cPQ1(c1))nN+o(nN).|x_{N}(s-s_{N-1})|<|\alpha|^{-(1-\frac{1}{c^{P-Q-1}(c-1)})n_{N}+o(n_{N})}.

Proof.

We have |xN||ANP+1ANP+Q||α|nNP+1++nNP+Q|x_{N}|\gg\ll|A_{N-P+1}\cdots A_{N-P+Q}||\alpha|^{n_{N-P+1}+\ldots+n_{N-P+Q}} and since the AkA_{k}’s are nonzero with height o(nk)o(n_{k}) we have |ANP+1ANP+Q|=|α|o(nN)|A_{N-P+1}\cdots A_{N-P+Q}|=|\alpha|^{o(n_{N})}. This proves part (i). Then Lemma 3.1 gives:

(8) nNP+1++nNP+Q<nNcPQ1(c1).n_{N-P+1}+\ldots+n_{N-P+Q}<\frac{n_{N}}{c^{P-Q-1}(c-1)}.

This proves part (iii).

For part (ii), we use part (i) together with:

|ssN1||CN/UN|=|α|nN+o(nN).|s-s_{N-1}|\gg\ll|C_{N}/U_{N}|=|\alpha|^{-n_{N}+o(n_{N})}.

Finally part (iv) follows from part (ii) and (8). ∎

4. The number ss is in (α){\mathbb{Q}}(\alpha)

We continue with the assumption and notation of Section 3, in particular ss is algebraic. Throughout this section, let Q<PQ<P be large, yet fixed, positive integers that will be specified later and let NN denote an arbitrarily large positive integer. In the various OO-notations and oo-notations, the implied constants might depend on the given data, PP, and QQ but they are independent of NN. In this section, we finish an important step toward the proof of Theorem 1.3 namely proving that s(α)s\in{\mathbb{Q}}(\alpha). This conclusion is similar to the one in the paper of Adamczewski-Bugeaud [AB07, Theorem 5]. In their paper, to obtain transcendence of the given number [AB07, Theorem 5A], they use a result of K. Schmidt [Sch80]. In this paper, we will need more sophisticated applications of the Subspace Theorem together with further combinatorial and Galois theoretic arguments in the next section to obtain the desired result.

As mentioned in the previous section, before applying the Subspace Theorem, we need to come up with an efficient way to “gather similar terms” in the LHS of (7). This is done first by proving the existence of a certain collection of data then choosing a minimal one among those collections.

Proposition 4.1.

Recall the xN,1x_{N,1} and x(N,1)=i=1QnNP+ix(N,1)=\displaystyle\sum_{i=1}^{Q}n_{N-P+i} in the expression for xNx_{N} in Lemma 3.7. There exist integers D,E,F0D,E,F\geq 0, tuples (γ1,,γE)(\gamma_{1},\ldots,\gamma_{E}) of elements of KK, an infinite set 𝒩\mathscr{N}, tuples (dN,1,,dN,D)(d_{N,1},\ldots,d_{N,D}), (d(N,1),,d(N,D))(d(N,1),\ldots,d(N,D)), (eN,1,,eN,E)(e_{N,1},\ldots,e_{N,E}), (e(N,1),,e(N,E))(e(N,1),\ldots,e(N,E)), (fN,1,,fN,F)(f_{N,1},\ldots,f_{N,F}), (f(N,1),,f(N,F))(f(N,1),\ldots,f(N,F)) for each N𝒩N\in\mathscr{N} with the following properties:

  • (i)

    D+E+F(2Q1)+(2Q1)+Q2Q1+2Qi=1PQ1(DN,Q+i+1)D+E+F\leq(2^{Q}-1)+(2^{Q}-1)+Q2^{Q-1}+2^{Q}\displaystyle\sum_{i=1}^{P-Q-1}(D_{N,Q+i}+1).

  • (ii)

    For every N𝒩N\in\mathscr{N}, the d(N,i)d(N,i)’s, e(N,j)e(N,j)’s, and f(N,)f(N,\ell)’s are integers for every i,j,i,j,\ell.

  • (iii)

    For every N𝒩N\in\mathscr{N}, nNP+1+maxi,j,{d(N,i),e(N,j),f(N,)}x(N,1)n_{N-P+1}+\displaystyle\max_{i,j,\ell}\{d(N,i),e(N,j),f(N,\ell)\}\leq x(N,1).

  • (iv)

    For every N𝒩N\in\mathscr{N}, mini,j,k{d(N,i),e(N,j),f(N,)}>3nN\min_{i,j,k}\{d(N,i),e(N,j),f(N,\ell)\}>-3n_{N}.

  • (v)

    For every N𝒩N\in\mathscr{N}, the dN,id_{N,i}’s, eN,je_{N,j}’s, and fN,f_{N,\ell}’s are elements of (α){\mathbb{Q}}(\alpha).

  • (vi)

    As NN\to\infty we have h(dN,i)/nN0h(d_{N,i})/n_{N}\to 0, h(eN,j)/nN0h(e_{N,j})/n_{N}\to 0, and h(fN,)/nN0h(f_{N,\ell})/n_{N}\to 0 for every i,j,i,j,\ell.

  • (vii)

    For all sufficiently large N𝒩N\in\mathscr{N}, we have

    (9) |sxN,1αx(N,1)+j=1EγjeN,jαe(N,j)sNPxN,1αx(N,1)i=1DsNPdN,iαd(N,i)=1FfN,αf(n,)|<|α|(11cPQ1(c1))nN+o(nN).\displaystyle\begin{split}&|sx_{N,1}\alpha^{x(N,1)}+\sum_{j=1}^{E}\gamma_{j}e_{N,j}\alpha^{e(N,j)}-s_{N-P}x_{N,1}\alpha^{x(N,1)}-\sum_{i=1}^{D}s_{N-P}d_{N,i}\alpha^{d(N,i)}\\ &-\sum_{\ell=1}^{F}f_{N,\ell}\alpha^{f(n,\ell)}|<|\alpha|^{-(1-\frac{1}{c^{P-Q-1}(c-1)})n_{N}+o(n_{N})}.\end{split}
Proof.

Recall the inequality (7):

|sxNsNPxNxNi=1PQ1j=0DN,Q+ixN(±1)nNP+Q+ijCNP+Q+iBNP+Q+ijANP+Q+ij+1α(2j+1)nNP+Q+i|\displaystyle\left|sx_{N}-s_{N-P}x_{N}-x^{\prime}_{N}-\sum_{i=1}^{P-Q-1}\sum_{j=0}^{D_{N,Q+i}}x_{N}\frac{(\pm 1)^{n_{N-P+Q+i}j}C_{N-P+Q+i}B_{N-P+Q+i}^{j}}{A_{N-P+Q+i}^{j+1}\alpha^{(2j+1)n_{N-P+Q+i}}}\right|
<|xN||α|nN+o(nN))<|α|(11cPQ1(c1))nN+o(nN)\displaystyle<|x_{N}||\alpha|^{-n_{N}+o(n_{N})})<|\alpha|^{-(1-\frac{1}{c^{P-Q-1}(c-1)})n_{N}+o(n_{N})}

where the last equality follows from Lemma 3.9 and holds when NN is sufficiently large. We now choose 𝒩\mathscr{N} to be the set of all sufficiently large integers, D=E=2Q1D=E=2^{Q}-1, and γ1==γE=s\gamma_{1}=\ldots=\gamma_{E}=s. We want the sum

sxN,1αx(N,1)+j=1EγjeN,jαe(N,j)=s(xN,1αx(N,1)+j=1EeN,jαe(N,j))sx_{N,1}\alpha^{x(N,1)}+\sum_{j=1}^{E}\gamma_{j}e_{N,j}\alpha^{e(N,j)}=s\left(x_{N,1}\alpha^{x(N,1)}+\sum_{j=1}^{E}e_{N,j}\alpha^{e(N,j)}\right)

to be sxNsx_{N}; therefore we simply choose the eN,je_{N,j}’s and e(N,j)e(N,j)’s for 1jE1\leq j\leq E to be respectively the terms xN,kx_{N,k}’s and x(N,k)x(N,k)’s for 2k2Q2\leq k\leq 2^{Q} in the expression for xNx_{N} in Lemma 3.7.

Similarly, we want the sum

sNPxN,1αx(N,1)+i=1DsNPdN,iαd(N,i)=sNP(xN,1αx(N,1)+i=1DdN,iαd(N,i))s_{N-P}x_{N,1}\alpha^{x(N,1)}+\sum_{i=1}^{D}s_{N-P}d_{N,i}\alpha^{d(N,i)}=s_{N-P}\left(x_{N,1}\alpha^{x(N,1)}+\sum_{i=1}^{D}d_{N,i}\alpha^{d(N,i)}\right)

to be sNPxNs_{N-P}x_{N}; therefore we simply choose the dN,id_{N,i}’s and d(N,i)d(N,i)’s for 1iD1\leq i\leq D to be respectively the terms xN,kx_{N,k}’s and x(N,k)x(N,k)’s for 2k2Q2\leq k\leq 2^{Q} above.

Finally, choose F=Q2Q1+ηF=Q2^{Q-1}+\eta (with η\eta in Lemma 3.8) and we want the sum =1FfN,αf(n,)\displaystyle\sum_{\ell=1}^{F}f_{N,\ell}\alpha^{f(n,\ell)} to be

xN+i=1PQ1j=0DN,Q+ixN(±1)nNP+Q+ijCNP+Q+iBNP+Q+ijANP+Q+ij+1α(2j+1)nNP+Q+ix^{\prime}_{N}+\sum_{i=1}^{P-Q-1}\sum_{j=0}^{D_{N,Q+i}}x_{N}\frac{(\pm 1)^{n_{N-P+Q+i}j}C_{N-P+Q+i}B_{N-P+Q+i}^{j}}{A_{N-P+Q+i}^{j+1}\alpha^{(2j+1)n_{N-P+Q+i}}}

using the expressions for xNx^{\prime}_{N} and and the double sum given in Lemma 3.7 and Lemma 3.8. All the properties (i)-(vii) follow from our choice and properties of the expressions for xNx_{N}, xNx^{\prime}_{N}, and the double sum given in the previous section. ∎

Remark 4.2.

In the proof of Proposition 4.1, we have that D+E+FD+E+F is exactly the RHS of (i) and the γi\gamma_{i}’s are exactly ss. However, relaxing these as in Proposition 4.1 allows us to work with more possible collections of data in order to choose a minimal one.

Remark 4.3.

Note that we allow any (or even all) of the D,E,FD,E,F to be 0 in the statement of Proposition 4.1. For example, if D=E=F=0D=E=F=0 then all the tuples are empty, the properties (i)-(vi) are vacuously true, and property (vii) becomes:

|(ssNP)xN,1αx(N,1)|<|α|(11cPQ1(c1))nN+o(nN).\displaystyle\left|(s-s_{N-P})x_{N,1}\alpha^{x(N,1)}\right|<|\alpha|^{-(1-\frac{1}{c^{P-Q-1}(c-1)})n_{N}+o(n_{N})}.

In the proof of Proposition 4.1, we prove the existence of the required data by crudely expanding out terms in xNx_{N}, xNx^{\prime}_{N}, and those in the double sum without any simplification whatsoever. The key trick is the following:

Definition 4.4.

Among all the collections of data (D,E,F,𝒩,)(D,E,F,\mathscr{N},\ldots) satisfying properties (i)-(vii) in Proposition 4.1, we choose one for which D+E+FD+E+F is minimal. By abusing the notation, we still use the same notation DD, EE, FF, 𝒩\mathscr{N}, γj\gamma_{j}’s, dN,id_{N,i}’s, d(N,i)d(N,i)’s, eN,je_{N,j}’s, e(N,j)e(N,j)’s, fN,f_{N,\ell}’s, and f(N,)f(N,\ell)’s for this chosen data with minimal D+E+FD+E+F.

Remark 4.5.

This trick is similar to the one in [KMN19, Proposition 3.4] in which the authors worked with a vector space with the minimal dimension among a certain family of finite-dimensional vector spaces so that any further non-trivial linear relation would not be possible.

Lemma 4.6.

There are at most finitely many NN in 𝒩\mathscr{N} such that one of the terms sNPxN,iαx(N,i)s_{N-P}x_{N,i}\alpha^{x(N,i)}, γjeN,jαe(N,j)\gamma_{j}e_{N,j}\alpha^{e(N,j)}, fN,αf(N,)f_{N,\ell}\alpha^{f(N,\ell)} for 1iD1\leq i\leq D, 1jE1\leq j\leq E, and 1F1\leq\ell\leq F is zero.

Proof.

If there is a term that is zero for an infinite subset 𝒩\mathscr{N}^{\prime} of 𝒩\mathscr{N}, then we have a new collection of data in which 𝒩\mathscr{N} is replaced by 𝒩\mathscr{N}^{\prime} and that zero term is removed. This violates the minimality of D+E+FD+E+F. ∎

The point of Definition 4.4 is that any non-trivial linear relation among the xN,1αx(N,1)x_{N,1}\alpha^{x(N,1)}, sNPxN,1αx(N,1)s_{N-P}x_{N,1}\alpha^{x(N,1)}, sNPdN,iαd(N,i)s_{N-P}d_{N,i}\alpha^{d(N,i)}, eN,jαe(N,j)e_{N,j}\alpha^{e(N,j)}, and fN,αf(N,)f_{N,\ell}\alpha^{f(N,\ell)} that holds for infinitely many NN must involve the first 2 terms.

Proposition 4.7.

Suppose there exist an infinite subset 𝒩\mathscr{N}^{\prime} of 𝒩\mathscr{N} and complex numbers λ1\lambda_{1}, λ2\lambda_{2}, d~i\tilde{d}_{i}, e~j\tilde{e}_{j}, f~\tilde{f}_{\ell} for 1iD1\leq i\leq D, 1jE1\leq j\leq E, and 1F1\leq\ell\leq F not all of which are zero such that:

(10) λ1xN,1αx(N,1)+λ2sNPxN,1αx(N,1)+i=1Dd~isNPdN,iαd(N,i)+j=1Ee~jeN,jαe(N,j)+=1Ff~fN,αf(N,)=0\displaystyle\begin{split}&\lambda_{1}x_{N,1}\alpha^{x(N,1)}+\lambda_{2}s_{N-P}x_{N,1}\alpha^{x(N,1)}+\sum_{i=1}^{D}\tilde{d}_{i}s_{N-P}d_{N,i}\alpha^{d(N,i)}\\ &+\sum_{j=1}^{E}\tilde{e}_{j}e_{N,j}\alpha^{e(N,j)}+\sum_{\ell=1}^{F}\tilde{f}_{\ell}f_{N,\ell}\alpha^{f(N,\ell)}=0\end{split}

for every N𝒩N\in\mathscr{N}^{\prime}. We have:

  • (i)

    There exist κ1\kappa_{1}, κ2\kappa_{2}, d~i\tilde{d}^{\prime}_{i}, e~j\tilde{e}^{\prime}_{j}, f~\tilde{f}^{\prime}_{\ell} for 1iD1\leq i\leq D, 1jE1\leq j\leq E, and 1F1\leq\ell\leq F not all of which are zero with the following properties:

    • (a)

      All the κ1\kappa_{1}, κ2\kappa_{2}, d~i\tilde{d}^{\prime}_{i}, e~j\tilde{e}^{\prime}_{j}, and f~\tilde{f}^{\prime}_{\ell} are in (α){\mathbb{Q}}(\alpha).

    • (b)

      For every N𝒩N\in\mathscr{N}^{\prime}:

      (11) κ1xN,1αx(N,1)+κ2sNPxN,1αx(N,1)+i=1Dd~isNPdN,iαd(N,i)+j=1Ee~jeN,jαe(N,j)+=1Ff~fN,αf(N,)=0\displaystyle\begin{split}&\kappa_{1}x_{N,1}\alpha^{x(N,1)}+\kappa_{2}s_{N-P}x_{N,1}\alpha^{x(N,1)}+\sum_{i=1}^{D}\tilde{d}^{\prime}_{i}s_{N-P}d_{N,i}\alpha^{d(N,i)}\\ &+\sum_{j=1}^{E}\tilde{e}^{\prime}_{j}e_{N,j}\alpha^{e(N,j)}+\sum_{\ell=1}^{F}\tilde{f}^{\prime}_{\ell}f_{N,\ell}\alpha^{f(N,\ell)}=0\end{split}
    • (c)

      κ1κ20\kappa_{1}\kappa_{2}\neq 0.

  • (ii)

    s(α)s\in{\mathbb{Q}}(\alpha).

Proof.

For part (i), since the terms xN,1αx(N,1)x_{N,1}\alpha^{x(N,1)}, sNPxN,1αx(N,1)s_{N-P}x_{N,1}\alpha^{x(N,1)}, sNPdN,iαd(N,i)s_{N-P}d_{N,i}\alpha^{d(N,i)}, eN,jαe(N,j)e_{N,j}\alpha^{e(N,j)}, and fN,αf(N,)f_{N,\ell}\alpha^{f(N,\ell)} are in (α){\mathbb{Q}}(\alpha), this establishes the existence of the λ1\lambda_{1}, λ2\lambda_{2}, d~i\tilde{d}^{\prime}_{i}, e~j\tilde{e}^{\prime}_{j}, f~\tilde{f}^{\prime}_{\ell} satisfying properties (a) and (b). We now prove κ1κ20\kappa_{1}\kappa_{2}\neq 0.

First, assume that κ1=κ2=0\kappa_{1}=\kappa_{2}=0. This means:

(12) i=1Dd~isNPdN,iαd(N,i)+j=1Ee~jeN,jαe(N,j)+=1Ff~fN,αf(N,)=0\displaystyle\sum_{i=1}^{D}\tilde{d}^{\prime}_{i}s_{N-P}d_{N,i}\alpha^{d(N,i)}+\sum_{j=1}^{E}\tilde{e}^{\prime}_{j}e_{N,j}\alpha^{e(N,j)}+\sum_{\ell=1}^{F}\tilde{f}^{\prime}_{\ell}f_{N,\ell}\alpha^{f(N,\ell)}=0

for N𝒩N\in\mathscr{N}^{\prime}. Now assume that d~i0\tilde{d}^{\prime}_{i^{*}}\neq 0 for some ii^{*}. Then for N𝒩N\in\mathscr{N}^{\prime}, equation (12) allows us to express sNPdN,iαd(N,i)s_{N-P}d_{N,i^{*}}\alpha^{d(N,i^{*})} as a linear combination of the sNPdN,iαd(N,i)s_{N-P}d_{N,i}\alpha^{d(N,i)} with iii\neq i^{*}, the eN,jαe(N,j)e_{N,j}\alpha^{e(N,j)}, and the fN,αf(N,)f_{N,\ell}\alpha^{f(N,\ell)} with coefficients in (α){\mathbb{Q}}(\alpha). This allows us to come up with a new data satisfying the properties in Proposition 4.1 in which 𝒩\mathscr{N} is replaced by 𝒩\mathscr{N}^{\prime} and DD is replaced by D1D-1. This contradicts the minimality of D+E+FD+E+F. Therefore dN,i=0d_{N,i}=0 for 1iD1\leq i\leq D and every N𝒩N\in\mathscr{N}^{\prime}. Similarly fN,=0f_{N,\ell}=0 for 1F1\leq\ell\leq F and every N𝒩N\in\mathscr{N}^{\prime}. So now (12) becomes:

j=1Ee~jeN,jαe(N,j)=0for N𝒩.\sum_{j=1}^{E}\tilde{e}^{\prime}_{j}e_{N,j}\alpha^{e(N,j)}=0\ \text{for $N\in\mathscr{N}^{\prime}$.}

Arguing as before, we obtain a contradiction to the minimality of D+E+FD+E+F. This proves at least κ1\kappa_{1} or κ2\kappa_{2} is non-zero.

We emphasize that the arguments should be run in the above order (i.e. obtaining dN,i=fN,=0d_{N,i}=f_{N,\ell}=0 first). Suppose one tried to prove all the eN,j=0e_{N,j}=0 first by using (12) to express some eN,jαe(N,j)e_{N,j^{*}}\alpha^{e(N,j^{*})} as a linear combination of the sNPdN,iαd(N,i)s_{N-P}d_{N,i}\alpha^{d(N,i)}, the eN,jαe(N,j)e_{N,j}\alpha^{e(N,j)} with jjj\neq j^{*}, and the fN,αf(N,)f_{N,\ell}\alpha^{f(N,\ell)}. Then due to the term γjeN,jαe(N,j)\gamma_{j^{*}}e_{N,j^{*}}\alpha^{e(N,j^{*})} in the LHS of (9) and since at the moment we do not necessarily have γj(α)\gamma_{j^{*}}\in{\mathbb{Q}}(\alpha), the “new” dN,id_{N,i} and fN,f_{N,\ell} would not remain in (α){\mathbb{Q}}(\alpha) and the new data would not satisfy all the properties in Proposition 4.1.

Suppose κ10\kappa_{1}\neq 0 while κ2=0\kappa_{2}=0. We have

κ1xN,1αx(N,1)+i=1Dd~isNPdN,iαd(N,i)+j=1Ee~jeN,jαe(N,j)+=1Ff~fN,αf(N,)=0\kappa_{1}x_{N,1}\alpha^{x(N,1)}+\sum_{i=1}^{D}\tilde{d}^{\prime}_{i}s_{N-P}d_{N,i}\alpha^{d(N,i)}+\sum_{j=1}^{E}\tilde{e}^{\prime}_{j}e_{N,j}\alpha^{e(N,j)}+\sum_{\ell=1}^{F}\tilde{f}^{\prime}_{\ell}f_{N,\ell}\alpha^{f(N,\ell)}=0

for N𝒩N\in\mathscr{N}^{\prime}. We divide by xN,1αx(N,1)x_{N,1}\alpha^{x(N,1)} and note that each of d(N,i)x(N,1)d(N,i)-x(N,1), e(N,j)x(N,1)e(N,j)-x(N,1), and f(N,)x(N,1)f(N,\ell)-x(N,1) is at most nNP+1nN/5P1-n_{N-P+1}\leq-n_{N}/5^{P-1} by (iii) in Proposition 4.1 and Proposition 3.4 while each of |sNPdN,i/xN,1||s_{N-P}d_{N,i}/x_{N,1}|, |eN,j/xN,1||e_{N,j}/x_{N,1}|, and |fN,/xN,1||f_{N,\ell}/x_{N,1}| is eo(nN)e^{o(n_{N})}. Therefore taking limit as NN\to\infty (and N𝒩N\in\mathscr{N}^{\prime}), we get κ1=0\kappa_{1}=0, contradiction. The case κ20\kappa_{2}\neq 0 and κ1=0\kappa_{1}=0 is ruled out in a similar way. This finishes the proof of (c).

For part (ii), we use (11), divide by xN,1αx(N,1)x_{N,1}\alpha^{x(N,1)} and let NN\to\infty as above to obtain

κ1+κ2s=0\kappa_{1}+\kappa_{2}s=0

and this proves s=κ1/κ2(α)s=-\kappa_{1}/\kappa_{2}\in{\mathbb{Q}}(\alpha). ∎

Proposition 4.8.

The number ss is in (α){\mathbb{Q}}(\alpha).

Proof.

We will obtain a non-trivial linear relation as in the statement of Proposition 4.7 for infinitely many NN and apply part (ii).

Let vv_{\infty} be the valuation on (α){\mathbb{Q}}(\alpha) corresponding to the usual |||\cdot| and let ww be the other archimedean one. Note that we follow the normalization in [BG06, Chapter 1], hence:

|x|v=|x|1/2and|x|w=|σ(x)|1/2.|x|_{v_{\infty}}=|x|^{1/2}\ \text{and}\ |x|_{w}=|\sigma(x)|^{1/2}.

The archimedean valuations on KK are denoted as v1,,vmv_{1},\ldots,v_{m} and w1,,wmw_{1},\ldots,w_{m} where the viv_{i}’s lie above vv_{\infty} and the wiw_{i}’s lie above ww. They correspond to the following real or one for each pair of complex-conjugate embeddings of KK into {\mathbb{C}}: τ1,,τm\tau_{1},\ldots,\tau_{m} and σ1,,σm\sigma_{1},\ldots,\sigma_{m}. In other words:

|x|vi=|τi(x)|d(vi)/[K:]and|x|wi=|σi(x)|d(wi)/[K:]|x|_{v_{i}}=|\tau_{i}(x)|^{d(v_{i})/[K:{\mathbb{Q}}]}\ \text{and}\ |x|_{w_{i}}=|\sigma_{i}(x)|^{d(w_{i})/[K:{\mathbb{Q}}]}

where d(vi)=[Kvi:]=1d(v_{i})=[K_{v_{i}}:{\mathbb{R}}]=1 or 22 depending on whether viv_{i} is real or complex and a similar definition for d(wi)d(w_{i}). Note that the τi\tau_{i}’s restrict to the identity automorphism on (α){\mathbb{Q}}(\alpha) while the σi\sigma_{i}’s restrict to σ\sigma on (α){\mathbb{Q}}(\alpha). In fact, since K/K/{\mathbb{Q}} is Galois, either all archimedean valuations are real or all are complex and we simply let δ\delta be the common value of the d(vi)d(v_{i}) and d(wi)d(w_{i}). We have:

(13) i=1md(vi)=i=1md(wi)=mδ=[K:(α)]=[K:]2.\sum_{i=1}^{m}d(v_{i})=\sum_{i=1}^{m}d(w_{i})=m\delta=[K:{\mathbb{Q}}(\alpha)]=\frac{[K:{\mathbb{Q}}]}{2}.

Our next step is to apply the Subspace Theorem using (9). Fix ϵ>0\epsilon>0 that will be specified later. Let

S=MK={v1,,vm,w1,,wm}.S=M_{K}^{\infty}=\{v_{1},\ldots,v_{m},w_{1},\ldots,w_{m}\}.

We will work with linear forms in variables

(T1,T2,X1,,XD,Y1,,YE,Z1,,ZF)(T_{1},T_{2},X_{1},\ldots,X_{D},Y_{1},\ldots,Y_{E},Z_{1},\ldots,Z_{F})

and the vectors

𝐯N=(\displaystyle{\mathbf{v}}_{N}=( xN,1αx(N,1),sNPxN,1,(sNPdN,iαd(N,i))1iD,(eN,jαe(N,j))1jE,\displaystyle x_{N,1}\alpha^{x(N,1)},-s_{N-P}x_{N,1},(-s_{N-P}d_{N,i}\alpha^{d(N,i)})_{1\leq i\leq D},(e_{N,j}\alpha^{e(N,j)})_{1\leq j\leq E},
(fN,αf(N,))1F)\displaystyle(-f_{N,\ell}\alpha^{f(N,\ell)})_{1\leq\ell\leq F})

for N𝒩N\in\mathscr{N}.

For each vSv\in S, the linear forms are denoted: Lv,T,1L_{v,T,1}, Lv,T,2L_{v,T,2}, Lv,X,iL_{v,X,i} for 1iD1\leq i\leq D, Lv,Y,jL_{v,Y,j} for 1jE1\leq j\leq E, and Lv,Z,L_{v,Z,\ell} for 1F1\leq\ell\leq F. The reason is that they will be defined as follows:

  • For any vSv\in S, Lv,T,2=T2L_{v,T,2}=T_{2}, Lv,X,i=XiL_{v,X,i}=X_{i}, Lv,Y,j=YjL_{v,Y,j}=Y_{j}, and Lv,Z,=ZL_{v,Z,\ell}=Z_{\ell} for any i,j,i,j,\ell.

  • If v=wkv=w_{k} for some 1km1\leq k\leq m, define Lv,T,1=T1L_{v,T,1}=T_{1} and we have

    |Lv,T,1(𝐯N)|v=|σ(xN,1αx(N,1))|δ/[K:]=o(1)|L_{v,T,1}({\mathbf{v}}_{N})|_{v}=|\sigma(x_{N,1}\alpha^{x(N,1)})|^{\delta/[K:{\mathbb{Q}}]}=o(1)

    since |σ(xN,1)|=eo(nN)|\sigma(x_{N,1})|=e^{o(n_{N})} while |σ(αx(N,1))|=|α|x(N,1)|\sigma(\alpha^{x(N,1)})|=|\alpha|^{-x(N,1)} and x(N,1)nNx(N,1)\gg n_{N}.

  • If v=vkv=v_{k} for some 1km1\leq k\leq m, define:

    Lv,T,1=\displaystyle L_{v,T,1}= τk1(s)T1+T2+X1++XD+τk1(γ1)Y1++τk1(γE)YE\displaystyle\tau_{k}^{-1}(s)T_{1}+T_{2}+X_{1}+\ldots+X_{D}+\tau_{k}^{-1}(\gamma_{1})Y_{1}+\ldots+\tau_{k}^{-1}(\gamma_{E})Y_{E}
    +Z1++ZF\displaystyle+Z_{1}+\ldots+Z_{F}

    so that |Lv,T,1(𝐯N)|v|L_{v,T,1}({\mathbf{v}}_{N})|_{v} is exactly the LHS of (9) to the power δ/[K:]{\delta/[K:{\mathbb{Q}}]}. Therefore

    |Lv,T,1(𝐯N)|v<|α|(11cPQ1(c1)ϵ)nNδ/[K:]|L_{v,T,1}({\mathbf{v}}_{N})|_{v}<|\alpha|^{-(1-\frac{1}{c^{P-Q-1}(c-1)}-\epsilon)n_{N}\delta/[K:{\mathbb{Q}}]}

    for all sufficiently large N𝒩N\in\mathscr{N}.

Combining with  (13), we have:

(14) vS|Lv,T,1(𝐯N)|v=O(|α|(11cPQ1(c1)ϵ)nN2).\prod_{v\in S}|L_{v,T,1}({\mathbf{v}}_{N})|_{v}=O(|\alpha|^{-(1-\frac{1}{c^{P-Q-1}(c-1)}-\epsilon)\frac{n_{N}}{2}}).

Now since α\alpha is an SS-unit (i.e. usual algebraic integer unit), by the product formula together with the fact that |sNP|=O(1)|s_{N-P}|=O(1) and the xN,1x_{N,1}, dN,id_{N,i}, eN,je_{N,j}, fN,f_{N,\ell} have height o(nN)o(n_{N}), we have that for all sufficiently large N𝒩N\in\mathscr{N}:

(15) vSL|L(𝐯N)|v=O(|α|(11cPQ1(c1)2ϵ)nN2)\prod_{v\in S}\prod_{L}|L({\mathbf{v}}_{N})|_{v}=O(|\alpha|^{-(1-\frac{1}{c^{P-Q-1}(c-1)}-2\epsilon)\frac{n_{N}}{2}})

where LL ranges over all the Lv,T,1L_{v,T,1}, Lv,T,2L_{v,T,2}, Lv,X,iL_{v,X,i}, Lv,Y,jL_{v,Y,j}, and Lv,Z,L_{v,Z,\ell}. Recall that 𝐯~N\tilde{{\mathbf{v}}}_{N} denotes the point in the projective space with coordinates 𝐯N{\mathbf{v}}_{N}. Since the |d(N,i)||d(N,i)|, |e(N,j)||e(N,j)|, and |f(N,)||f(N,\ell)| are less than 3nN3n_{N}, we have

H(𝐯~N)|α|5nNH(\tilde{{\mathbf{v}}}_{N})\leq|\alpha|^{5n_{N}}

for every N𝒩N\in\mathscr{N}.

We need to obtain ϵ>0\epsilon^{\prime}>0 such that:

(16) vSL|L(𝐯N)|v|𝐯N|v<H(𝐯~N)DEF2ϵ\prod_{v\in S}\prod_{L}\frac{|L({\mathbf{v}}_{N})|_{v}}{|{\mathbf{v}}_{N}|_{v}}<H(\tilde{{\mathbf{v}}}_{N})^{-D-E-F-2-\epsilon^{\prime}}

for all large NN in 𝒩\mathscr{N}. We have:

(17) H(𝐯~N)D+E+F+2vSL1|𝐯N|v=(vMKS|𝐯N|v)D+E+F+2.H(\tilde{{\mathbf{v}}}_{N})^{D+E+F+2}\prod_{v\in S}\prod_{L}\frac{1}{|{\mathbf{v}}_{N}|_{v}}=\left(\prod_{v\in M_{K}\setminus S}|{\mathbf{v}}_{N}|_{v}\right)^{D+E+F+2}.

Since all the xN,1x_{N,1}, dN,id_{N,i}, eN,je_{N,j}, and fN,f_{N,\ell} have multiplicative height eo(nN)e^{o(n_{N})}, we have:

(18) (vMKS|𝐯N|v)D+E+F+2H(sNP)D+E+F+2|α|ϵnN\left(\prod_{v\in M_{K}\setminus S}|{\mathbf{v}}_{N}|_{v}\right)^{D+E+F+2}\leq H(s_{N-P})^{D+E+F+2}|\alpha|^{\epsilon n_{N}}

for all sufficiently large N𝒩N\in\mathscr{N}. We have:

H(sNp)|α|n1++nNP+o(nNP)|α|nNP+1H(s_{N-p})\leq|\alpha|^{n_{1}+\ldots+n_{N-P}+o(n_{N-P})}\leq|\alpha|^{n_{N-P+1}}

for all large N𝒩N\in\mathscr{N}. From Proposition 4.1 and the definition of the DN,Q+iD_{N,Q+i}’s, we have:

D+E+F+2\displaystyle D+E+F+2 2Q+1+Q2Q1+2Qi=1PQ1(DN,Q+i+1)\displaystyle\leq 2^{Q+1}+Q2^{Q-1}+2^{Q}\sum_{i=1}^{P-Q-1}(D_{N,Q+i}+1)
2Q(P(Q/2)+1)+2Qi=1PQ1nN2nNP+Q+i\displaystyle\leq 2^{Q}(P-(Q/2)+1)+2^{Q}\sum_{i=1}^{P-Q-1}\frac{n_{N}}{2n_{N-P+Q+i}}
2QP+2Q1i=1PQ1nNnNP+Q+i\displaystyle\leq 2^{Q}P+2^{Q-1}\sum_{i=1}^{P-Q-1}\frac{n_{N}}{n_{N-P+Q+i}}

assuming Q2Q\geq 2. Hence for all large N𝒩N\in\mathscr{N},

(19) H(sNP)D+E+F+2|α|ΩNH(s_{N-P})^{D+E+F+2}\leq|\alpha|^{\Omega_{N}}

where

(20) ΩN:=nNP+12QP+2Q1(i=1PQ1nNP+1nNP+Q+i)nN2QPcP1nN+2Q1(i=1PQ11cQ+i1)nN(2QPcP1+2Q1cQ11(1/c))nN.\displaystyle\begin{split}\Omega_{N}:&=n_{N-P+1}2^{Q}P+2^{Q-1}\left(\sum_{i=1}^{P-Q-1}\frac{n_{N-P+1}}{n_{N-P+Q+i}}\right)n_{N}\\ &\leq\frac{2^{Q}P}{c^{P-1}}n_{N}+2^{Q-1}\left(\sum_{i=1}^{P-Q-1}\frac{1}{c^{Q+i-1}}\right)n_{N}\\ &\leq\left(\frac{2^{Q}P}{c^{P-1}}+\frac{2^{Q-1}}{c^{Q}}\cdot\frac{1}{1-(1/c)}\right)n_{N}.\end{split}

Finally, note that:

(21) H(𝐯~N)ϵ|α|5ϵnN.H(\tilde{{\mathbf{v}}}_{N})^{-\epsilon^{\prime}}\geq|\alpha|^{-5\epsilon^{\prime}n_{N}}.

In order to obtain ϵ\epsilon^{\prime} satisfying (16), we combine (15) and (17)–(21) and require that:

(22) (2QPcP1+2Q1cQ11(1/c))+ϵ12(11cPQ1(c1)2ϵ)<5ϵ.\left(\frac{2^{Q}P}{c^{P-1}}+\frac{2^{Q-1}}{c^{Q}}\cdot\frac{1}{1-(1/c)}\right)+\epsilon-\frac{1}{2}\left(1-\frac{1}{c^{P-Q-1}(c-1)}-2\epsilon\right)<-5\epsilon^{\prime}.

Such an ϵ\epsilon^{\prime} exists as long as the LHS is negative. Therefore, at the beginning of Section 3, we choose sufficiently large integers 2Q<P2\leq Q<P and here we choose a sufficiently small ϵ>0\epsilon>0 such that:

(23) 2QPcP1+2Q1cQ11(1/c)+2ϵ+12cPQ1(c1)<12;\frac{2^{Q}P}{c^{P-1}}+\frac{2^{Q-1}}{c^{Q}}\cdot\frac{1}{1-(1/c)}+2\epsilon+\frac{1}{2c^{P-Q-1}(c-1)}<\frac{1}{2};

this is possible since c>2c>2. Then we can apply the Subspace Theorem to have that there exists a non-trivial linear relation satisfied by the coordinates of 𝐯N{\mathbf{v}}_{N} for infinitely many N𝒩N\in\mathscr{N}. Then we use part (ii) of Proposition 4.7 to finish the proof. ∎

5. The proof of Theorem 1.3

We continue with the notations of Section 3 and ignore those in Section 4. We no longer assume the choice of PP, QQ, and ϵ\epsilon as in (22) and (23). However, we use the crucial result that s(α)s\in{\mathbb{Q}}(\alpha). While the arguments in Section 4 are valid for any sufficiently large Q<PQ<P (and sufficiently small ϵ\epsilon), those in this section require that Q=P1Q=P-1:

Assumption 5.1.

From now on, Q=P1Q=P-1. Therefore xN=i=1P1UNP+ix_{N}=\displaystyle\prod_{i=1}^{P-1}U_{N-P+i}, the x(N,i)sx(N,i)^{\prime}s are the numbers ±nNP+1±nNP+2±nN1\pm n_{N-P+1}\pm n_{N-P+2}\cdots\pm n_{N-1}, xNxN=i=1P1CNP+iUNP+i\displaystyle\frac{x^{\prime}_{N}}{x_{N}}=\sum_{i=1}^{P-1}\frac{C_{N-P+i}}{U_{N-P+i}}, and most importantly:

(24) (ssNP)xNxN=(ssN1)xN=xNk=NCkUk.(s-s_{N-P})x_{N}-x^{\prime}_{N}=(s-s_{N-1})x_{N}=x_{N}\sum_{k=N}^{\infty}\frac{C_{k}}{U_{k}}.

The following numbers x(N,+)x(N,+), x(N,)x(N,-), x~(N,+)\tilde{x}(N,+), and x~(N,)\tilde{x}(N,-) will play an important role:

Lemma 5.2.
  • (i)

    x(N,+):=nNP+1nN2+nN1x(N,+):=-n_{N-P+1}-\ldots-n_{N-2}+n_{N-1} is the smallest non-negative numbers while x(N,):=x(N,+)x(N,-):=-x(N,+) is the largest non-positive numbers among the x(N,i)x(N,i)’s.

  • (ii)

    Write x~(N,+):=nNP+1nN1+nN\tilde{x}(N,+):=-n_{N-P+1}-\ldots-n_{N-1}+n_{N} and write x~(N,):=x~(N,+)\tilde{x}(N,-):=-\tilde{x}(N,+). We have x~(N,)<x(N,)<x(N,+)<x~(N,+)\tilde{x}(N,-)<x(N,-)<x(N,+)<\tilde{x}(N,+). Moreover:

    x(N,+)x(N,)2(c2)c1nN12(c2)5(c1)nN,andx(N,+)-x(N,-)\geq\frac{2(c-2)}{c-1}n_{N-1}\geq\frac{2(c-2)}{5(c-1)}n_{N},\ \text{and}
    x(N,)x~(N,)=x~(N,+)x(N,+)=nN2nN1c2cnNx(N,-)-\tilde{x}(N,-)=\tilde{x}(N,+)-x(N,+)=n_{N}-2n_{N-1}\geq\frac{c-2}{c}n_{N}

    for all sufficiently large NN.

  • (iii)

    |(ssN1)xN|=|α|x~(N,)+o(nN)|(s-s_{N-1})x_{N}|=|\alpha|^{\tilde{x}(N,-)+o(n_{N})} for all sufficiently large NN.

Proof.

We have:

(25) x(N,+)x(N,)=2(nN1nN2nNP+1)2(c2)c1nN12(c2)5(c1)nN\displaystyle\begin{split}x(N,+)-x(N,-)&=2(n_{N-1}-n_{N-2}-\ldots-n_{N-P+1})\\ &\geq\frac{2(c-2)}{c-1}n_{N-1}\geq\frac{2(c-2)}{5(c-1)}n_{N}\end{split}

for all large NN by Lemma 3.1 and Proposition 3.4. Part (i) and the rest of part (ii) are elementary. Part (iii) is simply Lemma 3.9(ii) when Q=P1Q=P-1. ∎

First, we prove the existence of a certain expression and then choose a minimal one:

Proposition 5.3.

Note that Q=P1Q=P-1. There exist integers D,F0D,F\geq 0, an infinite set 𝒩\mathscr{N}, tuples (dN,1,,dN,D)(d_{N,1},\ldots,d_{N,D}), (fN,1,,fN,F)(f_{N,1},\ldots,f_{N,F}), (d(N,1),,d(N,D))(d(N,1),\ldots,d(N,D)) and (f(N,1),,f(N,F))(f(N,1),\ldots,f(N,F)) for each N𝒩N\in\mathscr{N} with the following properties:

  • (i)

    D+F2P1+(P1)2P2D+F\leq 2^{P-1}+(P-1)2^{P-2}.

  • (ii)

    For every N𝒩N\in\mathscr{N}, the d(N,i)d(N,i)’s and f(N,j)f(N,j)’s are integers for every i,ji,j.

  • (iii)

    For every N𝒩N\in\mathscr{N}, maxi|d(N,i)|nNP+1++nN1\max_{i}|d(N,i)|\leq n_{N-P+1}+\ldots+n_{N-1}.

  • (iv)

    For every N𝒩N\in\mathscr{N}, maxj|f(N,j)|nNP+2++nN1\max_{j}|f(N,j)|\leq n_{N-P+2}+\ldots+n_{N-1}.

  • (v)

    For every N𝒩N\in\mathscr{N}, the dN,id_{N,i}’s and fN,jf_{N,j}’s are elements of (α){\mathbb{Q}}(\alpha).

  • (vi)

    As NN\to\infty, we have h(dN,i)/nN0h(d_{N,i})/n_{N}\to 0 and h(fN,j)/nn0h(f_{N,j})/n_{n}\to 0 for every i,ji,j.

  • (vii)

    For every N𝒩N\in\mathscr{N}, we have

    (26) i=1D(ssNP)dN,iαd(N,i)j=1FfN,jαf(N,j)=(ssN1)xN.\displaystyle\sum_{i=1}^{D}(s-s_{N-P})d_{N,i}\alpha^{d(N,i)}-\sum_{j=1}^{F}f_{N,j}\alpha^{f(N,j)}=(s-s_{N-1})x_{N}.
Proof.

The proof is very similar to the proof of Proposition 4.1. Choose 𝒩\mathscr{N} to be the set of all sufficiently large integers and D=2P1D=2^{P-1}. We want the sum i=1D(ssNP)dN,iαd(N,i)\displaystyle\sum_{i=1}^{D}(s-s_{N-P})d_{N,i}\alpha^{d(N,i)} to be (ssNP)xN(s-s_{N-P})x_{N}. We simply choose the dN,id_{N,i} and d(N,i)d(N,i) to be the xN,ix_{N,i} and x(N,i)x(N,i).

Then we choose F=(P1)2P2F=(P-1)2^{P-2} and we want the sum j=1FfN,jαf(n,j)\displaystyle\sum_{j=1}^{F}f_{N,j}\alpha^{f(n,j)} to be

xN=j=1(P1)2P2xN,jαx(N,j),x^{\prime}_{N}=\sum_{j=1}^{(P-1)2^{P-2}}x^{\prime}_{N,j}\alpha^{x^{\prime}(N,j)},

so we simply take the fN,jf_{N,j} and f(N,j)f(N,j) to be the xN,jx^{\prime}_{N,j} and x(N,j)x(N,j). The properties (i)-(vi) follow from (24) and Lemma 3.7. ∎

Definition 5.4.

Among all the collections of data (D,F,𝒩,)(D,F,\mathscr{N},\ldots) satisfying properties (i)-(vii) in Proposition 5.3, we choose one for which D+FD+F is minimal. By abusing the notation, we still use the same notation DD, FF, 𝒩\mathscr{N}, dN,id_{N,i}’s, d(N,i)d(N,i)’s, fN,jf_{N,j}’s, and f(N,j)f(N,j)’s for this chosen data with minimal D+FD+F.

Remark 5.5.

As before, we allow the possibility that DD or FF is 0. Note that the scenario D=F=0D=F=0 cannot happen since the RHS of (26) is non-zero.

Lemma 5.6.

There are at most finitely many NN in 𝒩\mathscr{N} such that one of the terms (ssNP)dN,iαd(N,i)(s-s_{N-P})d_{N,i}\alpha^{d(N,i)}, fN,jαf(N,j)f_{N,j}\alpha^{f(N,j)} for 1iD1\leq i\leq D and 1jF1\leq j\leq F is zero.

Proof.

This is similar to the proof of Lemma 4.6. ∎

The reason we introduce the number x~(N,)\tilde{x}(N,-) is that the d(N,i)d(N,i) for 1iD1\leq i\leq D and f(N,j)f(N,j) for 1jF1\leq j\leq F are less than x~(N,)+o(nN)\tilde{x}(N,-)+o(n_{N}) for an appropriate choice of PP. This means that for any ϵ>0\epsilon>0, as long as PP is sufficiently large (depending on ϵ\epsilon), then the d(N,i)d(N,i)’s and f(N,j)f(N,j)’s are smaller than x~(N,)+ϵnN\tilde{x}(N,-)+\epsilon n_{N}. For our purpose, we state and prove the next result for the specific value c22c\displaystyle\frac{c-2}{2c} for ϵ\epsilon; note that c22cnN\displaystyle\frac{c-2}{2c}n_{N} is at most half of the gap between x~(N,)\tilde{x}(N,-) and x(N,)x(N,-) thanks to Lemma 5.2.

Proposition 5.7.

Assume that PP satisfies:

(27) 2P1+(P1)2P2cP1<c24c.\frac{2^{P-1}+(P-1)2^{P-2}}{c^{P-1}}<\frac{c-2}{4c}.

Then for all but finitely many N𝒩N\in\mathscr{N}, we have d(N,i)x~(N,)+c22cnN\displaystyle d(N,i)\leq\tilde{x}(N,-)+\frac{c-2}{2c}n_{N} and f(N,j)x~(N,)+c22cnNf(N,j)\leq\displaystyle\tilde{x}(N,-)+\frac{c-2}{2c}n_{N} for 1iD1\leq i\leq D and 1jF1\leq j\leq F.

Proof.

We prove by contradiction and without loss of generality, we only need to consider 2 cases:

  • Case 1: there exists an infinite subset 𝒩\mathscr{N}^{\prime} of 𝒩\mathscr{N} such that d(N,1)>x~(N,)+c22cnNd(N,1)>\tilde{x}(N,-)+\frac{c-2}{2c}n_{N} for every N𝒩N\in\mathscr{N}^{\prime}.

  • Case 2: there exists an infinite subset 𝒩\mathscr{N}^{\prime} of 𝒩\mathscr{N} such that f(N,1)>x~(N,)+c22cnNf(N,1)>\tilde{x}(N,-)+\frac{c-2}{2c}n_{N} for every N𝒩N\in\mathscr{N}^{\prime}.

First, we assume Case 1. Let ϵ>0\epsilon>0 be a small number that will be specified later. For all sufficiently large N𝒩N\in\mathscr{N}^{\prime}, we have:

(28) |i=1D(ssNP)dN,iαd(N,i)j=1FfN,jαf(N,j)|=|(ssN1)xN|<|α|x~(N,)+ϵnN\displaystyle\begin{split}|\sum_{i=1}^{D}(s-s_{N-P})d_{N,i}\alpha^{d(N,i)}-\sum_{j=1}^{F}f_{N,j}\alpha^{f(N,j)}|&=|(s-s_{N-1})x_{N}|\\ &<|\alpha|^{\tilde{x}(N,-)+\epsilon n_{N}}\end{split}

by (26) and Lemma 5.2. We now apply the Subspace Theorem over the field (α){\mathbb{Q}}(\alpha). Let S={v,w}S=\{v_{\infty},w\} be its archimedean places as described in the proof of Proposition 4.8. We work with linear forms in the variables:

(X1,,XD,Y1,YF)(X_{1},\ldots,X_{D},Y_{1}\ldots,Y_{F})

and the vectors

𝐯N=(((ssNP)dN,iαd(N,i))1iD,(fN,jαf(N,j))1jF)\displaystyle{\mathbf{v}}_{N}=(((s-s_{N-P})d_{N,i}\alpha^{d(N,i)})_{1\leq i\leq D},(-f_{N,j}\alpha^{f(N,j)})_{1\leq j\leq F})

for large N𝒩N\in\mathscr{N}^{\prime}. For vSv\in S, the linear forms are denoted Lv,X,iL_{v,X,i} for 1iD1\leq i\leq D, and Lv,Y,jL_{v,Y,j} for 1jF1\leq j\leq F. They are defined as follows:

  • For any vSv\in S, Lv,X,i=XiL_{v,X,i}=X_{i} for 2iD2\leq i\leq D and Lv,Y,j=YjL_{v,Y,j}=Y_{j} for 1jF1\leq j\leq F.

  • If v=wv=w, define Lv,X,1=X1L_{v,X,1}=X_{1}.

  • If v=vv=v_{\infty}, define

    Lv,X,1=X1++XD+Y1++YF.L_{v,X,1}=X_{1}+\ldots+X_{D}+Y_{1}+\ldots+Y_{F}.

Since ss is irrational by Corollary 3.3 and |ssNP|=o(1)|s-s_{N-P}|=o(1), we have |ssNP|w=O(1)|s-s_{N-P}|_{w}=O(1). Together with the fact that the dN,id_{N,i}’s and fN,jf_{N,j}’s have multiplicative height eo(nN)e^{o(n_{N})}, for all large N𝒩N\in\mathscr{N}^{\prime}, we have:

(29) vS|Lv,X,1(𝐯N)|v<|α|(d(N,1)+x~(N,)+ϵnN)/2<|α|((c2)/(2c)+ϵ)nN/2\prod_{v\in S}|L_{v,X,1}({\mathbf{v}}_{N})|_{v}<|\alpha|^{(-d(N,1)+\tilde{x}(N,-)+\epsilon n_{N})/2}<|\alpha|^{(-(c-2)/(2c)+\epsilon)n_{N}/2}

where the last inequality follows from the assumption in Case 1.

As in the proof of Proposition 4.8 and using H(ssNP)=O(H(sNP))H(s-s_{N-P})=O(H(s_{N-P})), we have:

(30) H(𝐯~N)D+FvSL1|𝐯N|v=(vMKS|𝐯N|v)D+FH(sNP)D+F|α|ϵnN|α|(D+F)nNP+1+ϵnN\displaystyle\begin{split}H(\tilde{{\mathbf{v}}}_{N})^{D+F}\prod_{v\in S}\prod_{L}\frac{1}{|{\mathbf{v}}_{N}|_{v}}&=\left(\prod_{v\in M_{K}\setminus S}|{\mathbf{v}}_{N}|_{v}\right)^{D+F}\\ &\leq H(s_{N-P})^{D+F}|\alpha|^{\epsilon n_{N}}\\ &\leq|\alpha|^{(D+F)n_{N-P+1}+\epsilon n_{N}}\end{split}

for all sufficiently large N𝒩N\in\mathscr{N}^{\prime}. Recall that D+F2P1+(P1)2P2D+F\leq 2^{P-1}+(P-1)2^{P-2}, as before, we can apply the Subspace Theorem if:

2P1+(P1)2P2cP1+ϵ+12(c22c+ϵ)<0\frac{2^{P-1}+(P-1)2^{P-2}}{c^{P-1}}+\epsilon+\frac{1}{2}\left(-\frac{c-2}{2c}+\epsilon\right)<0

or in other words

2P1+(P1)2P2cP1+32ϵ<c24c.\frac{2^{P-1}+(P-1)2^{P-2}}{c^{P-1}}+\frac{3}{2}\epsilon<\frac{c-2}{4c}.

We can choose such an ϵ\epsilon thanks to the given condition on PP. Then the Subspace Theorem yields a non-trivial linear relation over (α){\mathbb{Q}}(\alpha) among the (ssNP)dN,iαd(N,i)(s-s_{N-P})d_{N,i}\alpha^{d(N,i)}’s and fN,jαf(N,j)f_{N,j}\alpha^{f(N,j)}’s for NN in an infinite subset 𝒩′′\mathscr{N}^{\prime\prime} of 𝒩\mathscr{N}^{\prime}. This allows us to express one of them as a linear combination of the other terms and we obtain a new data satisfying the properties stated in Proposition 5.3 in which 𝒩\mathscr{N} is replaced by 𝒩′′\mathscr{N}^{\prime\prime} and D+FD+F is replaced by D+F1D+F-1, contradicting the minimality of D+FD+F. This shows that Case 1 cannot happen.

By similar arguments, we have that Case 2 cannot happen either. For Case 2, we consider the same 𝐯N{\mathbf{v}}_{N}’s, the variables X1,,XD,Y1,,YFX_{1},\ldots,X_{D},Y_{1},\ldots,Y_{F}, and SS as in Case 1 while the linear forms Lx,X,iL_{x,X,i}’s and Lv,Y,jL_{v,Y,j}’s are defined as folows:

  • For any vSv\in S, Lv,X,i=XiL_{v,X,i}=X_{i} for 1iD1\leq i\leq D and Lv,Y,j=YjL_{v,Y,j}=Y_{j} for 2jF2\leq j\leq F.

  • If v=wv=w, define Lv,Y,1=Y1L_{v,Y,1}=Y_{1}.

  • If v=vv=v_{\infty}, define Lv,Y,1=X1++XD+Y1++YFL_{v,Y,1}=X_{1}+\ldots+X_{D}+Y_{1}+\ldots+Y_{F}.

Then we proceed as before and arrive at a contradiction. ∎

We are now at the final stage of the proof of Theorem 1.3.

Notation 5.8.

Let i{1,,2P1}i^{-}\in\{1,\ldots,2^{P-1}\} such that that x(N,i)=x(N,)x(N,i^{-})=x(N,-). Let

I:={i{1,,2P1}:x(N,i)<x(N,i)}andI^{-}:=\{i\in\{1,\ldots,2^{P-1}\}:\ x(N,i)<x(N,i^{-})\}\text{and}
I+:={i{1,,2P1}:x(N,i)>x(N,i)}.I^{+}:=\{i\in\{1,\ldots,2^{P-1}\}:\ x(N,i)>x(N,i^{-})\}.

Note that |I|=2P21|I^{-}|=2^{P-2}-1 and |I+|=2P2|I^{+}|=2^{P-2}.

We apply σ\sigma to both sides of:

(31) i=1D(ssNP)dN,iαd(N,i)j=1FfN,jαf(N,j)=(ssN1)xN,\sum_{i=1}^{D}(s-s_{N-P})d_{N,i}\alpha^{d(N,i)}-\sum_{j=1}^{F}f_{N,j}\alpha^{f(N,j)}=(s-s_{N-1})x_{N},

and recall that σ(α)=β=±1α\sigma(\alpha)=\beta=\displaystyle\pm\frac{1}{\alpha} to get:

(32) i=1D(σ(s)sNP)σ(dN,i)(±1)d(N,i)αd(N,i)j=1Fσ(fN,j)(±1)f(N,j)αf(N,j)=(σ(s)sN1)xN.\displaystyle\begin{split}&\sum_{i=1}^{D}(\sigma(s)-s_{N-P})\sigma(d_{N,i})(\pm 1)^{d(N,i)}\alpha^{-d(N,i)}-\sum_{j=1}^{F}\sigma(f_{N,j})(\pm 1)^{f(N,j)}\alpha^{-f(N,j)}\\ &=(\sigma(s)-s_{N-1})x_{N}.\end{split}

Then taking the difference of the previous 2 equations, we get:

(33) (σ(s)s)xN=i=1D(σ(s)sNP)σ(dN,i)(±1)d(N,i)αd(N,i)j=1Fσ(fN,j)(±1)f(N,j)αf(N,j)i=1D(ssNP)dN,iαd(N,i)+j=1FfN,jαf(N,j).\displaystyle\begin{split}(\sigma(s)-s)x_{N}=&\sum_{i=1}^{D}(\sigma(s)-s_{N-P})\sigma(d_{N,i})(\pm 1)^{d(N,i)}\alpha^{-d(N,i)}\\ &-\sum_{j=1}^{F}\sigma(f_{N,j})(\pm 1)^{f(N,j)}\alpha^{-f(N,j)}\\ &-\sum_{i=1}^{D}(s-s_{N-P})d_{N,i}\alpha^{d(N,i)}+\sum_{j=1}^{F}f_{N,j}\alpha^{f(N,j)}.\end{split}

Our idea to conclude the proof of Theorem 1.3 is as follows. Note that σ(s)s0\sigma(s)-s\neq 0, hence the LHS of (33) contains the term (σ(s)s)xN,iαx(N,)(\sigma(s)-s)x_{N,i^{-}}\alpha^{x(N,-)}. Let δ2=2(c2)5P1(c1)\delta_{2}=\displaystyle\frac{2(c-2)}{5^{P-1}(c-1)} as in the proof of Lemma 3.7 (for Q=P1Q=P-1) and we have that there is a gap δ2nN\delta_{2}n_{N} between any two of the x(N,i)x(N,i)’s. However, since δ2\delta_{2} depends on PP, the previous method of choosing a sufficiently large PP does not work if we “play the x(N,i)x(N,i)’s against each other”. The whole point of Lemma 5.2 and Proposition 5.7 is that we now have a gap of at least c22cnN\displaystyle\frac{c-2}{2c}n_{N} between x(N,)x(N,-) and any of the ±d(N,i)\pm d(N,i)’s, ±f(N,j)\pm f(N,j)’s and a gap of at least 2(c2)5(c1)nN\displaystyle\frac{2(c-2)}{5(c-1)}n_{N} between x(N,)x(N,-) and any of the x(N,i)x(N,i) with iI+i\in I^{+}.

Proposition 5.9.

Recall δ2=2(c2)5P1(c1)\delta_{2}=\displaystyle\frac{2(c-2)}{5^{P-1}(c-1)} and let θ=min{c22c,2(c2)5(c1)}\theta=\min\left\{\displaystyle\frac{c-2}{2c},\displaystyle\frac{2(c-2)}{5(c-1)}\right\}. There exist integers A,B,C,V,W0A,B,C,V,W\geq 0, an infinite subset 𝒩\mathscr{N}^{\prime} of 𝒩\mathscr{N}, and tuples

  • (aN,1,,aN,A)(a_{N,1},\ldots,a_{N,A}), (a(N,1),,a(N,A))(a(N,1),\ldots,a(N,A)),

  • (bN,1,,bN,B)(b_{N,1},\ldots,b_{N,B}), (b(N,1),,b(N,B))(b(N,1),\ldots,b(N,B)),

  • (cN,1,,cN,C)(c_{N,1},\ldots,c_{N,C}), (c(N,1),,c(N,C))(c(N,1),\ldots,c(N,C)),

  • (vN,1,,vN,V)(v_{N,1},\ldots,v_{N,V}), (v(N,1),,v(N,V))(v(N,1),\ldots,v(N,V)),

  • (wN,1,,wN,W)(w_{N,1},\ldots,w_{N,W}) and (w(N,1),,w(N,W))(w(N,1),\ldots,w(N,W))

for each N𝒩N\in\mathscr{N}^{\prime} with the following properties:

  • (i)

    A+B+C+V+W2P11+2(D+F)2P+P2P11A+B+C+V+W\leq 2^{P-1}-1+2(D+F)\leq 2^{P}+P2^{P-1}-1.

  • (ii)

    For every N𝒩N\in\mathscr{N}^{\prime}, the a(N,i)a(N,i)’s, b(N,j)b(N,j)’s, c(N,k)c(N,k), v(N,)v(N,\ell)’s, and w(N,m)w(N,m)’s are integers for every i,j,k,,mi,j,k,\ell,m.

  • (iii)

    For every N𝒩N\in\mathscr{N}^{\prime}, we have:

    maxi,j,k,,m{|a(N,i)|,|b(N,j)|,|c(N,k)|,|v(N,)|,|w(N,m)|}<2nN.\max_{i,j,k,\ell,m}\{|a(N,i)|,|b(N,j)|,|c(N,k)|,|v(N,\ell)|,|w(N,m)|\}<2n_{N}.
  • (iv)

    For every N𝒩N\in\mathscr{N}^{\prime}, we have a(N,i)δ2nNa(N,i)\leq-\delta_{2}n_{N}, b(N,j)θnNb(N,j)\leq-\theta n_{N}, c(N,k)θnNc(N,k)\geq\theta n_{N}, v(N,)θnNv(N,\ell)\leq-\theta n_{N}, and w(N,m)θnNw(N,m)\geq\theta n_{N} for every i,j,k,,mi,j,k,\ell,m.

  • (v)

    For every N𝒩N\in\mathscr{N}^{\prime}, the aN,ia_{N,i}’s, bN,jb_{N,j}’s, cN,kc_{N,k}’s, vN,v_{N,\ell}’s, and wN,mw_{N,m}’s are elements of (α){\mathbb{Q}}(\alpha).

  • (vi)

    As NN\to\infty, we have h(aN,i)/nN0h(a_{N,i})/n_{N}\to 0, h(bN,j)/nN0h(b_{N,j})/n_{N}\to 0, h(cN,k)/nN0h(c_{N,k})/n_{N}\to 0, h(vN,)/nN0h(v_{N,\ell})/n_{N}\to 0, and h(wN,m)/nN0h(w_{N,m})/n_{N}\to 0 for every i,j,k,,mi,j,k,\ell,m.

  • (vii)

    For every N𝒩N\in\mathscr{N}^{\prime}, we have:

    (34) σ(s)s=i=1AaN,iαa(N,i)+j=1BbN,jαb(N,j)+k=1CcN,kαc(N,k)=1V(ssNP)vN,αv(N,)+m=1W(σ(s)sNP)wN,mαw(N,m).\displaystyle\begin{split}\sigma(s)-s=&\sum_{i=1}^{A}a_{N,i}\alpha^{a(N,i)}+\sum_{j=1}^{B}b_{N,j}\alpha^{b(N,j)}+\sum_{k=1}^{C}c_{N,k}\alpha^{c(N,k)}\\ &\sum_{\ell=1}^{V}(s-s_{N-P})v_{N,\ell}\alpha^{v(N,\ell)}+\sum_{m=1}^{W}(\sigma(s)-s_{N-P})w_{N,m}\alpha^{w(N,m)}.\end{split}
Proof.

We use (33) and the expression for xNx_{N} in Lemma 3.7, then divide both sides by xN,iαx(N,)x_{N,i^{-}}\alpha^{x(N,-)} to get:

(35) σ(s)s=iI(σ(s)s)xN,ixN,iαx(N,i)x(N,)iI+(σ(s)s)xN,ixN,iαx(N,i)x(N,i)i=1Fσ(fN,i)xN,i(±1)f(N,i)αf(N,i)x(N,)+i=1FfN,ixN,iαf(N,i)x(N,)i=1D(ssNP)dN,ixN,iαd(N,i)x(N,)+i=1D(σ(s)sNP)σ(dN,i)xN,i(±1)d(N,i)αd(N,i)x(N,)\displaystyle\begin{split}\sigma(s)-s=&-\sum_{i\in I^{-}}\frac{(\sigma(s)-s)x_{N,i}}{x_{N,i^{-}}}\alpha^{x(N,i)-x(N,-)}\\ &-\sum_{i\in I^{+}}\frac{(\sigma(s)-s)x_{N,i}}{x_{N,i^{-}}}\alpha^{x(N,i)-x(N,i^{-})}\\ &-\sum_{i=1}^{F}\frac{\sigma(f_{N,i})}{x_{N,i^{-}}}(\pm 1)^{f(N,i)}\alpha^{-f(N,i)-x(N,-)}\\ &+\sum_{i=1}^{F}\frac{f_{N,i}}{x_{N,i^{-}}}\alpha^{f(N,i)-x_{(}N,-)}\\ &-\sum_{i=1}^{D}(s-s_{N-P})\frac{d_{N,i}}{x_{N,i^{-}}}\alpha^{d(N,i)-x(N,-)}\\ &+\sum_{i=1}^{D}(\sigma(s)-s_{N-P})\frac{\sigma(d_{N,i})}{x_{N,i^{-}}}(\pm 1)^{d(N,i)}\alpha^{-d(N,i)-x(N,-)}\end{split}

Let 𝒩\mathscr{N}^{\prime} be the set of all sufficiently large N𝒩N\in\mathscr{N}; in the following, NN is an element of 𝒩\mathscr{N}^{\prime}. We want i=1AaN,iαa(N,i)\displaystyle\sum_{i=1}^{A}a_{N,i}\alpha^{a(N,i)} to be iI(σ(s)s)xN,ixN,iαx(N,i)x(N,)\displaystyle-\sum_{i\in I^{-}}\frac{(\sigma(s)-s)x_{N,i}}{x_{N,i^{-}}}\alpha^{x(N,i)-x(N,-)}. Therefore we let A=|I|=2P21A=|I^{-}|=2^{P-2}-1, let the aN,ia_{N,i} and a(N,i)a(N,i) for 1iA1\leq i\leq A be respectively the (σ(s)s)xN,ixN,i-\displaystyle\frac{(\sigma(s)-s)x_{N,i}}{x_{N,i^{-}}} and x(N,i)x(N,)x(N,i)-x(N,-) for iIi\in I^{-}. By Lemma 3.7 and the definition of II^{-}, we have a(N,i)δ2nNa(N,i)\leq-\delta_{2}n_{N}.

We want j=1BbN,jαb(N,j)\displaystyle\sum_{j=1}^{B}b_{N,j}\alpha^{b(N,j)} to be i=1FfN,ixN,iαf(N,i)x(N,)\displaystyle\sum_{i=1}^{F}\frac{f_{N,i}}{x_{N,i^{-}}}\alpha^{f(N,i)-x_{(}N,-)}. Therefore we let B=FB=F, let the bN,jb_{N,j} and b(N,j)b(N,j) for 1jB1\leq j\leq B be respectively the fN,ixN,i\displaystyle\frac{f_{N,i}}{x_{N,i^{-}}} and f(N,i)x(N,)f(N,i)-x(N,-) for 1iF1\leq i\leq F. By Lemma 5.2 and Proposition 5.7, we have:

f(N,i)x(N,)x~(N,)+c22cnNx(N,)c22cnNθnNf(N,i)-x(N,-)\leq\tilde{x}(N,-)+\frac{c-2}{2c}n_{N}-x(N,-)\leq-\frac{c-2}{2c}n_{N}\leq-\theta n_{N}

for 1iF1\leq i\leq F.

We want k=1CcN,kαc(N,k)\displaystyle\sum_{k=1}^{C}c_{N,k}\alpha^{c(N,k)} to be:

iI+(σ(s)s)xN,ixN,iαx(N,i)x(N,i)i=1Fσ(fN,i)xN,i(±1)f(N,i)αf(N,i)x(N,).-\sum_{i\in I^{+}}\frac{(\sigma(s)-s)x_{N,i}}{x_{N,i^{-}}}\alpha^{x(N,i)-x(N,i^{-})}-\sum_{i=1}^{F}\frac{\sigma(f_{N,i})}{x_{N,i^{-}}}(\pm 1)^{f(N,i)}\alpha^{-f(N,i)-x(N,-)}.

We let C=|I+|+F=2P2+FC=|I^{+}|+F=2^{P-2}+F and specify the cN,kc_{N,k} and c(N,k)c(N,k) in the same manner as before. Note that x(N,+)x(N,+) is the minimum among the x(N,i)x(N,i) for iI+i\in I^{+} while Lemma 5.2 and Proposition 5.7 yields

f(N,i)x~(N,)c22cnN=x~(N,+)c22cnN>x(N,+).-f(N,i)\geq-\tilde{x}(N,-)-\frac{c-2}{2c}n_{N}=\tilde{x}(N,+)-\frac{c-2}{2c}n_{N}>x(N,+).

This guarantees c(N,k)x(N,+)x(N,)2(c2)5(c1)nNθnNc(N,k)\geq x(N,+)-x(N,-)\geq\displaystyle\frac{2(c-2)}{5(c-1)}n_{N}\geq\theta n_{N}.

Finally, we want =1V(ssNP)vN,αv(N,)\displaystyle\sum_{\ell=1}^{V}(s-s_{N-P})v_{N,\ell}\alpha^{v(N,\ell)} to be

i=1D(ssNP)dN,ixN,iαd(N,i)x(N,)\displaystyle-\sum_{i=1}^{D}(s-s_{N-P})\frac{d_{N,i}}{x_{N,i^{-}}}\alpha^{d(N,i)-x(N,-)}

and want m=1W(σ(s)sNP)wN,mαw(N,m)\displaystyle\sum_{m=1}^{W}(\sigma(s)-s_{N-P})w_{N,m}\alpha^{w(N,m)} to be

i=1D(σ(s)sNP)σ(dN,i)xN,i(±1)d(N,i)αd(N,i)x(N,).\sum_{i=1}^{D}(\sigma(s)-s_{N-P})\frac{\sigma(d_{N,i})}{x_{N,i^{-}}}(\pm 1)^{d(N,i)}\alpha^{-d(N,i)-x(N,-)}.

We let V=W=DV=W=D and similar arguments can be used to finish the proof; note that with our choice:

A+B+C+V+W=2P11+2(D+F)2P+P2P11A+B+C+V+W=2^{P-1}-1+2(D+F)\leq 2^{P}+P2^{P-1}-1

where the last inequality follows from Proposition 5.3. ∎

Definition 5.10.

Among all the collections of data (A,B,C,V,W,𝒩,)(A,B,C,V,W,\mathscr{N}^{\prime},\ldots) satisfying properties (i)-(vii) in Proposition 5.9, we choose one for which A+B+C+V+WA+B+C+V+W is minimal. By abusing the notation, we still use the same notation AA, BB, CC, VV, WW, 𝒩\mathscr{N}^{\prime}, aN,ia_{N,i}’s, a(N,i)a(N,i)’s, bN,jb_{N,j}’s, b(N,j)b(N,j)’s, cN,kc_{N,k}’s, c(N,k)c(N,k)’s, vN,v_{N,\ell}’s, v(N,)v(N,\ell)’s, wN,mw_{N,m}, and w(N,m)w(N,m)’s for this chosen data with minimal A+B+C+V+WA+B+C+V+W.

Another application of the Subspace Theorem yields the following:

Proposition 5.11.

Recall that θ=min{c22c,2(c2)5(c1)}\theta=\displaystyle\min\left\{\frac{c-2}{2c},\frac{2(c-2)}{5(c-1)}\right\}. Assume that PP satisfies:

(36) 2P+P2P11cP1<θ2.\frac{2^{P}+P2^{P-1}-1}{c^{P-1}}<\frac{\theta}{2}.

Then we have B=C=V=W=0B=C=V=W=0.

Proof.

First, suppose that B>0B>0. Let ϵ>0\epsilon>0 be a small number that will be specified later. We apply the Subspace Theorem over the field (α){\mathbb{Q}}(\alpha) and let S={v,w}S=\{v_{\infty},w\} be as before. We work with linear forms in the variables:

((Xi)1iA,(Yj)1jB,(Zk)1kC,(R)1V,(Tm)1mW)((X_{i})_{1\leq i\leq A},(Y_{j})_{1\leq j\leq B},(Z_{k})_{1\leq k\leq C},(R_{\ell})_{1\leq\ell\leq V},(T_{m})_{1\leq m\leq W})

and the vectors

𝐯N=(\displaystyle{\mathbf{v}}_{N}=( (aN,iαa(N,i))1iA,(bN,jαb(N,j))1jB,((cN,kαc(N,k))1kC,\displaystyle(a_{N,i}\alpha^{a(N,i)})_{1\leq i\leq A},(b_{N,j}\alpha^{b(N,j)})_{1\leq j\leq B},((c_{N,k}\alpha^{c(N,k)})_{1\leq k\leq C},
((ssNP)vN,αv(N,))1V,((σ(s)sNP)wN,mαw(N,m))1mW)\displaystyle((s-s_{N-P})v_{N,\ell}\alpha^{v(N,\ell)})_{1\leq\ell\leq V},((\sigma(s)-s_{N-P})w_{N,m}\alpha^{w(N,m)})_{1\leq m\leq W})

for N𝒩N\in\mathscr{N}^{\prime}. For vSv\in S, the linear forms are denoted Lv,X,iL_{v,X,i}, Lv,Y,jL_{v,Y,j}, Lv,Z,kL_{v,Z,k}, Lv,R,L_{v,R,\ell}, and Lv,T,mL_{v,T,m} for 1iA1\leq i\leq A, 1jB1\leq j\leq B, 1kC1\leq k\leq C, 1V1\leq\ell\leq V and 1mW1\leq m\leq W and they are defined as follows:

  • For any vSv\in S, Lv,X,i=XiL_{v,X,i}=X_{i}, Lv,Y,j=YjL_{v,Y,j}=Y_{j}, Lv,Z,k=ZkL_{v,Z,k}=Z_{k}, Lv,R,=RL_{v,R,\ell}=R_{\ell}, and Lv,T,m=TmL_{v,T,m}=T_{m} for every i,j,k,,mi,j,k,\ell,m except when j=1j=1.

  • If v=vv=v_{\infty}, define Lv,Y,1=Y1L_{v,Y,1}=Y_{1}.

  • If v=wv=w, define

    Lv,Y,1=i=1AXi+j=1BYj+k=1CZk+=1VR+m=1WTm.L_{v,Y,1}=\sum_{i=1}^{A}X_{i}+\sum_{j=1}^{B}Y_{j}+\sum_{k=1}^{C}Z_{k}+\sum_{\ell=1}^{V}R_{\ell}+\sum_{m=1}^{W}T_{m}.

Therefore if v=vv=v_{\infty}, we have |Lv,Y,1(𝐯N)|v=|bN,1|1/2|α|b(N,1)/2|α|(θnN/2)+o(nN)|L_{v,Y,1}({\mathbf{v}}_{N})|_{v}=|b_{N,1}|^{1/2}|\alpha|^{b(N,1)/2}\leq|\alpha|^{(-\theta n_{N}/2)+o(n_{N})} since bN,1b_{N,1} has height o(nN)o(n_{N}) while b(N,1)θnNb(N,1)\leq-\theta n_{N}. If v=wv=w, we have Lv,Y,1(𝐯N)=σ(s)sL_{v,Y,1}({\mathbf{v}}_{N})=\sigma(s)-s thanks to (34). Thus, arguing as before, we have:

(37) vSL|L(𝐯N)|v<|α|(θ+ϵ)nN/2\displaystyle\prod_{v\in S}\prod_{L}|L({\mathbf{v}}_{N})|_{v}<|\alpha|^{(-\theta+\epsilon)n_{N}/2}

for all sufficiently large N𝒩N\in\mathscr{N}^{\prime} where LL ranges over all the Lv,X,iL_{v,X,i}’s, Lv,Y,jL_{v,Y,j}’s, Lv,Z,kL_{v,Z,k}’s, Lv,R,L_{v,R,\ell}’s, and Lv,T,mL_{v,T,m}. On the other hand,

(38) H(𝐯~N)A+B+C+V+WvSL1|𝐯N|v=(vMKS|𝐯N|v)A+B+C+V+WH(sNP)A+B+C+V+WαϵnN|α|(A+B+C+V+W)nNP+1+ϵnN.\displaystyle\begin{split}H(\tilde{{\mathbf{v}}}_{N})^{A+B+C+V+W}\prod_{v\in S}\prod_{L}\frac{1}{|{\mathbf{v}}_{N}|_{v}}&=\left(\prod_{v\in M_{K}\setminus S}|{\mathbf{v}}_{N}|_{v}\right)^{A+B+C+V+W}\\ &\leq H(s_{N-P})^{A+B+C+V+W}\alpha^{\epsilon n_{N}}\\ &\leq|\alpha|^{(A+B+C+V+W)n_{N-P+1}+\epsilon n_{N}}.\end{split}

Since A+B+C+V+W2P+P2P11A+B+C+V+W\leq 2^{P}+P2^{P-1}-1, we can apply the Subspace Theorem if:

2P+P2P11cP1+ϵ+θ+ϵ2<0.\frac{2^{P}+P2^{P-1}-1}{c^{P-1}}+\epsilon+\frac{-\theta+\epsilon}{2}<0.

At the beginning of the proof, can choose an ϵ\epsilon satisfying the above inequality thanks to the condition on PP. Then the Subspace Theorem implies that the coordinates of 𝐯N{\mathbf{v}}_{N} satisfies a non-trivial linear relation over (α){\mathbb{Q}}(\alpha) for every NN in an infinite subset 𝒩′′\mathscr{N}^{\prime\prime} of 𝒩\mathscr{N}^{\prime}. Then we have a new data satisfying the properties in Proposition 5.9 in which 𝒩\mathscr{N}^{\prime} is replaced by 𝒩′′\mathscr{N}^{\prime\prime} and A+B+C+V+WA+B+C+V+W is replaced by A+B+C+V+W1A+B+C+V+W-1; this contradicts the minimality of A+B+C+V+WA+B+C+V+W.

For the case C>0C>0, V>0V>0, or W>0W>0, we use the same vectors 𝐯N{\mathbf{v}}_{N} and the same notation for the variables and linear forms. In the case C>0C>0, the linear forms are:

  • For any vSv\in S, Lv,X,i=XiL_{v,X,i}=X_{i}, Lv,Y,j=YjL_{v,Y,j}=Y_{j}, Lv,Z,k=ZkL_{v,Z,k}=Z_{k}, Lv,R,=RL_{v,R,\ell}=R_{\ell}, and Lv,T,m=TmL_{v,T,m}=T_{m} for every i,j,k,,mi,j,k,\ell,m except when k=1k=1.

  • If v=wv=w, define Lv,Z,1=Z1L_{v,Z,1}=Z_{1}.

  • If v=vv=v_{\infty}, define

    Lv,Z,1=i=1AXi+j=1BYj+k=1CZk+=1VR+m=1WTm.L_{v,Z,1}=\sum_{i=1}^{A}X_{i}+\sum_{j=1}^{B}Y_{j}+\sum_{k=1}^{C}Z_{k}+\sum_{\ell=1}^{V}R_{\ell}+\sum_{m=1}^{W}T_{m}.

In the case V>0V>0, the linear forms are:

  • For any vSv\in S, Lv,X,i=XiL_{v,X,i}=X_{i}, Lv,Y,j=YjL_{v,Y,j}=Y_{j}, Lv,Z,k=ZkL_{v,Z,k}=Z_{k}, Lv,R,=RL_{v,R,\ell}=R_{\ell}, and Lv,T,m=TmL_{v,T,m}=T_{m} for every i,j,k,,mi,j,k,\ell,m except when =1\ell=1.

  • If v=vv=v_{\infty}, define Lv,R,1=R1L_{v,R,1}=R_{1}.

  • If v=wv=w, define

    Lv,R,1=i=1AXi+j=1BYj+k=1CZk+=1VR+m=1WTm.L_{v,R,1}=\sum_{i=1}^{A}X_{i}+\sum_{j=1}^{B}Y_{j}+\sum_{k=1}^{C}Z_{k}+\sum_{\ell=1}^{V}R_{\ell}+\sum_{m=1}^{W}T_{m}.

Finally, in the case W>0W>0, the linear forms are:

  • For any vSv\in S, Lv,X,i=XiL_{v,X,i}=X_{i}, Lv,Y,j=YjL_{v,Y,j}=Y_{j}, Lv,Z,k=ZkL_{v,Z,k}=Z_{k}, Lv,R,=RL_{v,R,\ell}=R_{\ell}, and Lv,T,m=TmL_{v,T,m}=T_{m} for every i,j,k,,mi,j,k,\ell,m except when m=1m=1.

  • If v=wv=w, define Lv,T,1=T1L_{v,T,1}=T_{1}.

  • If v=vv=v_{\infty}, define

    Lv,T,1=i=1AXi+j=1BYj+k=1CZk+=1VR+m=1WTm.L_{v,T,1}=\sum_{i=1}^{A}X_{i}+\sum_{j=1}^{B}Y_{j}+\sum_{k=1}^{C}Z_{k}+\sum_{\ell=1}^{V}R_{\ell}+\sum_{m=1}^{W}T_{m}.

Then similar arguments as before lead to a contradiction. This finishes the proof. ∎

Completion of the proof of Theorem 1.3.

At the beginning of this section, we fix a sufficiently large PP satisfying both (27) and (36). Then the previous results show that there exist an infinite set of positive integers 𝒩\mathscr{N}^{\prime}, an integer A0A\geq 0, tuples (aN,1,,aN,A)(a_{N,1},\ldots,a_{N,A}) and (a(N,1),,a(N,A))(a(N,1),\ldots,a(N,A)) satisfying the conditions of Proposition 5.9; in particular:

σ(s)s=aN,1αa(N,1)++aN,Aαa(N,A)\sigma(s)-s=a_{N,1}\alpha^{a(N,1)}+\ldots+a_{N,A}\alpha^{a(N,A)}

for every N𝒩N\in\mathscr{N}^{\prime}. However, each |aN,i|=|α|o(nN)|a_{N,i}|=|\alpha|^{o(n_{N})} as NN\to\infty while each a(N,i)<2(c2)5P1(c1)nNa(N,i)<\displaystyle-\frac{2(c-2)}{5^{P-1}(c-1)}n_{N}. Let NN\to\infty the we have

σ(s)s=0\sigma(s)-s=0

contradicting the earlier results that s(α)s\in{\mathbb{Q}}(\alpha) is irrational. This finishes the proof. ∎

References

  • [AB07] B. Adamczewski and Y. Bugeaud, On the complexity of algebraic numbers I. Expansions in integer bases, Ann. of Math. (2) 165 (2007), 547–565.
  • [Bad93] C. Badea, A theorem on irrationality of infinite series and applications, Acta Arith. 63 (1993), 313–323.
  • [BG06] E. Bombieri and W. Gubler, Heights in Diophantine geometry, New Mathematical Monographs, vol. 4, Cambridge University Press, Cambridge, 2006.
  • [BMS06] Y. Bugeaud, M. Mignotte, and S. Siksek, Classical and modular approaches to exponential Diophantine equations I. Fibonacci and Lucas perfect powers., Ann. of Math. (2) 163 (2006), 969–1018.
  • [BT94] P.-G. Becker and T. Töpfer, Transcendency results for sums of reciprocals of linear recurrences, Math. Nachr. 168 (1994), 5–17.
  • [CZ02] P. Corvaja and U. Zannier, Some new applications of the Subspace Theorem, Compos. Math. 131 (2002), 319–340.
  • [CZ04] by same author, On the rational approximations to the powers of an algebraic number: Solution of two problems of Mahler and Mendès France, Acta Math. 193 (2004), 175–191.
  • [DKT02] D. Duverney, T. Kanoko, and T. Tanaka, Transcendence of certain reciprocal sums of linear recurrences, Monatsh. Math. 137 (2002), 115–128.
  • [EG80] P. Erdös and G. L. Graham, Old and new problems and results in combinatorial number theory, L’Enseignement mathémathique, vol. 28, Universite Genève, 1980.
  • [ES02] J.-H. Evertse and H. P. Schlickewei, A quantitative version of the Absolute Subspace Theorem, J. Reine Angew Math. 548 (2002), 21–127.
  • [Eve96] J.-H. Evertse, An improvement of the quantitative subspace theorem, Compos. Math. 101 (1996), 225–311.
  • [KKS09] T. Kanoko, T. Kurosawa, and I. Shiokawa, Transcendence of reciprocal sums of binary recurrences, Monatsh. Math. 157 (2009), 323–334.
  • [KMN19] A. Kulkarni, N. M. Mavraki, and K. D. Nguyen, Algebraic approximations to linear combinations of powers: an extension of results by Mahler and Corvaja-Zannier, Trans. Amer. Math. Soc. 371 (2019), 3787–3804.
  • [Mah75] K. Mahler, On the transcendency of the solutions of a special class of functional equations, Bull. Aust. Math. Soc. 13 (1975), 389–410.
  • [Mig71] M. Mignotte, Quelques problèmes d’effectivité en théorie des nombres, Ph.D. thesis, L’Université de Paris XIII, 1971.
  • [Mig77] by same author, An Application of W. M. Schmidt’s Theorem. Transcendental Numbers and Golden Number, Fibonacci Quart. 15 (1977), 15–16.
  • [Nis96] K. Nishioka, Mahler functions and transcendence, Lecture Notes in Math., vol. 1631, Springer-Verlag, 1996.
  • [Sch70] W. M. Schmidt, Simultaneous approximation to algebraic numbers by rationals, Acta Math. 125 (1970), 189–201.
  • [Sch80] K. Schmidt, On periodic expansions of Pisot numbers and Salem numbers, Bull. Lond. Math. Soc. 12 (1980), 269–278.
  • [Sch92] H. P. Schlickewei, The quantitative subspace theorem for number fields, Compos. Math. 82 (1992), 245–273.
  • [Ste13] C. L. Stewart, On divisors of Lucas and Lehmer numbers, Acta Math. 211 (2013), 291–314.
  • [TZ99] G. Troi and U. Zannier, Note on the density constant in the distribution of self-numbers. II, Boll. Unione Mat. Ital. (8) 2-B (1999), 397–399.