这是indexloc提供的服务,不要输入任何密码

Hermite’s approach to Abelian integrals revisited

Makoto Kawashima
Abstract

In this article, we establish a new linear independence criterion for the values of certain Lauricella hypergeometric series FDF_{D} with rational parameters, in both the complex and pp-adic settings, over an algebraic number field. This result generalizes a theorem of C. Hermite [16] on the linear independence of certain Abelian integrals. Our proof relies on explicit Padé type approximations to solutions of a reducible Jordan-Pochhammer differential equation, which extends the Padé approximations for certain Abelian integrals in [16]. The main novelty of our approach lies in the proof of the non-vanishing of the determinants associated with these Padé type approximants.

Key words and phrases: Lauricella hypergeometric series, Jordan-Pochhammer differential equation, Padé approximation, Rodrigues formula, GG-functions, linear independence.

1 Introduction

In this paper we extend Hermite’s construction of Padé approximants for Abelian integrals to a broad class of Lauricella hypergeometric series, and derive an explicit criterion (with an effective measure) for the linear independence of their values over algebraic number fields - simultaneously in the complex and pp-adic contexts.

Padé approximation, originating in the works of C. Hermite in his study of the transcendence of ee [15] and H. Padé [24, 25], has long played a central role in Diophantine approximation and transcendental number theory. In arithmetic applications, one typically constructs explicit systems of Padé approximations to certain functions, often by linear algebraic arguments combined with bounds derived from Siegel’s lemma via Dirichlet’s box principle. However, these general constructions are not always sufficient for arithmetic purposes, such as establishing linear independence results. In such cases, it becomes essential to construct explicit Padé approximations providing analytic estimates sharp enough for the intended application–a task that can usually be carried out only for special classes of functions.

Over the complex field, Hermite [17] established a criterion for the \mathbb{Q}-linear independence of the values of certain Abelian integrals associated with the Jordan-Pochhammer differential equation, by explicitly constructing Padé approximations of these integrals. In his work, all parameters sis_{i} determining the exponents of the Jordan-Pochhammer equation (see equation (2)) are assumed to be of the form 1/k1/k for positive integers kk. Following Hermite’s approach, G. Rhin and P. Toffin [28] proved a linear independence criterion for distinct logarithmic values over imaginary quadratic fields by constructing explicit Padé-type approximants of logarithmic functions, corresponding to the case si=0s_{i}=0. Later, M. Huttner [17] refined Hermite’s method and, through a detailed analysis of Hermite’s approximants, obtained sharper measures of linear independence than those of Hermite’s original results (see also [18]). In [20], the author established a linear independence criterion for the values of Gauss’ hypergeometric functions with varying parameters, both in the complex and in the pp-adic settings, for a special case where all si=1/ms_{i}=-1/m with mm the number of parameters. Related Diophantine properties of the values of solutions to Jordan-Pochhammer equations were investigated by G. V. and D. V. Chudnovsky in [8, Section 2].

In the present work, we extend Hermite’s criterion to the case where the parameters sis_{i} are arbitrary rational numbers whose sum does not lie in <1\mathbb{Z}_{<-1}. This leads to a new linear independence criterion for the corresponding Abelian integrals over algebraic number fields, valid simultaneously in both the complex and pp-adic settings. Our approach relies on a detailed study of Padé-type approximants (beyond the classical Padé approximants) and their algebraic and analytic structures, following the framework developed in [20].

To investigate the arithmetic nature of the values of holonomic Laurent series, it is crucial to construct their Padé-type approximants and to elucidate their analytic and algebraic properties. In [20], explicit constructions were given for holonomic series on which a first-order differential operator with polynomial coefficients acts so that the resulting function becomes a polynomial. In this paper, we establish more general sufficient conditions ensuring the non-vanishing of determinants associated with such Padé-type approximants. This analysis is motivated by the goal of extending the method to broader classes of holonomic Laurent series and by the desire to understand the arithmetic implications of non-vanishing–a property that plays a central role in Siegel’s method [29].

A key technical ingredient is the introduction of the formal ff-integration map φf\varphi_{f} (see equation (5)), which enables the explicit–yet purely formal–construction of Padé-type approximants and the proof of their principal properties. This concept appears in various forms in earlier works of S. David, N. Hirata-Kohno, and the author [9, 10, 11, 12], as well as in [21, 22] by A. Poëls and the author.

To apply Siegel’s method [29] in proving our main theorem, it is essential to establish the non-vanishing of the determinant formed by these Padé-type approximants. In previous works such as [9, 10, 11, 21, 20], this step required explicit computation of the determinant, which is often a delicate and technically demanding task. Here, we develop a new approach based on the study of the kernel of the formal ff-integration map, as introduced in [22] and [12]. This method allows us to prove the non-vanishing property in a simpler and more conceptual way, without computing the explicit determinant values. We emphasize that the non-vanishing condition is governed entirely by the first-order differential operators involved–specifically, by the coefficients of operator (see Theorem 3.1).

1.1 Main result

Our main result concerns a linear independence criterion for the values of the Lauricella hypergeometric series over an algebraic number field, in both the complex and pp-adic settings. We begin by recalling the definition of the mm-variable Lauricella hypergeometric series with parameters α,β1,,βm,γ\alpha,\beta_{1},\ldots,\beta_{m},\gamma, defined by

FD(m)(α,β1,,βm,γ;z1,,zm)=k1,,km=0(α)k1++km(β1)k1(βm)km(γ)k1++kmk1!km!z1k1zmkm,F^{(m)}_{D}\!\left(\alpha,\beta_{1},\ldots,\beta_{m},\gamma;z_{1},\ldots,z_{m}\right)=\sum_{k_{1},\ldots,k_{m}=0}^{\infty}\dfrac{(\alpha)_{k_{1}+\cdots+k_{m}}(\beta_{1})_{k_{1}}\cdots(\beta_{m})_{k_{m}}}{(\gamma)_{k_{1}+\cdots+k_{m}}\,k_{1}!\cdots k_{m}!}\,z_{1}^{k_{1}}\cdots z_{m}^{k_{m}},

where (a)k(a)_{k} denotes the Pochhammer symbol, given by (a)0=1(a)_{0}=1 and (a)k=a(a+1)(a+k1)(a)_{k}=a(a+1)\cdots(a+k-1) for k1k\geq 1.

To state our main result, we first fix notation. Let KK be an algebraic number field and denote 𝔐K\mathfrak{M}_{K} the set of places of KK, and for v𝔐Kv\in\mathfrak{M}_{K}, let KvK_{v} denote the completion of KK at vv. We normalize the absolute value ||v|\cdot|_{v} by

|p|v=p[Kv:p][K:]if vp,|x|v=|ιv(x)|[Kv:][K:]if v,|p|_{v}=p^{-\tfrac{[K_{v}:\mathbb{Q}_{p}]}{[K:\mathbb{Q}]}}\quad\text{if }v\mid p,\qquad|x|_{v}=|\iota_{v}(x)|^{\tfrac{[K_{v}:\mathbb{R}]}{[K:\mathbb{Q}]}}\quad\text{if }v\mid\infty,

where pp is a rational prime and ιv:K\iota_{v}:K\hookrightarrow\mathbb{C} the embedding associated to vv. On KvnK_{v}^{n}, the norm ||v|\cdot|_{v} is taken to be the supremum norm.

For 𝜷=(β1,,βm)Km\boldsymbol{\beta}=(\beta_{1},\ldots,\beta_{m})\in K^{m}, we define the logarithmic vv-adic height and the global logarithmic height by

hv(𝜷)\displaystyle{\mathrm{h}}_{v}(\boldsymbol{\beta}) =logmax{1,|β1|v,,|βm|v},\displaystyle=\log\max\{1,|\beta_{1}|_{v},\ldots,|\beta_{m}|_{v}\},
h(𝜷)\displaystyle{\mathrm{h}}(\boldsymbol{\beta}) =v𝔐Khv(𝜷).\displaystyle=\sum_{v\in\mathfrak{M}_{K}}{\mathrm{h}}_{v}(\boldsymbol{\beta}).

We also define the denominator of 𝜷\boldsymbol{\beta} as

den(𝜷)=min{nn>0such thatnβiare algebraic integers for all 1im}.{\rm{den}}(\boldsymbol{\beta})=\min\{n\in\mathbb{Z}\mid n>0\ \text{such that}\ n\beta_{i}\ \text{are algebraic integers for all }1\leq i\leq m\}.

For a rational number α\alpha, we put

μ(α)=den(α)q:primeqden(α)q1q1.\mu(\alpha)={\rm{den}}(\alpha)\prod_{\begin{subarray}{c}q:\,\text{prime}\\ q\mid{\rm{den}}(\alpha)\end{subarray}}q^{\tfrac{1}{q-1}}.

Now we are ready to state our main result. Let m2m\geq 2 be a fixed positive integer. Fix an algebraic number field KK. Consider polynomials a(z),b(z)K[z]a(z),b(z)\in K[z], where a(z)a(z) is assumed to be monic of degree mm and b(z)b(z) satisfies degbm1\deg b\leq m-1. We denote the derivative of a(z)a(z) by a(z)a^{\prime}(z), and by bm1b_{m-1} the coefficient of zm1z^{m-1} in b(z)b(z). Note that bm1b_{m-1} can be 0. Assume that a(z)a(z) is decomposable over KK, and let α1,,αmK\alpha_{1},\ldots,\alpha_{m}\in K denote its roots, counted with multiplicity. We impose the following conditions:

(1) α1,,αmare pairwise distinct,\displaystyle\alpha_{1},\ldots,\alpha_{m}\ \text{are pairwise distinct},
(2) si:=b(αi)a(αi)1for all 1im,\displaystyle s_{i}:=\dfrac{b(\alpha_{i})}{a^{\prime}(\alpha_{i})}\in\mathbb{Q}\setminus\mathbb{Z}_{\leq-1}\quad\text{for all }1\leq i\leq m,
(3) bm1<1.\displaystyle b_{m-1}\notin\mathbb{Z}_{<-1}.

For v𝔐Kv\in\mathfrak{M}_{K} and βK{0}\beta\in K\setminus\{0\}, we introduce the quantity

(4) Vv(β)=\displaystyle V_{v}(\beta)= mhv(β)(m1)h(β)i=1mh(αi)m(h(𝜶)+i=1mlogμ(si)+log4)\displaystyle\,m\,{\rm{h}}_{v}(\beta)-(m-1){\rm{h}}(\beta)-\sum_{i=1}^{m}{\rm{h}}(\alpha_{i})-m\left({\rm{h}}(\boldsymbol{\alpha})+\sum_{i=1}^{m}\log\mu(s_{i})+\log 4\right)
(m1)den(bm1)φ(den(bm1))1jden(bm1)(j,den(bm1))=11j,\displaystyle-(m-1)\dfrac{{\rm{den}}(b_{m-1})}{\varphi({\rm{den}}(b_{m-1}))}\!\sum_{\begin{subarray}{c}1\leq j\leq{\rm{den}}(b_{m-1})\\ (j,{\rm{den}}(b_{m-1}))=1\end{subarray}}\dfrac{1}{j},

where 𝜶=(α1,,αm)Km\boldsymbol{\alpha}=(\alpha_{1},\ldots,\alpha_{m})\in K^{m} and φ\varphi denotes Euler’s totient function.

Theorem 1.1.

Retain the above notation and assumptions (1), (2) and (3). For 0jm20\leq j\leq m-2, we define formal Laurent series fj(z)f_{j}(z) by

i=1m(1αiz)si1zj+1FD(m)(bm1+j+1, 1+s1,,1+sm;bm1+j+2;α1z,,αmz).\prod_{i=1}^{m}\left(1-\frac{\alpha_{i}}{z}\right)^{s_{i}}\cdot\frac{1}{z^{j+1}}\cdot F^{(m)}_{D}\!\left(b_{m-1}+j+1,\,1+s_{1},\ldots,1+s_{m};\,b_{m-1}+j+2;\,\frac{\alpha_{1}}{z},\ldots,\frac{\alpha_{m}}{z}\right).

Let v0𝔐Kv_{0}\in\mathfrak{M}_{K}. We assume Vv0(β)>0V_{v_{0}}(\beta)>0***The positivity condition Vv0(β)>0V_{v_{0}}(\beta)>0 roughly means that β\beta is arithmetically large enough at v0v_{0} compared with the heights of the parameters. This guarantees both the convergence of the involved Laurent series at z=βz=\beta and the arithmetic control required in the application of Siegel’s method. . Then each series fj(z)f_{j}(z) converges at z=βz=\beta in Kv0K_{v_{0}} and the mm elements of Kv0K_{v_{0}}::

1,f0(β),f1(β),,fm2(β),1,f_{0}(\beta),f_{1}(\beta),\ldots,f_{m-2}(\beta),

are linearly independent over KK.

We will give a proof of Theorem 1.1 together with a linear independence measure (see Theorem 5.1).

Remark 1.2.

We denote LL by the differential operator of order 11 with polynomial coefficients a(z)ddz+b(z)-a(z)\tfrac{d}{dz}+b(z). A result due to S. Fischler and T. Rivoal [14, Proposition 33 (ii)] establishes that LL is a GG-operator (see the definition of GG-operator [1, IV]) under the assumptions (1) and (2). We observe that the Laurent series fjf_{j} satisfies LfjK[z]L\cdot f_{j}\in K[z] with degree mj2m-j-2 (see Lemma 4.1). In particular, fjf_{j} is a solution of a reducible Jordan-Pochhammer equation (ddz)m1L\left(\tfrac{d}{dz}\right)^{m-1}L (see Example 6.1). Combining these results, by a well-known theorem of Y. André, G. V. & D. V. Chudnovsky and N. Katz (refer [2, Théorème 3.53.5]), we conclude fjf_{j} is a GG-function in the sense of C. F. Siegel [29].

Outline of this article. Section 2 is based on the results given in [20]. We begin by introducing the Padé-type approximants of Laurent series. In Subsection 2.1, we introduce the formal ff-integration map associated with a Laurent series ff, which plays a central role throughout this paper, and we describe its fundamental properties in case of ff is holonomic. Subsection 2.3 provides an overview of Padé-type approximants and Padé-type approximation for Laurent series that become polynomials under the action of a first-order differential operator with polynomial coefficients. Section 3 formulates, in terms of the coefficients of the differential operator, sufficient conditions ensuring the non-vanishing of the determinant formed by the Padé-type approximants constructed in Section 2 (see Theorem 3.1). This part constitutes the main novel contribution of the present work. In Section 4, we treat the case where the differential operator considered in Section 3 is a GG-operator. We give explicit expressions for the corresponding Laurent series and establish estimates for their Padé-type approximants and Padé approximations with respect to both Archimedean and non-Archimedean valuations. Section 5 is devoted to the proof of our main theorem, together with a quantitative measure of linear independence. Finally, Section 6 serves as an appendix, summarizing some basic facts on the Jordan-Pochhammer equation.

2 Padé-type approximants of Laurent series

Throughout this section, we fix a field KK of characteristic 0. We denote the formal power series ring of variable 1/z1/z with coefficients KK by K[[1/z]]K[[1/z]] and the field of fractions by K((1/z))K((1/z)). We say an element of K((1/z))K((1/z)) is a formal Laurent series. We define the order function at z=z=\infty by

ord:K((1/z)){};kakzkmin{k{}ak0}.{\rm{ord}}_{\infty}:K((1/z))\longrightarrow\mathbb{Z}\cup\{\infty\};\sum_{k}\dfrac{a_{k}}{z^{k}}\mapsto\min\{k\in\mathbb{Z}\cup\{\infty\}\mid a_{k}\neq 0\}.

Note that, for fK((1/z))f\in K((1/z)), ordf={\rm{ord}}_{\infty}\,f=\infty if and only if f=0f=0. We recall without proof the following elementary fact :

Lemma 2.1.

Let mm be a nonnegative integer, f1(z),,fm(z)(1/z)K[[1/z]]f_{1}(z),\ldots,f_{m}(z)\in(1/z)\cdot K[[1/z]] and 𝐧=(n1,,nm)m\boldsymbol{n}=(n_{1},\ldots,n_{m})\in\mathbb{N}^{m}. Put N=j=1mnjN=\sum_{j=1}^{m}n_{j}. For a nonnegative integer MM with MNM\geq N, there exist polynomials (P,Q1,,Qm)K[z]m+1{𝟎}(P,Q_{1},\ldots,Q_{m})\in K[z]^{m+1}\setminus\{\boldsymbol{0}\} satisfying the following conditions:

(i)(i) degPM{\rm{deg}}P\leq M,

(ii)(ii) ord(P(z)fj(z)Qj(z))nj+1for 1jm{\rm{ord}}_{\infty}\left(P(z)f_{j}(z)-Q_{j}(z)\right)\geq n_{j}+1\ \ \text{for}\ \ 1\leq j\leq m.

Definition 2.2.

We say that a vector of polynomials (P,Q1,,Qm)K[z]m+1(P,Q_{1},\ldots,Q_{m})\in K[z]^{m+1} satisfying properties (i)(i) and (ii)(ii) is a weight 𝒏\boldsymbol{n} and degree MM Padé-type approximant of (f1,,fm)(f_{1},\ldots,f_{m}). For such approximants (P,Q1,,Qm)(P,Q_{1},\ldots,Q_{m}) of (f1,,fm)(f_{1},\ldots,f_{m}), we call the formal Laurent series (P(z)fj(z)Qj(z))1jm(P(z)f_{j}(z)-Q_{j}(z))_{1\leq j\leq m}, that is to say remainders, as weight 𝒏\boldsymbol{n} degree MM Padé-type approximations of (f1,,fm)(f_{1},\ldots,f_{m}).

2.1 Formal ff-integration map

Let f(z)=k=0fk/zk+1(1/z)K[[1/z]]f(z)=\sum_{k=0}^{\infty}f_{k}/z^{k+1}\in(1/z)\cdot K[[1/z]]. We define a KK-linear map φfHomK(K[t],K)\varphi_{f}\in{\rm{Hom}}_{K}(K[t],K) by

(5) φf:K[t]K;tkfk(k0).\displaystyle\varphi_{f}:K[t]\longrightarrow K;\ \ \ t^{k}\mapsto f_{k}\ \ \ (k\geq 0).

The above linear map extends naturally a K[z]K[z]-linear map φf:K[z,t]K[z]\varphi_{f}:K[z,t]\rightarrow K[z], and then to a K[z][[1/z]]K[z][[1/z]]-linear map φf:K[z,t][[1/z]]K[z][[1/z]]\varphi_{f}:K[z,t][[1/z]]\rightarrow K[z][[1/z]]. With this notation, the formal Laurent series f(z)f(z) satisfies the following crucial identities (see [23, (6.2)(6.2) page 60 and (5.7)(5.7) page 52]):

f(z)=φf(1zt),P(z)f(z)φf(P(z)P(t)zt)(1/z)K[[1/z]]for anyP(z)K[z].\displaystyle f(z)=\varphi_{f}\left(\dfrac{1}{z-t}\right),\ \ \ P(z)f(z)-\varphi_{f}\left(\dfrac{P(z)-P(t)}{z-t}\right)\in(1/z)\cdot K[[1/z]]\ \ \text{for any}\ \ P(z)\in K[z].

Let us recall a condition, based on the morphism φf\varphi_{f}, for given polynomials to be Padé approximants.

Lemma 2.3 (confer [20, Lemma 2.3]).

Let m,Mm,M be positive integers, f1(z),,fm(z)(1/z)K[[1/z]]f_{1}(z),\ldots,f_{m}(z)\in(1/z)\cdot K[[1/z]] and 𝐧=(n1,,nm)m\boldsymbol{n}=(n_{1},\ldots,n_{m})\in\mathbb{N}^{m} with j=1mnjM\sum_{j=1}^{m}n_{j}\leq M. Let P(z)K[z]P(z)\in K[z] be a non-zero polynomial with degPM\deg P\leq M, and put Qj(z)=φfj((P(z)P(t))/(zt))K[z]Q_{j}(z)=\varphi_{f_{j}}\left((P(z)-P(t))/(z-t)\right)\in K[z] for 1jm1\leq j\leq m. The following assertions are equivalent.

(i)(i) The vector of polynomials (P,Q1,,Qm)(P,Q_{1},\ldots,Q_{m}) is a weight 𝐧\boldsymbol{n} Padé-type approximants of (f1,,fm)(f_{1},\ldots,f_{m}).

(ii)(ii) We have tkP(t)kerφfjt^{k}P(t)\in{\rm{ker}}\,\varphi_{f_{j}} for any pair of integers (j,k)(j,k) with 1jm1\leq j\leq m and 0knj10\leq k\leq n_{j}-1.

2.2 Holonomic series and kernel of ff-integration map

Lemma 2.3 implies that the study of the kernel of the formal ff-integration map is essential for constructing Padé approximants of Laurent series. This aspect has been explored in [20] for holonomic Laurent series. In this subsection, we recall a result from [20, Corollary 2.62.6].

We describe the action of a differential operator LL on a function ff, such as a polynomial or a Laurent series, by LfL\cdot f. Consider the map

(6) ι:K(z)[ddz]K(t)[ddt];jPj(z)djdzjj(1)jdjdtjPj(t).\displaystyle\iota:K(z)[\tfrac{d}{dz}]\longrightarrow K(t)[\tfrac{d}{dt}];\ \ \sum_{j}P_{j}(z)\tfrac{d^{j}}{dz^{j}}\mapsto\sum_{j}(-1)^{j}\tfrac{d^{j}}{dt^{j}}P_{j}(t).

Note, for LK(z)[ddz]L\in K(z)[\tfrac{d}{dz}], ι(L)\iota(L) is called the adjoint of LL and relates to the dual of differential module K(z)[ddz]/K(z)[ddz]LK(z)[\tfrac{d}{dz}]/K(z)[\tfrac{d}{dz}]L (see [1, III Exercises 3)3)]). For LK(z)[ddz]L\in K(z)[\tfrac{d}{dz}], we denote ι(L)\iota(L) by LL^{*}. Notice that we have (L1L2)=L2L1(L_{1}L_{2})^{*}=L^{*}_{2}L^{*}_{1} for any L1,L2K(z)[ddz]L_{1},L_{2}\in K(z)[\tfrac{d}{dz}].

Proposition 2.4.

(([20, Corollary 2.62.6])) Let f(z)(1/z)K[[1/z]]f(z)\in(1/z)\cdot K[[1/z]] and LK[z,ddz]L\in K[z,\tfrac{d}{dz}]. Then the following are equivalent:

(i)(i) Lf(z)K[z]L\cdot f(z)\in K[z].

(ii)(ii) LK[t]kerφfL^{*}\cdot K[t]\subseteq{\rm{ker}}\,\varphi_{f}.

2.3 Order 11 differential operator and Padé-type approximants

Let KK be a field of characteristic 0 and a(z),b(z)K[z]a(z),b(z)\in K[z]. Put dega=u,degb=v{\rm{deg}}\,a=u,{\rm{deg}}\,b=v, w=max{u2,v1}w=\max\{u-2,v-1\} and

(7) a(z)=i=0uaizi,b(z)=j=0vbjzj.\displaystyle a(z)=\sum_{i=0}^{u}a_{i}z^{i},\ \ b(z)=\sum_{j=0}^{v}b_{j}z^{j}.

We now assume

(8) w0,\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ w\geq 0,
(9) au(k+u)+bv0for allk0ifu1=v.\displaystyle a_{u}(k+u)+b_{v}\neq 0\ \ \ \text{for all}\ \ k\geq 0\ \ \text{if}\ \ u-1=v.

Define the order 11 differential operator with polynomial coefficients

L=a(z)ddz+b(z)K[z,ddz].L=-a(z)\dfrac{d}{dz}+b(z)\in K\left[z,\tfrac{d}{dz}\right].

We consider the Laurent series f(z)(1/z)K[[1/z]]f(z)\in(1/z)\cdot K[[1/z]] such that Lf(z)K[z]L\cdot f(z)\in K[z] and construct the Padé approximation of f(z)f(z). The following lemma is proven in [20, Lemma 4.14.1]. For the sake of completion, let us recall the proof in the present article.

Lemma 2.5.

We keep the notation above. Then there exist f0(z),,fw(z)(1/z)K[[1/z]]f_{0}(z),\ldots,f_{w}(z)\in(1/z)\cdot K[[1/z]] that are linearly independent over KK and satisfy Lfj(z)K[z]L\cdot f_{j}(z)\in K[z] with at most degree ww for 0jw0\leq j\leq w.

Proof..

Let f(z)=k=0fk/zk+1(1/z)K[[1/z]]f(z)=\sum_{k=0}^{\infty}f_{k}/z^{k+1}\in(1/z)\cdot K[[1/z]] be a Laurent series. There exists a polynomial A(z)K[z]A(z)\in K[z] that depends on the operator LL and ff with degAw{\rm{deg}}\,A\leq w and satisfying

(10) Lf(z)=A(z)+k=0i=0uai(k+i)fk+i1+j=0vbjfk+jzk+1.\displaystyle L\cdot f(z)=A(z)+\sum_{k=0}^{\infty}\dfrac{\sum_{i=0}^{u}a_{i}(k+i)f_{k+i-1}+\sum_{j=0}^{v}b_{j}f_{k+j}}{z^{k+1}}.

Put

i=0uai(k+i)fk+i1+j=0vbjfk+j=ck,0fk1++ck,wfk+w+ck,w+1fk+w+1fork0,\sum_{i=0}^{u}a_{i}(k+i)f_{k+i-1}+\sum_{j=0}^{v}b_{j}f_{k+j}=c_{k,0}f_{k-1}+\cdots+c_{k,w}f_{k+w}+c_{k,w+1}f_{k+w+1}\ \ \ \text{for}\ \ \ k\geq 0,

with c0,0=0c_{0,0}=0. We remark that ck,lc_{k,l} depends only on a(z),b(z)a(z),b(z). Notice that ck,w+1c_{k,w+1} is au(k+u)a_{u}(k+u) if u2>v1u-2>v-1, bvb_{v} if u2<v1u-2<v-1 and au(k+u)+bva_{u}(k+u)+b_{v} if u2=v1u-2=v-1. Then the assumption (9) ensures min{k0ck,w+10for allkk}=0\min\{k\geq 0\mid c_{k^{\prime},w+1}\neq 0\ \text{for all}\ k^{\prime}\geq k\}=0 and thus the KK-linear map:

(11) Kw+1{f(1/z)K[[1/z]]LfK[z]};(f0,,fw)k=0fkzk+1,\displaystyle K^{w+1}\longrightarrow\{f\in(1/z)\cdot K[[1/z]]\mid L\cdot f\in K[z]\};\ \ (f_{0},\ldots,f_{w})\mapsto\sum_{k=0}^{\infty}\dfrac{f_{k}}{z^{k+1}},

where, for kw+1k\geq w+1, fkf_{k} is determined inductively by

(12) i=0uai(k+i)fk+i1+j=0vbjfk+j=0fork0\displaystyle\sum_{i=0}^{u}a_{i}(k+i)f_{k+i-1}+\sum_{j=0}^{v}b_{j}f_{k+j}=0\ \ \text{for}\ \ k\geq 0

is an isomorphism. This completes the proof of Lemma 2.5. ∎

Let us fix Laurent series f0(z),,fw(z)(1/z)K[[1/z]]f_{0}(z),\ldots,f_{w}(z)\in(1/z)\cdot K[[1/z]] such that these series are linearly independent over KK and Lfj(z)K[z]L\cdot f_{j}(z)\in K[z] for 0jw0\leq j\leq w. We denote the formal integration associated to fjf_{j} by φfj=φj\varphi_{f_{j}}=\varphi_{j}. For a non-negative integer nn, we denote the nn-th Rodriges operator associated with LL (refer [4, Equation (2.5)(2.5)] and [20, Definition 3.13.1]) by

Rn=1n!(ddz+b(z)a(z))na(z)n.\displaystyle R_{n}=\dfrac{1}{n!}\left(\dfrac{d}{dz}+\dfrac{b(z)}{a(z)}\right)^{n}a(z)^{n}.

The differential operator RnR_{n} can be decomposed as follows:

Lemma 2.6.

(([20, Lemma 4.34.3])) Let w(z)w(z) be a non-zero solution of the differential operator LL in a differential extension 𝒦\mathcal{K} of K(z)K(z). In the ring 𝒦[ddz]\mathcal{K}[\tfrac{d}{dz}], we have the equalities:

Rn=1n!w(z)1dndznw(z)a(z)n=1n!R1(R1+a(z))(R1+(n1)a(z)).R_{n}=\dfrac{1}{n!}w(z)^{-1}\dfrac{d^{n}}{dz^{n}}w(z)a(z)^{n}=\dfrac{1}{n!}R_{1}(R_{1}+a^{\prime}(z))\cdots(R_{1}+(n-1)a^{\prime}(z)).

Making use of the Rodrigues operator associated with LL, we construct explicit Padé approximants of (fj)j(f_{j})_{j}. The following theorem is a particular case of [20, Theorem 4.24.2, Lemma 5.15.1].

Theorem 2.7.

For a non-negative integer \ell, define polynomials

(13) Pn,(z)=Rnz,Qn,j,(z)=φj(Pn,(z)Pn,(t)zt)for 0jw.\displaystyle P_{n,\ell}(z)=R_{n}\cdot z^{\ell},\ \ \ Q_{n,j,\ell}(z)=\varphi_{j}\left(\dfrac{P_{n,\ell}(z)-P_{n,\ell}(t)}{z-t}\right)\ \ \text{for}\ \ 0\leq j\leq w.

Then the following properties hold.

(i)(i) The vector of polynomials (Pn,(z),Qn,j,(z))(P_{n,\ell}(z),Q_{n,j,\ell}(z)) constitutes a weight nn Padé-type approximant for f0(z),,fw(z)f_{0}(z),\dots,f_{w}(z).

(ii)(ii) Put the Padé-type approximation of (fj)j(f_{j})_{j} by

(14) n,j,(z)=Pn,(z)fj(z)Qn,j,(z)for 0jw.\displaystyle\mathfrak{R}_{n,j,\ell}(z)=P_{n,\ell}(z)f_{j}(z)-Q_{n,j,\ell}(z)\ \ \text{for}\ \ 0\leq j\leq w.

We have

n,j,(z)=(1)nk=n(kn)φj(tk+na(t)n)zk+1.\displaystyle\mathfrak{R}_{n,j,\ell}(z)=(-1)^{n}\sum_{k=n}^{\infty}\binom{k}{n}\dfrac{\varphi_{j}(t^{k+\ell-n}{a}(t)^{n})}{z^{k+1}}.
Proof..

(i)(i) The statement is derived by applying [20, Theorem 4.24.2] in the special case where d=1d=1 and D1=LD_{1}=L.

(ii)(ii) By definition of n,j,\mathfrak{R}_{n,j,\ell}, we see n,j,(z)=φj(Pn,(t)/(zt))\mathfrak{R}_{n,j,\ell}(z)=\varphi_{j}\left(P_{n,\ell}(t)/(z-t)\right). Substituting the expansion 1/(zt)=k=0tk/zk+11/(z-t)=\sum_{k=0}^{\infty}t^{k}/z^{k+1} into this expression, we obtain

n,j,(z)=k=nφj(tkPn,(t))zk+1.\mathfrak{R}_{n,j,\ell}(z)=\sum_{k=n}^{\infty}\dfrac{\varphi_{j}(t^{k}P_{n,\ell}(t))}{z^{k+1}}.

Above equality implies it is sufficient to prove

(15) φj(tkPn,(t))=(1)n(kn)φj(t+kna(t)n)forkn.\displaystyle\varphi_{j}(t^{k}P_{n,\ell}(t))=(-1)^{n}\binom{k}{n}\varphi_{j}(t^{\ell+k-n}{a}(t)^{n})\ \ \text{for}\ \ k\geq n.

Now we fix a positive integer knk\geq n. To prove equation (15), we define the differential operator a,b=ddt+b(t)/a(t)K(t)[ddt]\mathcal{E}_{a,b}=\tfrac{d}{dt}+b(t)/a(t)\in K(t)[\tfrac{d}{dt}]. Notice L=a,ba(t)L^{*}=\mathcal{E}_{a,b}a(t). By [20, Proposition 3.2(i)3.2\,(i)], there exists a set of integers {cn,k,i; 0in}\{c_{n,k,i};\,0\leq i\leq n\} such that cn,k,n=(1)nk(k1)(kn+1)c_{n,k,n}=(-1)^{n}k(k-1)\cdots(k-n+1) and

tka,bn=i=0ncn,k,ia,bnitki.t^{k}\mathcal{E}^{n}_{a,b}=\sum_{i=0}^{n}c_{n,k,i}\mathcal{E}_{a,b}^{n-i}t^{k-i}.

This implies

(16) tkPn,(t)=tkn!a,bna(t)nt=i=0ncn,k,in!a,bnitkia(t)nt.\displaystyle t^{k}P_{n,\ell}(t)=\dfrac{t^{k}}{n!}\mathcal{E}^{n}_{a,b}a(t)^{n}\cdot t^{\ell}=\sum_{i=0}^{n}\dfrac{c_{n,k,i}}{n!}\mathcal{E}_{a,b}^{n-i}t^{k-i}a(t)^{n}\cdot t^{\ell}.

From [20, Proposition 3.2(ii)3.2\,(ii)], we know

(17) a,bnitkia(t)ntLK[t]for 0in1.\displaystyle\mathcal{E}_{a,b}^{n-i}t^{k-i}a(t)^{n}\cdot t^{\ell}\in L^{*}\cdot K[t]\ \ \text{for}\ \ 0\leq i\leq n-1.

Combining Equations (16), (17) together with LK[t]kerφjL^{*}\cdot K[t]\subseteq{\rm{ker}}\,\varphi_{j} (refer Proposition 2.4) deduces

φj(tkPn,(t))=φj(cn,k,nn!tk+na(t)n)=(1)n(kn)φj(tk+na(t)n).\varphi_{j}(t^{k}P_{n,\ell}(t))=\varphi_{j}\left(\dfrac{c_{n,k,n}}{n!}t^{k+\ell-n}a(t)^{n}\right)=(-1)^{n}\binom{k}{n}\varphi_{j}(t^{k+\ell-n}a(t)^{n}).

This establishes equation (15) and completes the proof of (ii)(ii). ∎

Let nn be a non-negative integer. Denote the determinant of the matrix formed by the Padé approximants of (fj)j(f_{j})_{j} in Theorem 2.7 by

Δn(z)=det(Pn,0(z)Pn,1(z)Pn,w+1(z)Qn,0,0(z)Qn,0,1(z)Qn,0,w+1(z)Qn,w,0(z)Qn,w,1(z)Qn,w,w+1(z)).\Delta_{n}(z)={\rm{det}}\begin{pmatrix}P_{n,0}(z)&P_{n,1}(z)&\cdots&P_{n,w+1}(z)\\ Q_{n,0,0}(z)&Q_{n,0,1}(z)&\cdots&Q_{n,0,w+1}(z)\\ \vdots&\vdots&\ddots&\vdots\\ Q_{n,w,0}(z)&Q_{n,w,1}(z)&\cdots&Q_{n,w,w+1}(z)\end{pmatrix}.

In the next section, let us consider the non-vanishing of Δn(z)\Delta_{n}(z).

3 Linear independence of Padé-type approximants

In this section, we keep the notation in Subsection 2.3. Recall the polynomials a(z),b(z)K[z]a(z),b(z)\in K[z] defined in equation (7) which satisfy (8) and (9). This section is devoted to prove the next theorem.

Theorem 3.1.

For the polynomials a(z),b(z)a(z),b(z), we assume

(18) na(z)+b(z)andb(z)are coprime for any positive integern,\displaystyle na^{\prime}(z)+b(z)\ \text{and}\ b(z)\ \text{are coprime for any positive integer}\ n,
(19) au((k+1)u+n)+bv0for anyk,n0ifu1=v.\displaystyle a_{u}((k+1)u+n)+b_{v}\neq 0\ \ \text{for any}\ \ k,n\geq 0\ \ \text{if}\ \ u-1=v.

Then Δn(z)K{0}\Delta_{n}(z)\in K\setminus\{0\} for any n0n\geq 0.

Our strategy of the proof of Theorem 3.1 is the following. By definition Δn(z)\Delta_{n}(z) is a polynomial. We show in Proposition 3.3, which is essentially an application of [10, Lemma 4.2 (ii)], that this polynomial is a constant, and we reduce the problem to showing that another determinant detn{\rm{det}}\,\mathcal{M}_{n} is non-zero. This last property, established in Subsection 3.2, will be a consequence of Theorem 3.1. In order to prove above results, let us prepare the following lemma.

Lemma 3.2.

Denote the differential operator a(z)ddz+b(z)-a(z)\tfrac{d}{dz}+b(z) by LL. Let kk\in\mathbb{N} and P(t)K[t]{0}P(t)\in K[t]\setminus\{0\}. For the polynomials a(z),b(z)a(z),b(z), we assume equation (19) holds. Then we have

deg(L+ka(t))P(t)=degP+w+1.{\rm{deg}}\,(L^{*}+ka^{\prime}(t))\cdot P(t)={\rm{deg}}\,P+w+1.
Proof..

We may assume P(t)=tNP(t)=t^{N} for a non-negative integer NN. Put

𝒫(t)=(L+ka(t))tN.\mathcal{P}(t)=\left(L^{*}+ka^{\prime}(t)\right)\cdot t^{N}.

Then a direct computation yields that deg𝒫N+w+1{\rm{deg}}\,\mathcal{P}\leq N+w+1 and the coefficient of tN+w+1t^{N+w+1} of 𝒫\mathcal{P} is au(u(k+1)+N)+bva_{u}(u(k+1)+N)+b_{v} if u1=vu-1=v, bvb_{v} if v>u1v>u-1 and au((k+1)u+N)a_{u}((k+1)u+N) if u1>vu-1>v. The assumption (19) ensures the assertion. This completes the proof of the statement. ∎

Let nn be a non-negative integer. Recall the polynomials Pn,(z),Qn,j,(z)P_{n,\ell}(z),Q_{n,j,\ell}(z) defined in equation (13). Combining Lemmas 2.6 and 3.2 yields

(20) degPn,=+n(w+1).\displaystyle{\rm{deg}}\,P_{n,\ell}=\ell+n(w+1).

Put the following w+1w+1 by w+1w+1 matrix by

n=(φ0(a(t)n)φ0(twa(t)n)φw(a(t)n)φw(twa(t)n))Matw+1(K).\displaystyle\mathcal{M}_{n}=\begin{pmatrix}\varphi_{0}(a(t)^{n})&\cdots&\varphi_{{0}}(t^{w}a(t)^{n})\\ \vdots&\ddots&\vdots\\ \varphi_{{w}}(a(t)^{n})&\cdots&\varphi_{{w}}(t^{w}a(t)^{n})\end{pmatrix}\in{\rm{Mat}}_{w+1}(K).

We now prove Δn(z)\Delta_{n}(z) is a constant.

Proposition 3.3.

There exists a constant cKc\in K such that cc is non-zero under the assumption (19) and

Δn(z)=cdetn.\Delta_{n}(z)=c\cdot{\rm{det}}\,\mathcal{M}_{n}.

In particular, we have Δn(z)K\Delta_{n}(z)\in K.

Proof..

Note that this proposition is a particular case of [20, Proposition 5.25.2]. For the matrix in the definition of Δn(z)\Delta_{n}(z), adding fj(z)-f_{j}(z) times first row to j+1j+1 th row for each 0jw0\leq j\leq w,

Δn(z)=(1)wdet(Pn,0(z)Pn,w+1(z)n,0,0(z)n,0,w+1(z)n,w,0(z)n,w,w+1(z)).\Delta_{n}(z)=(-1)^{w}{\rm{det}}{\begin{pmatrix}P_{n,0}(z)&\dots&P_{n,w+1}(z)\\ \mathfrak{R}_{n,0,0}(z)&\dots&\mathfrak{R}_{n,0,w+1}(z)\\ \vdots&\ddots&\vdots\\ \mathfrak{R}_{n,w,0}(z)&\dots&\mathfrak{R}_{n,w,w+1}(z)\end{pmatrix}}.

We denote the (s,t)(s,t)th cofactor of the matrix in the right hand side of above equality by Δs,t(z)\Delta_{s,t}(z). Then we have, developing along the first row

(21) Δn(z)=(1)w(l=0wPn,(z)Δ1,+1(z)).\displaystyle\Delta_{n}(z)=(-1)^{w}\left(\sum_{l=0}^{{{w}}}P_{n,\ell}(z)\Delta_{1,\ell+1}(z)\right).

The property of the Padé approximation ordn,j,(z)n+1{\rm{ord}}_{\infty}\,{{\mathfrak{R}}}_{n,j,\ell}(z)\geq n+1 for 0jw0\leq j\leq w, 0w+10\leq\ell\leq w+1 implies

ordΔ1,+1(z)(n+1)(w+1)for 0w.{\rm{ord}}_{\infty}\,\Delta_{1,\ell+1}(z)\geq(n+1)(w+1)\ \ \text{for}\ \ 0\leq\ell\leq w.

Combining equation (20) and above inequality yields

Pn,(z)Δ1,+1(z)(1/z)K[[1/z]]for 0w,P_{n,\ell}(z)\Delta_{1,\ell+1}(z)\in(1/z)\cdot K[[1/z]]\ \ \text{for}\ \ 0\leq\ell\leq w,

and

Pn,w+1(z)Δ1,w+1(z)K[[1/z]].P_{n,w+1}(z)\Delta_{1,{{w+1}}}(z)\in K[[1/z]].

Note that in above relation, the constant term of Pw(z)Δ1,w+1(z)P_{w}(z)\Delta_{1,{{w+1}}}(z) is

(22) Coefficient ofz(n+1)(w+1)ofPn,w+1(z)′′Coefficient of 1/z(n+1)(w+1)ofΔ1,w+1(z)′′.``\text{Coefficient of}\ z^{(n+1)(w+1)}\ \text{of}\ P_{n,w+1}(z)^{\prime\prime}\cdot``\text{Coefficient of}\ 1/z^{(n+1)(w+1)}\ \text{of}\ \Delta_{1,{{w+1}}}(z)^{\prime\prime}.

Notice that the coefficient of z(n+1)(w+1)z^{(n+1)(w+1)} of Pn,w+1(z)P_{n,w+1}(z) is non-zero under the assumption (19) by Lemma 3.2. equation (21)(\ref{formal power series rep delta}) implies Δn(z)\Delta_{n}(z) is a polynomial in zz with non-positive valuation with respect to ord{\rm{ord}}_{\infty}. Thus, it has to be a constant. Finally, by Theorem 2.7 (ii)(ii), the coefficient of 1/z(n+1)(w+1)1/z^{(n+1)(w+1)} of Δ1,w+1(z)\Delta_{1,{{w+1}}}(z) is

det((1)nφ0(a(t)n)(1)nφ0(twa(t)n)(1)nφw(a(t)n)(1)nφw(twa(t)n))=(1)wndetn.\displaystyle{\rm{det}}\begin{pmatrix}(-1)^{n}\varphi_{0}(a(t)^{n})&\ldots&(-1)^{n}\varphi_{0}(t^{w}a(t)^{n})\\ \vdots&\ddots&\vdots\\ (-1)^{n}\varphi_{w}(a(t)^{n})&\ldots&(-1)^{n}\varphi_{w}(t^{w}a(t)^{n})\\ \end{pmatrix}=(-1)^{wn}\cdot{\rm{det}}\,\mathcal{M}_{n}.

Combining Equations (21)(\ref{formal power series rep delta}), (22)(\ref{what const}) and above equality yields the assertion. This completes the proof of Proposition 3.3. ∎

3.1 Study of kernels of φj\varphi_{j}

Let nn be a non-negative integer. Denote K[t]nK[t]K[t]_{\leq n}\subset K[t] the KK-vector space of polynomials of degree at most nn. In this subsection, we consider the kernel of φj\varphi_{j} and prove the following crucial lemma.

Lemma 3.4.

We have

j=0wkerφj=LK[t].\bigcap_{j=0}^{w}{\rm{ker}}\,\varphi_{j}=L^{*}\cdot K[t].
Proof..

Denote the KK-vector space j=0wkerφj\cap_{j=0}^{w}{\rm{ker}}\,\varphi_{j} by WW. Since Lfj(z)K[z]L\cdot f_{j}(z)\in K[z], Proposition 2.4 yields LK[t]WL^{*}\cdot K[t]\subseteq W. Let us show the opposite inclusion. Let P(t)WP(t)\in W. Applying Lemma 3.2 for k=0k=0, we see that there exists a polynomial P~(t)LK[t]\tilde{P}(t)\in L^{*}\cdot K[t] such that P(t)P~(t)K[t]wP(t)-\tilde{P}(t)\in K[t]_{\leq w}. This implies that it is sufficient to prove the following equality:

(23) WK[t]w={0}.\displaystyle W\bigcap K[t]_{\leq w}=\{0\}.

Put fj(z)=k=0fj,k/zk+1f_{j}(z)=\sum_{k=0}^{\infty}f_{j,k}/z^{k+1}. Then, by the definition of φj\varphi_{j},

0=(f0,0f0,wfw,0fw,w)Matw+1(K).\mathcal{M}_{0}=\begin{pmatrix}f_{0,0}&\cdots&f_{0,w}\\ \vdots&\ddots&\vdots\\ f_{w,0}&\cdots&f_{w,w}\end{pmatrix}\in{\rm{Mat}}_{w+1}(K).

Since the Laurent series fjf_{j} is determined by (fj,k)0kw(f_{j,k})_{0\leq k\leq w} (see the KK-isomorphism in equation (11)) and the Laurent series {fj(z)}0jw\{f_{j}(z)\}_{0\leq j\leq w} are linearly independent over KK, we have det00{\rm{det}}\,\mathcal{M}_{0}\neq 0. For P(t)=k=0wpjtjWK[t]wP(t)=\sum_{k=0}^{w}p_{j}t^{j}\in W\cap K[t]_{\leq w}, we put 𝒑=(p0,,pw)t\boldsymbol{p}={}^{t}(p_{0},\ldots,p_{w}). By the definition of WW, we have

0𝒑=𝟎,\mathcal{M}_{0}\cdot\boldsymbol{p}=\boldsymbol{0},

and thus 𝒑=𝟎\boldsymbol{p}=\boldsymbol{0}. This implies P=0P=0 and equation (23) holds. ∎

Remark 3.5.

The non-vanishing of the determinant of 0\mathcal{M}_{0} yields the non-vanishing of Δ0(z)\Delta_{0}(z).

3.2 Proof of Theorem 3.1

Let nn be a positive integer. By Proposition 3.3, it is sufficient to show the non-vanishing of detn{\rm{det}}\,\mathcal{M}_{n} to prove Theorem 3.1. Let 𝒒=(q0,,qw)tKw\boldsymbol{q}={}^{t}(q_{0},\ldots,q_{w})\in K^{w} such that

n𝒒=𝟎.\mathcal{M}_{n}\cdot\boldsymbol{q}=\boldsymbol{0}.

Put Q(t)=k=0wqktkK[t]Q(t)=\sum_{k=0}^{w}q_{k}t^{k}\in K[t]. Then the linearity of φj\varphi_{j} derives φj(Q(t)a(t)n)=0\varphi_{j}(Q(t)a(t)^{n})=0 for 0jw0\leq j\leq w and thus, using Lemma 3.4,

(24) Q(t)a(t)nj=0wkerφj=LK[t].\displaystyle Q(t)a(t)^{n}\in\bigcap_{j=0}^{w}{\rm{ker}}\,\varphi_{j}=L^{*}\cdot K[t].

We complete the proof of Theorem 3.1 by showing Q(t)=0Q(t)=0. In order to prove Q(t)=0Q(t)=0, we prepare the following key lemma.

Lemma 3.6.

Let a(t),b(t)K[t]a(t),b(t)\in K[t]. Assume Na(t)+b(t)Na^{\prime}(t)+b(t) and a(t)a(t) are coprime for any positive integer NN. Let nn be a positive integer and P(t),Q(t)K[t]P(t),Q(t)\in K[t] such that Q(t)a(t)n=LP(t)Q(t)a(t)^{n}=L^{*}\cdot P(t). Then P(t)P(t) is divisible by a(t)na(t)^{n}.

Proof..

Let us prove the statement by induction on nn. Let n=1n=1. Then

Q(t)a(t)=(a(t)+b(t))P(t)+a(t)P(t).Q(t)a(t)=(a^{\prime}(t)+b(t))P(t)+a(t)P^{\prime}(t).

Since a(t)+b(t)a^{\prime}(t)+b(t) is coprime with a(t)a(t), the polynomial P(t)P(t) is divisible by a(t)a(t). Assume the claim holds for n1n\geq 1. Let us consider the case n+1n+1. The induction hypothesis yields P(t)P(t) is divisible by a(t)na(t)^{n}. Put P(t)=a(t)nP~(t)P(t)=a(t)^{n}\tilde{P}(t). Then

Q(t)a(t)n+1=((n+1)a(t)+b(t))P~(t)a(t)n+a(t)n+1P~(t).Q(t)a(t)^{n+1}=\left((n+1)a^{\prime}(t)+b(t)\right)\tilde{P}(t)a(t)^{n}+a(t)^{n+1}\tilde{P}^{\prime}(t).

Combining the hypothesis of a(t),b(t)a(t),b(t) and this equality yields P~(t)\tilde{P}(t) is divisible by a(t)a(t) and thus P(t)P(t) is divisible by a(t)n+1a(t)^{n+1}. ∎

We now finish the proof of Theorem 3.1.

Proof of Theorem 3.1.

Recall a(z),b(z)a(z),b(z) be polynomials defined in equation (7) satisfying the assumptions (8), (18) and (19). For a vector 𝒒=(q0,,qw)t\boldsymbol{q}={}^{t}(q_{0},\ldots,q_{w}) satisfying n𝒒=𝟎\mathcal{M}_{n}\cdot\boldsymbol{q}=\boldsymbol{0}, we put Q(t)=k=0wqktkQ(t)=\sum_{k=0}^{w}q_{k}t^{k}. Assume Q(t)0Q(t)\neq 0. From equation (24), there exists a non-zero polynomial P(t)K[t]P(t)\in K[t] such that

Q(t)a(t)n=LP(t).Q(t)a(t)^{n}=L^{*}\cdot P(t).

Using the assumption (18), by Lemma 3.6, the polynomial P(t)P(t) must be divisible by a(t)na(t)^{n} and thus we have degPnu{\rm{deg}}\,P\geq nu. Using the assumption (19) together with Lemma 3.2 for k=0k=0, we have

degLP(t)nu+w+1.{\rm{deg}}\,L^{*}\cdot P(t)\geq nu+w+1.

However, by the definition of Q(t)Q(t), we know

degQ(t)a(t)nnu+w.{\rm{deg}}\,Q(t)a(t)^{n}\leq nu+w.

This leads to a contradiction, as the degrees cannot simultaneously satisfy both conditions. Thus we conclude Q(t)=0Q(t)=0, completing the proof of Theorem 3.1. ∎

Remark 3.7.

Let KK be an algebraic number field, and let a(z),b(z)K[z]a(z),b(z)\in K[z] satisfy the assumptions (7), (8) and (9). Consider the differential operator L=a(z)ddz+b(z)L=-a(z)\tfrac{d}{dz}+b(z). In Theorem 1.1, we apply Theorem 3.1 to the case where dega=m{\deg}\,a=m and degbm1{\deg}\,b\leq m-1, under the assumptions (1), (2) and (3), and to Laurent series fj(z)(1/z)K[[1/z]]f_{j}(z)\in(1/z)\cdot K[[1/z]] that are linearly independent over KK and satisfy LfjK[z]L\cdot f_{j}\in K[z]. In this setting, as mentioned in Remark 1.2, the Laurent series fjf_{j} are all GG-functions.

In [2], André introduced the notion of arithmetic Gevrey series as a general framework encompassing the theories of EE-functions and GG-functions originally developed by Siegel [29] in his study of transcendental number theory (refer [3]). It would be interesting to explore possible applications of Theorem 3.1 to the study of the arithmetic properties of other classes of arithmetic Gevrey series f(z)(1/z)K[[1/z]]f(z)\in(1/z)\cdot K[[1/z]] satisfying LfK[z]L\cdot f\in K[z].

4 Estimates

Let us recall notation in Subsection 1.1. We fix an positive integer m2m\geq 2. Let KK be an algebraic number field and a(z),b(z)K[z]a(z),b(z)\in K[z] with a(z)a(z) being a monic polynomial satisfying

dega=manddegbm1.{\rm{deg}}\,a=m\ \ \ \text{and}\ \ {\rm{deg}}\,b\leq m-1.

Including multiplicities, denote the roots of the polynomial a(z)a(z) in ¯\overline{\mathbb{Q}} by α1,,αm\alpha_{1},\ldots,\alpha_{m} and

a(z)=i=0maizi,b(z)=j=0m1bjzj.a(z)=\sum_{i=0}^{m}a_{i}z^{i},\ \ b(z)=\sum_{j=0}^{m-1}b_{j}z^{j}.

Assume (1), (2) and (3). Denote the differential operator of order 11

L=a(z)ddz+b(z).L=-a(z)\dfrac{d}{dz}+b(z).

The assumption (3) together with Lemma 2.5 ensures that there exist Laurent series f0(z),,fm2(z)(1/z)K[[1/z]]f_{0}(z),\ldots,f_{m-2}(z)\in(1/z)\cdot K[[1/z]] such that fjf_{j} are linearly independent over KK and

(25) Lfj(z)K[z]for 0jm2.\displaystyle L\cdot f_{j}(z)\in K[z]\ \ \text{for}\ \ 0\leq j\leq m-2.

We now compute the exact form of fj(z)f_{j}(z). Recall si:=b(αi)/a(αi)s_{i}:=b(\alpha_{i})/a^{\prime}(\alpha_{i}) for 1im1\leq i\leq m and the quantity bm1=s1++smb_{m-1}=s_{1}+\ldots+s_{m} does not belong to <1\mathbb{Z}_{<-1}.

Lemma 4.1.

For 0jm20\leq j\leq m-2, the formal Laurent series

fj(z):=i=1m(1αiz)si1zj+1FD(m)(bm1+j+1, 1+s1,,1+sm,bm1+j+2;α1z,,αmz),f_{j}(z):=\prod_{i=1}^{m}\left(1-\frac{\alpha_{i}}{z}\right)^{s_{i}}\cdot\frac{1}{z^{j+1}}F^{(m)}_{D}\!\left(b_{m-1}+j+1,\,1+s_{1},\ldots,1+s_{m},\,b_{m-1}+j+2;\,\frac{\alpha_{1}}{z},\ldots,\frac{\alpha_{m}}{z}\right),

are linearly independent over KK and satisfy (25).

Proof..

Since ordfj=j+1{\rm{ord}}_{\infty}\,f_{j}=j+1, it is easy to see that fjf_{j} are linearly independent over KK. Let us show fjf_{j} satisfy (25). Set

Φ(z)\displaystyle\Phi(z) =i=1m(1αiz)si,\displaystyle=\prod_{i=1}^{m}\left(1-\frac{\alpha_{i}}{z}\right)^{s_{i}},
gj(z)\displaystyle g_{j}(z) =1zj+1FD(m)(bm1+j+1, 1+s1,,1+sm,bm1+j+2;α1z,,αmz).\displaystyle=\frac{1}{z^{j+1}}F^{(m)}_{D}\!\left(b_{m-1}+j+1,\,1+s_{1},\ldots,1+s_{m},\,b_{m-1}+j+2;\,\frac{\alpha_{1}}{z},\ldots,\frac{\alpha_{m}}{z}\right).

It is sufficient to prove that fj(z)=Φ(z)gj(z)f_{j}(z)=\Phi(z)g_{j}(z) satisfies

(26) (ddzb(z)a(z))Φ(z)gj(z)=(bm1+j+1)zmj2a(z).\left(\frac{d}{dz}-\frac{b(z)}{a(z)}\right)\Phi(z)g_{j}(z)=\frac{(b_{m-1}+j+1)z^{m-j-2}}{a(z)}.

A direct computation yields

(27) ddzΦ(z)\displaystyle\frac{d}{dz}\Phi(z) =Φ(z)zj=1mαjsjzαj,\displaystyle=\frac{\Phi(z)}{z}\sum_{j=1}^{m}\frac{\alpha_{j}s_{j}}{\,z-\alpha_{j}},
(zddz+bm1)gj(z)\displaystyle\Bigl(-z\frac{d}{dz}+b_{m-1}\Bigr)g_{j}(z) =bm1+j+1zj+1i=1m(1αiz)1si\displaystyle=-\frac{b_{m-1}+j+1}{z^{j+1}}\prod_{i=1}^{m}\left(1-\frac{\alpha_{i}}{z}\right)^{-1-s_{i}}
=(bm1+j+1)zmj1a(z)Φ(z).\displaystyle=-\frac{(b_{m-1}+j+1)z^{m-j-1}}{a(z)\Phi(z)}.

From the second equality we obtain

(28) gj(z)=bm1zgj(z)+(bm1+j+1)zmj2a(z)Φ(z).g^{\prime}_{j}(z)=\frac{b_{m-1}}{z}\,g_{j}(z)+\frac{(b_{m-1}+j+1)z^{m-j-2}}{a(z)\Phi(z)}.

Now, applying (27) and (28), we compute

(ddzb(z)a(z))Φ(z)gj(z)\displaystyle\left(\frac{d}{dz}-\frac{b(z)}{a(z)}\right)\Phi(z)g_{j}(z) =Φ(z)gj(z)+Φ(z)gj(z)b(z)a(z)Φ(z)gj(z)\displaystyle=\Phi^{\prime}(z)g_{j}(z)+\Phi(z)g^{\prime}_{j}(z)-\frac{b(z)}{a(z)}\Phi(z)g_{j}(z)
=[1z(j=1mαjsjzαj+bm1)b(z)a(z)]Φ(z)gj(z)+(bm1+j+1)zmj2a(z).\displaystyle=\left[\frac{1}{z}\left(\sum_{j=1}^{m}\frac{\alpha_{j}s_{j}}{z-\alpha_{j}}+b_{m-1}\right)-\frac{b(z)}{a(z)}\right]\Phi(z)g_{j}(z)+\frac{(b_{m-1}+j+1)z^{m-j-2}}{a(z)}.

Finally, observe the identity

zb(z)a(z)=zj=1msjzαj=j=1m(αjsjzαj+sj)=j=1mαjsjzαj+bm1,\frac{zb(z)}{a(z)}=z\sum_{j=1}^{m}\frac{s_{j}}{z-\alpha_{j}}=\sum_{j=1}^{m}\left(\frac{\alpha_{j}s_{j}}{z-\alpha_{j}}+s_{j}\right)=\sum_{j=1}^{m}\frac{\alpha_{j}s_{j}}{z-\alpha_{j}}+b_{m-1},

which implies

1z(j=1mαjsjzαj+bm1)b(z)a(z)=0.\frac{1}{z}\left(\sum_{j=1}^{m}\frac{\alpha_{j}s_{j}}{z-\alpha_{j}}+b_{m-1}\right)-\frac{b(z)}{a(z)}=0.

Therefore, (26) holds. This completes the proof of Lemma 4.1. ∎

Now we fix the Laurent series fj(z)f_{j}(z) in Lemma 4.1 and denote φfj\varphi_{f_{j}} by φj\varphi_{j}. We define the polynomials

(29) Pn,(z)=1n!(ddz+b(z)a(z))na(z)nz,Qn,j,(z)=φj(Pn,(z)Pn,(t)zt)for 0jm2.\displaystyle P_{n,\ell}(z)=\dfrac{1}{n!}\left(\dfrac{d}{dz}+\dfrac{b(z)}{a(z)}\right)^{n}a(z)^{n}\cdot z^{\ell},\ \ \ Q_{n,j,\ell}(z)=\varphi_{j}\left(\dfrac{P_{n,\ell}(z)-P_{n,\ell}(t)}{z-t}\right)\ \ \text{for}\ \ 0\leq j\leq m-2.

Then, by Theorem 2.7(i)\ref{Pade fj}\,(i), the vector of polynomials (Pn,(z),Qn,j,(z))0jm2(P_{n,\ell}(z),Q_{n,j,\ell}(z))_{0\leq j\leq m-2} is a weight nn Padé-type approximants of (fj)j(f_{j})_{j}. Denote the Padé-type approximations of (fj)j(f_{j})_{j} by

(30) n,j,(z)=Pn,(z)fj(z)Qn,j,(z)for 0jm2.\displaystyle\mathfrak{R}_{n,j,\ell}(z)=P_{n,\ell}(z)f_{j}(z)-Q_{n,j,\ell}(z)\ \ \text{for}\ 0\leq j\leq m-2.

A direct computation shows the algebraic function

w(z)=i=1m(zαi)siw(z)=\prod_{i=1}^{m}(z-\alpha_{i})^{s_{i}}

is a non-zero solution of the differential operator LL ((refer [14, Proposition 33 (i)])). Notice that Lemma 2.6 implies

(31) Pn,(z)=w(z)1n!dndznw(z)a(z)nz\displaystyle P_{n,\ell}(z)=\dfrac{w(z)^{-1}}{n!}\dfrac{d^{n}}{dz^{n}}w(z)\cdot a(z)^{n}z^{\ell}

and, applying the Leibniz formula to equation (31) along with the expression a(z)=i=1m(zαi)a(z)=\prod_{i=1}^{m}(z-\alpha_{i}) and the definition of w(z)w(z), the following equality holds.

Lemma 4.2.

We have

Pn,(z)=k=0n(1)k0k1,,kmkki=k0j1,,jm+1nkji=nk(jm+1)[i=1m(si)kiki!(nji)(zαi)njiki]zjm+1,\displaystyle P_{n,\ell}(z)=\sum_{k=0}^{n}(-1)^{k}\sum_{\begin{subarray}{c}0\leq k_{1},\ldots,k_{m}\leq k\\ \sum{k_{i}}=k\end{subarray}}\sum_{\begin{subarray}{c}0\leq j_{1},\ldots,j_{m+1}\leq n-k\\ \sum{j_{i}}=n-k\end{subarray}}\binom{\ell}{j_{m+1}}\left[\prod_{i=1}^{m}\dfrac{(-s_{i})_{k_{i}}}{k_{i}!}\binom{n}{j_{i}}(z-\alpha_{i})^{n-j_{i}-k_{i}}\right]z^{\ell-j_{m+1}},

with the convention (jm+1)=0\binom{\ell}{j_{m+1}}=0 if jm+1>j_{m+1}>\ell.

Proof..

The Leibniz formula yields

(32) dndznw(z)a(z)nz\displaystyle\dfrac{d^{n}}{dz^{n}}\cdot w(z)a(z)^{n}z^{\ell} =k=0n(nk)dkdzkw(z)×dnkdznka(z)nz.\displaystyle=\sum_{k=0}^{n}\binom{n}{k}\dfrac{d^{k}}{dz^{k}}\cdot w(z)\times\dfrac{d^{n-k}}{dz^{n-k}}\cdot a(z)^{n}z^{\ell}.

From the definition of w(z)w(z) and the equality a(z)=i=1m(zαi)a(z)=\prod_{i=1}^{m}(z-\alpha_{i}), we compute

dkdzkw(z)=w(z)0k1,,kmkki=k(1)kk!k1!km!i=1m(si)ki(zαi)ki,\displaystyle\dfrac{d^{k}}{dz^{k}}\cdot w(z)=w(z)\sum_{\begin{subarray}{c}0\leq k_{1},\ldots,k_{m}\leq k\\ \sum k_{i}=k\end{subarray}}(-1)^{k}\dfrac{k!}{k_{1}!\cdots k_{m}!}\prod_{i=1}^{m}\dfrac{(-s_{i})_{k_{i}}}{(z-\alpha_{i})^{k_{i}}},
dnkdznka(z)nz=(nk)!0j1,,jm+1nkji=nk(jm+1)zjm+1i=1m(nji)(zαi)nji.\displaystyle\dfrac{d^{n-k}}{dz^{n-k}}\cdot a(z)^{n}z^{\ell}=(n-k)!\sum_{\begin{subarray}{c}0\leq j_{1},\ldots,j_{m+1}\leq n-k\\ \sum{j_{i}}=n-k\end{subarray}}\binom{\ell}{j_{m+1}}z^{\ell-j_{m+1}}\prod_{i=1}^{m}\binom{n}{j_{i}}(z-\alpha_{i})^{n-j_{i}}.

Substituting (32) into equation (31) together with above equalities, we derive the expansion of Pn,(z)P_{n,\ell}(z). ∎

4.1 Absolute values of the Padé approximants

Let vv be a place of KK. In this subsection, we describe the asymptotic behavior, as nn goes to infinity, of the vv-adic absolue values of the polynomials Pn,(z)P_{n,\ell}(z) and Qn,j,(z)Q_{n,j,\ell}(z) evaluated at a fixed βK\beta\in K.

Now we use the following notations. Denote the vv-adic Weil height of mm-tuple of algebraic number 𝜶=(α1,,αm)\boldsymbol{\alpha}=(\alpha_{1},\ldots,\alpha_{m}) by Hv(𝜶)=exp(hv(𝜶)){\rm{H}}_{v}(\boldsymbol{\alpha})=\exp({\rm{h}}_{v}(\boldsymbol{\alpha})). For a polynomial P=k=0npkzkK[z]P=\sum_{k=0}^{n}p_{k}z^{k}\in K[z] and a valuation vv of KK, we denote the maximal vv-adic modulus of the coefficients of PP by

Pv:=max0kn{|pk|v}.||P||_{v}:=\max_{0\leq k\leq n}\{|p_{k}|_{v}\}.

The floor function is denoted by \lfloor\cdot\rfloor. For a non-negative integer nn, ss\in\mathbb{Q} and b1b\in\mathbb{Q}\setminus\mathbb{Z}_{\leq-1}, we denote

μn(s)\displaystyle\mu_{n}(s) =den(s)nq:primeqden(s)qn/q1,\displaystyle={\rm{den}}(s)^{n}\prod_{\begin{subarray}{c}q:\text{prime}\\ q\mid{\rm{den}}(s)\end{subarray}}q^{\lfloor n/q-1\rfloor},
dn(b)\displaystyle d_{n}(b) =den(1b+1,,1b+n+1).\displaystyle={\rm{den}}\left(\dfrac{1}{b+1},\ldots,\dfrac{1}{b+n+1}\right).

The aim of this section is to prove the following proposition.

Proposition 4.3.

Let vv be a place of KK and βK\beta\in K.

(i)(i) Assume vv is non-Archimedean. Then

logmax{|Pn,(β)|v,|Qn,j,(β)|v}\displaystyle\log\max\{|P_{n,\ell}(\beta)|_{v},|Q_{n,j,\ell}(\beta)|_{v}\} n(i=1mhv(αi)+(m1)(hv(𝜶)+hv(β)))\displaystyle\leq n\left(\sum_{i=1}^{m}{\rm{h}}_{v}(\alpha_{i})+(m-1)({\rm{h}}_{v}(\boldsymbol{\alpha})+{\rm{h}}_{v}(\beta))\right)
+mi=1mlog|μn(si)|v1+log|d(m1)(n+1)(bm1)|v+o(n),\displaystyle+m\sum_{i=1}^{m}\log|\mu_{n}(s_{i})|_{v}^{-1}+\log|d_{(m-1)(n+1)}(b_{m-1})|_{v}+o(n),

where o(n)=0o(n)=0 for almost all finite places vv.

(ii)(ii) Assume vv is Archimedean. Then

logmax{|Pn,(β)|v,|Qn,j,(β)|v}n(i=1mhv(αi)+(m1)(hv(𝜶)+hv(β))+mlog|4|v+o(1)).\displaystyle\log\max\{\left|P_{n,\ell}(\beta)\right|_{v},\left|Q_{n,j,\ell}(\beta)\right|_{v}\}\leq n\left(\sum_{i=1}^{m}{\rm{h}}_{v}(\alpha_{i})+(m-1)({\rm{h}}_{v}(\boldsymbol{\alpha})+{\rm{h}}_{v}(\beta))+m\log|4|_{v}+o(1)\right).

4.1.1 Proof of Proposition 4.3 (i)(i)

We will use the following classical lemma to control the denominator of (α)n/n!(\alpha)_{n}/n! and the growth of dn(b)d_{n}(b).

Lemma 4.4.

Let α\alpha\in\mathbb{Q} and b1b\in\mathbb{Q}\setminus\mathbb{Z}_{\leq-1}.

(i)(i) Let nn be a non-negative integer. For k=0,,nk=0,\dots,n, we have

μn(α)(α)kk!.\displaystyle\mu_{n}(\alpha)\dfrac{(\alpha)_{k}}{k!}\in\mathbb{Z}.

(ii)(ii) Let kk be an integer and n1,n2n_{1},n_{2} be positive integers. Then

μn(α+k)=μn(α)andμn1+n2(α)is divisible byμn1(α)μn2(α).\displaystyle\mu_{n}(\alpha+k)=\mu_{n}(\alpha)\ \ \text{and}\ \ \mu_{n_{1}+n_{2}}(\alpha)\ \text{is divisible by}\ \mu_{n_{1}}(\alpha)\mu_{n_{2}}(\alpha).

(iii)(iii) We have

lim supn1nlogdn(b)=den(b)φ(den(b))1jden(bm1)(j,den(bm1))=11j,\limsup_{n\to\infty}\dfrac{1}{n}\log d_{n}(b)=\dfrac{{\rm{den}}(b)}{\varphi({\rm{den}}(b))}\sum_{\begin{subarray}{c}1\leq j\leq{\rm{den}}(b_{m-1})\\ (j,{\rm{den}}(b_{m-1}))=1\end{subarray}}\dfrac{1}{j},

where φ\varphi is the Euler’s totient function.

(iv)(iv) Let pp be a prime. We have

limn|dn(b)|p1n=0.\displaystyle{\lim_{n\to\infty}}\dfrac{|d_{n}(b)|^{-1}_{p}}{n}=0.
Proof..

(i)(i) This property was proved in [6, Lemma 2.2].

(ii)(ii) These properties follow directly from the definition of μn(α)\mu_{n}(\alpha).

(iii)(iii), (iv)(iv) These statements were proved in [5]. ∎

Lemma 4.5.

Let vv be a non-Archimedean place of KK. Let f(z)=k=0fk/zk+1K[[1/z]]f(z)=\sum_{k=0}^{\infty}f_{k}/z^{k+1}\in K[[1/z]] and P(z)K[z]P(z)\in K[z] with degP=N{\rm{deg}}\,P=N. Put Q(z)=φf(P(z)P(t)zt)Q(z)=\varphi_{f}\left(\tfrac{P(z)-P(t)}{z-t}\right). Then we have

QvPvmax0kN1{|fk|v}.||Q||_{v}\leq||P||_{v}\cdot\max_{0\leq k\leq N-1}\{|f_{k}|_{v}\}.
Proof..

Put P(z)=i=0NpiziP(z)=\sum_{i=0}^{N}p_{i}z^{i}. Then, using the identity ziti=(zt)k=0i1ti1kzkz^{i}-t^{i}=(z-t)\sum_{k=0}^{i-1}t^{i-1-k}z^{k} for 1i1\leq i, we get

(33) Q(z)\displaystyle Q(z) =φf(k=0N1(i=k+1Npkti1k)zk)=k=0N1(i=k+1Npkfi1k)zk.\displaystyle=\varphi_{f}\left(\sum_{k=0}^{N-1}\left(\sum_{i=k+1}^{N}p_{k}t^{i-1-k}\right)z^{k}\right)=\sum_{k=0}^{N-1}\left(\sum_{i=k+1}^{N}p_{k}f_{i-1-k}\right)z^{k}.

Combining above equality with strong triangle inequality leads us to get

Qv=max0kN1{|i=k+1Npkfi1k|v}Pvmax0kN1{|fk|}.||Q||_{v}=\max_{0\leq k\leq N-1}\Biggl\{\left|\sum_{i=k+1}^{N}p_{k}f_{i-1-k}\right|_{v}\Biggr\}\leq||P||_{v}\cdot\max_{0\leq k\leq N-1}\{|f_{k}|\}.

Lemma 4.6.

Let f(z)=k=0fk/zk+1K[[1/z]]f(z)=\sum_{k=0}^{\infty}f_{k}/z^{k+1}\in K[[1/z]] be a Laurent series satisfying Lf(z)K[z]L\cdot f(z)\in K[z]. For any positive integer nn, there exists a positive integer DD which depends only on ff such that

(34) max0kn{|fk|}|D|v1Hv(𝜶)ni=1m|μn(si)|v1|dn(bm1)|v1\displaystyle\max_{0\leq k\leq n}\{|f_{k}|\}\leq|D|^{-1}_{v}{\rm{H}}_{v}(\boldsymbol{\alpha})^{n}\cdot\prod_{i=1}^{m}|\mu_{n}(s_{i})|^{-1}_{v}\cdot|d_{n}(b_{m-1})|^{-1}_{v}
Proof..

Put

fj(z)=k=0fj,kzk+1.f_{j}(z)=\sum_{k=0}^{\infty}\frac{f_{j,k}}{z^{k+1}}.

Since the Laurent series f(z)f(z) can be expressed as a KK-linear combination of fj(z)f_{j}(z) for 0jm20\leq j\leq m-2 (cf. isomorphism (11)), it suffices to prove that

(35) max0kn{|fj,k|}|bm1|v1Hv(𝜶)ni=1m|μn(si)|v1|dn(bm1)|v1.\displaystyle\max_{0\leq k\leq n}\{|f_{j,k}|\}\leq|b_{m-1}|^{-1}_{v}{\rm{H}}_{v}(\boldsymbol{\alpha})^{n}\cdot\prod_{i=1}^{m}|\mu_{n}(s_{i})|^{-1}_{v}\cdot|d_{n}(b_{m-1})|^{-1}_{v}.

We have the following expansions:

i=1m(1αiz)si=k=0[0kiki=ki=1m(si)kiki!αiki]1zk,\displaystyle\prod_{i=1}^{m}\left(1-\frac{\alpha_{i}}{z}\right)^{s_{i}}=\sum_{k=0}^{\infty}\left[\sum_{\begin{subarray}{c}0\leq k_{i}\\ \sum k_{i}=k\end{subarray}}\prod_{i=1}^{m}\frac{(-s_{i})_{k_{i}}}{k_{i}!}\alpha_{i}^{k_{i}}\right]\frac{1}{z^{k}},
FD(m)(bm1+j+1, 1+s1,,1+sm;bm1+j+2;α1z,,αmz)\displaystyle F^{(m)}_{D}\!\left(b_{m-1}+j+1,\,1+s_{1},\ldots,1+s_{m};\,b_{m-1}+j+2;\,\frac{\alpha_{1}}{z},\ldots,\frac{\alpha_{m}}{z}\right)
=k=0[0kiki=ki=1m(1+si)kiki!αiki]bm1+j+1bm1+j+k+11zk.\displaystyle=\sum_{k=0}^{\infty}\left[\sum_{\begin{subarray}{c}0\leq k_{i}\\ \sum k_{i}=k\end{subarray}}\prod_{i=1}^{m}\frac{(1+s_{i})_{k_{i}}}{k_{i}!}\alpha_{i}^{k_{i}}\right]\frac{b_{m-1}+j+1}{b_{m-1}+j+k+1}\frac{1}{z^{k}}.

From these we deduce that, for 0jm20\leq j\leq m-2 and k0k\geq 0,

(36) fj,k+j=w=0k[0kiwki=wi=1m(si)kiki!αiki][0likwli=kwi=1m(1+si)lili!αili]bm1+j+1bm1+j+kw+1.\displaystyle f_{j,k+j}=\sum_{w=0}^{k}\left[\sum_{\begin{subarray}{c}0\leq k_{i}\leq w\\ \sum k_{i}=w\end{subarray}}\prod_{i=1}^{m}\frac{(-s_{i})_{k_{i}}}{k_{i}!}\alpha_{i}^{k_{i}}\right]\cdot\left[\sum_{\begin{subarray}{c}0\leq l_{i}\leq k-w\\ \sum l_{i}=k-w\end{subarray}}\prod_{i=1}^{m}\frac{(1+s_{i})_{l_{i}}}{l_{i}!}\alpha_{i}^{l_{i}}\right]\cdot\frac{b_{m-1}+j+1}{b_{m-1}+j+k-w+1}.

Now let 0wkn0\leq w\leq k\leq n. For any choice of positive integers ki,lik_{i},l_{i} with ki=w\sum k_{i}=w and li=kw\sum l_{i}=k-w, Lemma 4.4(i)(i) and (ii)(ii) implies

|i=1m(si)kiki!(1+si)lili!αiki+li|vi=1m|μn(si)|v1Hv(𝜶)n.\left|\prod_{i=1}^{m}\frac{(-s_{i})_{k_{i}}}{k_{i}!}\frac{(1+s_{i})_{l_{i}}}{l_{i}!}\alpha_{i}^{k_{i}+l_{i}}\right|_{v}\leq\prod_{i=1}^{m}|\mu_{n}(s_{i})|^{-1}_{v}\,{\rm{H}}_{v}(\boldsymbol{\alpha})^{n}.

Therefore, by (36) and the strong triangle inequality, we get

max0kn{|fj,k|v}\displaystyle\max_{0\leq k\leq n}\{|f_{j,k}|_{v}\} |bm1|vi=1m|μn(si)|vHv(𝜶)nmax0kn{|1bm1+k+1|v}\displaystyle\leq|b_{m-1}|_{v}\prod_{i=1}^{m}|\mu_{n}(s_{i})|_{v}\,{\rm{H}}_{v}(\boldsymbol{\alpha})^{n}\cdot\max_{0\leq k\leq n}\Biggl\{\left|\frac{1}{b_{m-1}+k+1}\right|_{v}\Biggr\}
=|bm1|vi=1m|μn(si)|v1Hv(𝜶)n|dn(bm1)|v1.\displaystyle=|b_{m-1}|_{v}\prod_{i=1}^{m}|\mu_{n}(s_{i})|^{-1}_{v}\,{\rm{H}}_{v}(\boldsymbol{\alpha})^{n}\cdot|d_{n}(b_{m-1})|^{-1}_{v}.

This completes the proof of the assertion. ∎

Lemma 4.7.

Let vv be a non-Archimedean place and nn be a positive integer. Then

logmax0m{Pn,v}i=1m(log|μn(si)|v1+nhv(αi)).\log\max_{0\leq\ell\leq m}\{||P_{n,\ell}||_{v}\}\leq\sum_{i=1}^{m}\left(\log|\mu_{n}(s_{i})|_{v}^{-1}+n{\rm{h}}_{v}(\alpha_{i})\right).
Proof..

Put the equality

(zαi)njiki=wi=0njiki(njikiwi)(αi)njikiwizwifor 0kikn, 0jinki,(z-\alpha_{i})^{n-j_{i}-k_{i}}=\sum_{w_{i}=0}^{n-j_{i}-k_{i}}\binom{n-j_{i}-k_{i}}{w_{i}}(-\alpha_{i})^{n-j_{i}-k_{i}-w_{i}}z^{w_{i}}\ \ \text{for}\ 0\leq k_{i}\leq k\leq n,\ 0\leq j_{i}\leq n-k_{i},

into the expression of Pn,P_{n,\ell} obtained in Lemma 4.2. We then have

(37) Pn,(z)=k(1)kkiji(jm+1)[i=1m(si)kiki!(nji)wi(njikiwi)(αi)njikiwizwi]zjm+1.\displaystyle P_{n,\ell}(z)=\sum_{k}(-1)^{k}\sum_{k_{i}}\sum_{j_{i}}\binom{\ell}{j_{m+1}}\left[\prod_{i=1}^{m}\dfrac{(-s_{i})_{k_{i}}}{k_{i}!}\binom{n}{j_{i}}\sum_{w_{i}}\binom{n-j_{i}-k_{i}}{w_{i}}(-\alpha_{i})^{n-j_{i}-k_{i}-w_{i}}z^{w_{i}}\right]z^{\ell-j_{m+1}}.

Making use of Lemma 4.4 (i)(i) together with the strong triangle inequality for the above identity, we obtain the desire estimate. ∎

Proof of Proposition 4.3 (i)(i).

We apply Lemma 4.5 for P=Pn,P=P_{n,\ell}. Since degPn,=+(m1)n{\rm{deg}}\,P_{n,\ell}=\ell+(m-1)n, using Lemma 4.6, the value logmaxj,{Qn,j,v}\log\max_{j,\ell}\{||Q_{n,j,\ell}||_{v}\} is bounded by

Pn,v+(m1)nhv(𝜶)+i=1mlog|μ(m1)(n+1)(si)|v1+log|d(m1)(n+1)(bm1)|v+|bm1|v.||P_{n,\ell}||_{v}+(m-1)n{\rm{h}}_{v}(\boldsymbol{\alpha})+\sum_{i=1}^{m}\log|\mu_{(m-1)(n+1)}(s_{i})|^{-1}_{v}+\log|d_{(m-1)(n+1)}(b_{m-1})|_{v}+|b_{m-1}|_{v}.

Finally, Lemma 4.7 together with the strong triangle inequality yields

logmax{|Pn,(β)|v,|Qn,j,(β)|v}\displaystyle\log\max\{|P_{n,\ell}(\beta)|_{v},|Q_{n,j,\ell}(\beta)|_{v}\}\leq
i=1m(mlog|μn(si)|v1+nhv(αi))+(m1)n(hv(𝜶)+hv(β))+log|d(m1)(n+1)(bm1)|v+o(n),\displaystyle\sum_{i=1}^{m}\left(m\log|\mu_{n}(s_{i})|_{v}^{-1}+n{\rm{h}}_{v}(\alpha_{i})\right)+(m-1)n({\rm{h}}_{v}(\boldsymbol{\alpha})+{\rm{h}}_{v}(\beta))+\log|d_{(m-1)(n+1)}(b_{m-1})|_{v}+o(n),

where o(n)=0o(n)=0 if vv does not divide den(bm1,s1,,sm){\rm{den}}(b_{m}-1,s_{1},\ldots,s_{m}). This completes the proof of Proposition 4.3 (i)(i). ∎

4.1.2 Poincaré-Perron type recurrence

We now turn attention to the proof of Proposition 4.3 (ii)(ii). To this end, in this subsection, let us consider the following Poincaré-type recurrence of some order m>0m>0.

(38) am(n)u(n+m)+am1(n)u(n+m1)++a0(n)u(n)=0\displaystyle a_{m}(n)u(n+m)+a_{m-1}(n)u(n+m-1)+\cdots+a_{0}(n)u(n)=0

for large enough nn, where the coefficients ai(t)[t]a_{i}(t)\in\mathbb{C}[t] are polynomials and am(t)0a_{m}(t)\neq 0. Then, we can apply Perron’s Second Theorem below (see [26] and [27, Theorem C]) to estimate precisely the growth of a solution of the above recurrence.

Theorem 4.8 (Perron’s Second Theorem).

Let mm be a positive integer. Assume that for i=0,,mi=0,\dots,m there exist a function ai:0a_{i}:\mathbb{Z}_{\geq 0}\rightarrow\mathbb{C} and aia_{i}\in\mathbb{C} such that

limnai(n)=ai,\displaystyle\lim_{n\to\infty}a_{i}(n)=a_{i}\in\mathbb{C},

with am0a_{m}\neq 0. Denote by λ1,,λm\lambda_{1},\dots,\lambda_{m} the (not necessarily distinct) roots of the characteristic polynomial

χ(z)=amzm+am1zm1++a0.\displaystyle\chi(z)=a_{m}z^{m}+a_{m-1}z^{m-1}+\cdots+a_{0}.

Then, there exist mm linearly independent solutions u1,,umu_{1},\dots,u_{m} of (38), such that, for each i=1,,mi=1,\dots,m,

lim supn|ui(n)|1/n=|λi|.\displaystyle\limsup_{n\to\infty}|u_{i}(n)|^{1/n}=|\lambda_{i}|.

In particular, any solution uu of (38) satisfies lim supn|u(n)|1/nmax1im{|λi|}\limsup_{n\to\infty}|u(n)|^{1/n}\leq\max_{1\leq i\leq m}\{|\lambda_{i}|\}.

Remark 4.9.

In the above theorem, there are no restriction on the roots of χ(z)\chi(z), whereas in Poincaré’s Theorem and Perron’s First Theorem, we ask that

(39) |λi||λj| for ij,\displaystyle|\lambda_{i}|\neq|\lambda_{j}|\textrm{ for }i\neq j,

see [27, Theorem A and B].

Corollary 4.10.

Let us fix an embedding KK into \mathbb{C}. Let f(z)=k=0fk/zk+1(1/z)[[1/z]]f(z)=\sum_{k=0}^{\infty}f_{k}/z^{k+1}\in(1/z)\cdot\mathbb{C}[[1/z]] be a Laurent series satisfying Lf(z)[z]L\cdot f(z)\in\mathbb{C}[z]. Then we have

lim supn|fn|1/nmax1im{|αi|}.\limsup_{n\to\infty}|f_{n}|^{1/n}\leq\max_{1\leq i\leq m}\{|\alpha_{i}|\}.
Proof..

Equation (12) in Lemma 2.5 yields that the vector (fn)n1(f_{n})_{n\geq-1} with f1=0f_{-1}=0 is a solution of the recurrence equation:

i=0mai(k+i)fk+i1+j=0m1bjfk+j=0fork0.\sum_{i=0}^{m}a_{i}(k+i)f_{k+i-1}+\sum_{j=0}^{m-1}b_{j}f_{k+j}=0\ \ \text{for}\ \ k\geq 0.

A straightforward computation yields that this recurrence equation is Poincaré type and the characteristic polynomial χ(z)\chi(z) is a(z)a(z). Thus Perron’s Second Theorem ensures

lim supn|fn|1/nmax1im{|αi|}.\limsup_{n\to\infty}|f_{n}|^{1/n}\leq\max_{1\leq i\leq m}\{|\alpha_{i}|\}.

This completes the proof of Corollary 4.10. ∎

4.1.3 Proof of Proposition 4.3 (ii)(ii)

Lemma 4.11.

Let vv be an Archimedean valuation of KK. Define a constant cv(𝛂)c_{v}(\boldsymbol{\alpha}) which depends on 𝛂=(α1,,αm)\boldsymbol{\alpha}=(\alpha_{1},\ldots,\alpha_{m}) and vv by

(40) cv(𝜶)=i=1mhv(αi)+mlog|4|v.\displaystyle c_{v}(\boldsymbol{\alpha})=\sum_{i=1}^{m}{\rm{h}}_{v}(\alpha_{i})+m\log|4|_{v}.

Then we have

logmax0m1{Pn,(z)v}ncv(𝜶)+o(n).\log\max_{0\leq\ell\leq m-1}\{||P_{n,\ell}(z)||_{v}\}\leq nc_{v}(\boldsymbol{\alpha})+o(n).
Proof..

We apply the estimate

(s)kk!eo(n),(nk)2nforsand 0kn,\displaystyle\dfrac{(s)_{k}}{k!}\leq e^{o(n)},\ \ \binom{n}{k}\leq 2^{n}\ \ \text{for}\ \ s\in\mathbb{Q}\ \ \text{and}\ \ 0\leq k\leq n,

together with the triangular inequality for equation (37), we conclude the desire estimate. ∎

Lemma 4.12.

Let nn be a positive integer and P(z)K[z]P(z)\in K[z] with degP=n{\rm{deg}}\,P=n. Let vv be an Archimedean valuation of KK and βK\beta\in K. Let f(z)=k=0fk/zk+1(1/z)K[[1/z]]f(z)=\sum_{k=0}^{\infty}f_{k}/z^{k+1}\in(1/z)\cdot K[[1/z]] be a Laurent series satisfying Lf(z)K[z]L\cdot f(z)\in K[z]. Put

Q(z)=φf(P(z)P(t)zt).Q(z)=\varphi_{f}\left(\dfrac{P(z)-P(t)}{z-t}\right).

Then the following inequalities hold.

(i)(i) |P(β)|v(n+1)PvHv(β)n|P(\beta)|_{v}\leq(n+1)||P||_{v}{\rm{H}}_{v}(\beta)^{n}.

(ii)(ii) QvnPvHv(𝛂)n+o(n)||Q||_{v}\leq n||P||_{v}{\rm{H}}_{v}(\boldsymbol{\alpha})^{n+o(n)}.

(iii) |Q(β)|vn2PvHv(𝛂)n+o(n)Hv(β)n1|Q(\beta)|_{v}\leq n^{2}||P||_{v}{\rm{H}}_{v}(\boldsymbol{\alpha})^{n+o(n)}{\rm{H}}_{v}(\beta)^{n-1}.

Proof..

(i)(i) This is an easy consequence of the triangle inequality.

(ii)(ii) Equation (33) together with the triangle inequality and the estimate

max0kn1{|fk|v}Hv(𝜶)n+o(n)\max_{0\leq k\leq n-1}\{|f_{k}|_{v}\}\leq{\rm{H}}_{v}(\boldsymbol{\alpha})^{n+o(n)}

deduced from Corollary 4.10 allows us to obtain the desire inequality.

(iii) Combining (i)(i) and (ii)(ii) yields

|Q(β)|vnQvHv(β)n1n2PvHv(𝜶)n+o(n)Hv(β)n1.|Q(\beta)|_{v}\leq n||Q||_{v}{\rm{H}}_{v}(\beta)^{n-1}\leq n^{2}||P||_{v}{\rm{H}}_{v}(\boldsymbol{\alpha})^{n+o(n)}{\rm{H}}_{v}(\beta)^{n-1}.

This completes the proof of Lemma 4.12. ∎

Proof of Proposition 4.3 (ii)(ii).

Let cv(𝜶)c_{v}(\boldsymbol{\alpha}) be the constant defined in (40). We apply Lemma 4.12 (i)(i) with P=Pn,P=P_{n,\ell}. Using Lemma 4.11 and the equality degPn,=+(m1)n{\rm{deg}}\,P_{n,\ell}=\ell+(m-1)n yields

(41) log|Pn,(β)|vn(cv(𝜶)+(m1)hv(β)+o(1)).\displaystyle\log\,|P_{n,\ell}(\beta)|_{v}\leq n\left(c_{v}(\boldsymbol{\alpha})+(m-1){\rm{h}}_{v}(\beta)+o(1)\right).

For the estimating of Qn,j,Q_{n,j,\ell}, we invoke Lemma 4.12 (iii), which gives

(42) log|Qn,j,(β)|vn(cv(𝜶)+(m1)(hv(𝜶)+hv(β))+o(1)).\displaystyle\log\,|Q_{n,j,\ell}(\beta)|_{v}\leq n\left(c_{v}(\boldsymbol{\alpha})+(m-1)\left({\rm{h}}_{v}(\boldsymbol{\alpha})+{\rm{h}}_{v}(\beta)\right)+o(1)\right).

Combining inequalities (41) and (42), we arrive at the desired estimate. ∎

4.2 Absolute values of the Padé approximations

For a rational number ss and a place vv of KK, we define

μv(s)={1if v𝔐K or v𝔐Kf & |α|v1,|den(s)|v|p|v1p1if v𝔐Kf & |α|v>1 where p is the prime below v.\mu_{v}(s)=\left\{\begin{array}[]{ll}1&\mbox{if $v\in\mathfrak{M}^{\infty}_{K}$ or $v\in\mathfrak{M}^{f}_{K}$ \& $|\alpha|_{v}\leq 1$},\\ |{\rm{den}}(s)|_{v}|p|_{v}^{\tfrac{1}{p-1}}&\mbox{if $v\in\mathfrak{M}^{f}_{K}$ \& $|\alpha|_{v}>1$ where $p$ is the prime below $v$}.\end{array}\right.
Proposition 4.13.

Let vv be a place of KK and βK\beta\in K.

(i)(i) Assume vv is non-Archimedean and

(43) |β|v>i=1mμv(si)1Hv(𝜶).\displaystyle|\beta|_{v}>\prod_{i=1}^{m}\mu_{v}(s_{i})^{-1}\cdot{\rm{H}}_{v}(\boldsymbol{\alpha}).

Then the series n,j,(z)\mathfrak{R}_{n,j,\ell}(z) converges to an element of KvK_{v} at z=βz=\beta and

logmax0jm20m1{|n,j,(β)|v}n(log|β|v+i=1mhv(αi)+mhv(𝜶)mi=1mlogμv(si)+o(1)).\log\max_{\begin{subarray}{c}0\leq j\leq m-2\\ 0\leq\ell\leq m-1\end{subarray}}\{|\mathfrak{R}_{n,j,\ell}(\beta)|_{v}\}\leq n\left(-\log|\beta|_{v}+\sum_{i=1}^{m}{\rm{h}}_{v}(\alpha_{i})+m{\rm{h}}_{v}(\boldsymbol{\alpha})-m\sum_{i=1}^{m}\log\mu_{v}(s_{i})+o(1)\right).

(ii)(ii) Assume vv is Archimedean and |β|v>max{|αi|v}|\beta|_{v}>\max\{|\alpha_{i}|_{v}\}. Then the series n,j,(z)\mathfrak{R}_{n,j,\ell}(z) converges to an element of KvK_{v} at z=βz=\beta and

logmax0jm20m1{|n,j,(β)|v}n(log|β|v+i=1mhv(αi)+mhv(𝜶)+mlog|2|v+o(1)).\log\max_{\begin{subarray}{c}0\leq j\leq m-2\\ 0\leq\ell\leq m-1\end{subarray}}\{|\mathfrak{R}_{n,j,\ell}(\beta)|_{v}\}\leq n\left(-\log|\beta|_{v}+\sum_{i=1}^{m}{\rm{h}}_{v}(\alpha_{i})+m{\rm{h}}_{v}(\boldsymbol{\alpha})+m\log|2|_{v}+o(1)\right).
Proof..

Before starting the proof, we give a expansion of n,j,(z)\mathfrak{R}_{n,j,\ell}(z). Combining the expansion

tk+a(t)n=w=0mnkinki=wi=1m(nki)(αi)nkitw+k+,t^{k+\ell}a(t)^{n}=\sum_{w=0}^{mn}\sum_{\begin{subarray}{c}k_{i}\leq n\\ \sum{k_{i}=w}\end{subarray}}\prod_{i=1}^{m}\binom{n}{k_{i}}(-\alpha_{i})^{n-k_{i}}t^{w+k+\ell},

and Theorem 2.7 (ii)(ii) implies

(44) n,j,(z)=(1)nzn+1k=0(k+nn)(w=0mnkinki=wi=1m(nki)(αi)nkifw+k+)zk.\displaystyle\mathfrak{R}_{n,j,\ell}(z)=\dfrac{(-1)^{n}}{z^{n+1}}\sum_{k=0}^{\infty}\binom{k+n}{n}\left(\sum_{w=0}^{mn}\sum_{\begin{subarray}{c}k_{i}\leq n\\ \sum{k_{i}}=w\end{subarray}}\prod_{i=1}^{m}\binom{n}{k_{i}}(-\alpha_{i})^{n-k_{i}}f_{w+k+\ell}\right)\cdot z^{-k}.

(i)(i) Let vv be a non-Archimedean valuation. By the strong triangle inequality together with Lemma 4.6, we obtain

(45) |(k+nn)w=0mnkinki=wi=1m(nki)(αi)nkifw+k+βk|v\displaystyle\left|\binom{k+n}{n}\sum_{w=0}^{mn}\sum_{\begin{subarray}{c}k_{i}\leq n\\ \sum k_{i}=w\end{subarray}}\prod_{i=1}^{m}\binom{n}{k_{i}}(-\alpha_{i})^{n-k_{i}}f_{w+k+\ell}\,\beta^{-k}\right|_{v}\leq
i=1mHv(αi)nHv(𝜶)mn+k+mi=1m|μmn+k+m(si)|v1|dmn+k+m(bm1)|v1|bm1|v1|β|vk,\displaystyle\quad\prod_{i=1}^{m}{\rm H}_{v}(\alpha_{i})^{n}\cdot{\rm H}_{v}(\boldsymbol{\alpha})^{mn+k+m}\cdot\prod_{i=1}^{m}|\mu_{mn+k+m}(s_{i})|_{v}^{-1}\cdot|d_{mn+k+m}(b_{m-1})|_{v}^{-1}\cdot|b_{m-1}|_{v}^{-1}|\beta|_{v}^{-k},

for every k0k\geq 0.

By Lemma 4.4 (ii) and the divisibility relation

dmn+k+m(bm1)dm(n+1)(bm1)dk(bm1+m(n+1)),d_{mn+k+m}(b_{m-1})\mid d_{m(n+1)}(b_{m-1})\,d_{k}(b_{m-1}+m(n+1)),

we have

i=1m|μmn+k+m(si)|v1|dmn+k+m(bm1)|v1\displaystyle\prod_{i=1}^{m}|\mu_{mn+k+m}(s_{i})|_{v}^{-1}\,|d_{mn+k+m}(b_{m-1})|_{v}^{-1}
i=1m|μm(n+1)(si)|v1|d(m+1)n(bm1)|v1i=1m|μk(si)|v1|dk(bm1+m(n+1))|v1.\displaystyle\qquad\leq\prod_{i=1}^{m}|\mu_{m(n+1)}(s_{i})|_{v}^{-1}\,|d_{(m+1)n}(b_{m-1})|_{v}^{-1}\prod_{i=1}^{m}|\mu_{k}(s_{i})|_{v}^{-1}\,|d_{k}(b_{m-1}+m(n+1))|_{v}^{-1}.

Combining this with (45), we obtain

|(k+nn)w=0mnkinki=wi=1m(nki)(αi)nkifw+k+βk|v\displaystyle\left|\binom{k+n}{n}\sum_{w=0}^{mn}\sum_{\begin{subarray}{c}k_{i}\leq n\\ \sum k_{i}=w\end{subarray}}\prod_{i=1}^{m}\binom{n}{k_{i}}(-\alpha_{i})^{n-k_{i}}f_{w+k+\ell}\,\beta^{-k}\right|_{v} i=1m|μm(n+1)(si)|v1|d(m+1)n(bm1)|v1i=1mHv(αi)nHv(𝜶)m(n+1)\displaystyle\leq\prod_{i=1}^{m}|\mu_{m(n+1)}(s_{i})|_{v}^{-1}|d_{(m+1)n}(b_{m-1})|_{v}^{-1}\prod_{i=1}^{m}{\rm H}_{v}(\alpha_{i})^{n}{\rm H}_{v}(\boldsymbol{\alpha})^{m(n+1)}
|bm1|v1|dk(bm1+m(n+1))|v1Hv(𝜶)ki=1m|μk(si)|v1|β|vk.\displaystyle\qquad\cdot|b_{m-1}|_{v}^{-1}|d_{k}(b_{m-1}+m(n+1))|_{v}^{-1}{\rm{H}}_{v}(\boldsymbol{\alpha})^{k}\prod_{i=1}^{m}|\mu_{k}(s_{i})|_{v}^{-1}|\beta|_{v}^{-k}.

Since Lemma 4.4 (iv) implies that

|dk(bm1+m(n+1))|v1=o(k)(k),|d_{k}(b_{m-1}+m(n+1))|_{v}^{-1}=o(k)\qquad(k\to\infty),

and by assumption (43), we deduce that

lim supk|bm1|v1|dk(bm1+m(n+1))|v1Hv(𝜶)ki=1m|μk(si)|v1|β|vk=0.\limsup_{k\to\infty}|b_{m-1}|_{v}^{-1}|d_{k}(b_{m-1}+m(n+1))|_{v}^{-1}{\rm{H}}_{v}(\boldsymbol{\alpha})^{k}\prod_{i=1}^{m}|\mu_{k}(s_{i})|_{v}^{-1}|\beta|_{v}^{-k}=0.

Hence n,j,(z)\mathfrak{R}_{n,j,\ell}(z) converges to an element of KvK_{v} at z=βz=\beta, and

log|n,j,(β)|v\displaystyle\log|\mathfrak{R}_{n,j,\ell}(\beta)|_{v} log|β|vn1+logmaxk0{|(k+nn)w=0mnkinki=wi=1m(nki)(αi)nkifw+k+βk|v}\displaystyle\leq\log|\beta|_{v}^{-n-1}+\log\max_{k\geq 0}\left\{\left|\binom{k+n}{n}\sum_{w=0}^{mn}\sum_{\begin{subarray}{c}k_{i}\leq n\\ \sum k_{i}=w\end{subarray}}\prod_{i=1}^{m}\binom{n}{k_{i}}(-\alpha_{i})^{n-k_{i}}f_{w+k+\ell}\,\beta^{-k}\right|_{v}\right\}
n(log|β|v+i=1mhv(αi)+mhv(𝜶)mi=1mlogμv(si)+o(1)).\displaystyle\leq n\left(-\log|\beta|_{v}+\sum_{i=1}^{m}{\rm h}_{v}(\alpha_{i})+m\,{\rm h}_{v}(\boldsymbol{\alpha})-m\sum_{i=1}^{m}\log\mu_{v}(s_{i})+o(1)\right).

(ii)(ii) Let vv be an Archimedean valuation. The triangle inequality and Corollary 4.10 lead us to

|w=0mnkinki=wi=1m(nki)(αi)nkifw+k+|veo(n)|2|vmni=1mHv(αi)nmax{|αi|v}mn+k\left|\sum_{w=0}^{mn}\sum_{\begin{subarray}{c}k_{i}\leq n\\ \sum{k_{i}}=w\end{subarray}}\prod_{i=1}^{m}\binom{n}{k_{i}}(-\alpha_{i})^{n-k_{i}}f_{w+k+\ell}\right|_{v}\leq e^{o(n)}|2|^{mn}_{v}\prod_{i=1}^{m}{\rm{H}}_{v}(\alpha_{i})^{n}\max\{|\alpha_{i}|_{v}\}^{mn+k}

for any k0k\geq 0. Therefore combining (44) and the assumption |β|v>max{|αi|v}|\beta|_{v}>\max\{|\alpha_{i}|_{v}\} yields that n,j,(z)\mathfrak{R}_{n,j,\ell}(z) converges at β\beta in KvK_{v} and

|n,j,(β)|v\displaystyle\left|\mathfrak{R}_{n,j,\ell}(\beta)\right|_{v} eo(n)|β|vn|2|vmni=1mHv(αi)nmax{|αi|v}mnk=0(k+nn)(max{|αi|v}|β|v)k\displaystyle\leq e^{o(n)}|\beta|^{-n}_{v}|2|^{mn}_{v}\prod_{i=1}^{m}{\rm{H}}_{v}(\alpha_{i})^{n}\max\{|\alpha_{i}|_{v}\}^{mn}\sum_{k=0}^{\infty}\binom{k+n}{n}\left(\dfrac{\max\{|\alpha_{i}|_{v}\}}{|\beta|_{v}}\right)^{k}
eo(n)|β|vn|2|vmni=1mHv(αi)nHv(𝜶)mn.\displaystyle\leq e^{o(n)}|\beta|^{-n}_{v}|2|^{mn}_{v}\prod_{i=1}^{m}{\rm{H}}_{v}(\alpha_{i})^{n}{\rm{H}}_{v}(\boldsymbol{\alpha})^{mn}.

Taking the logarithm, we obtain the conclusion of (ii)(ii). ∎

5 Proof of Theorem 1.1

We keep the notation in Section 4. For a positive integer nn, we recall that the polynomials Pn,(z)P_{n,\ell}(z) and Qn,j,(z)Q_{n,j,\ell}(z) are defined in equation (29). Let us fix a place v0v_{0} of KK and let βK\beta\in K. Define the following mm by mm matrix MnM_{n} as

Mn=(Pn,0(β)Pn,1(β)Pn,m1(β)Qn,0,0(β)Qn,0,1(β)Qn,0,m1(β)Qn,m2,0(β)Qn,m2,1(β)Qn,m2,m1(β))Matm(K).M_{n}=\begin{pmatrix}P_{n,0}(\beta)&P_{n,1}(\beta)&\cdots&P_{n,m-1}(\beta)\\ Q_{n,0,0}(\beta)&Q_{n,0,1}(\beta)&\cdots&Q_{n,0,m-1}(\beta)\\ \vdots&\vdots&\ddots&\vdots\\ Q_{n,m-2,0}(\beta)&Q_{n,m-2,1}(\beta)&\cdots&Q_{n,m-2,m-1}(\beta)\end{pmatrix}\in{\rm{Mat}}_{m}(K).

Our proof relies on a qualitative linear independence criterion [9, Proposition 5.65.6] which is based on the method of C. F. Siegel (see [29]). Define real numbers:

𝔸v0(β)={log|β|v0i=1mhv0(αi)mhv0(𝜶)+mi=1mlogμv0(si)ifv,log|β|v0i=1mhv0(αi)mhv0(𝜶)mlog|2|v0ifv,\displaystyle\mathbb{A}_{v_{0}}(\beta)=\begin{cases}\log|\beta|_{v_{0}}-\sum_{i=1}^{m}{\rm{h}}_{v_{0}}(\alpha_{i})-m{\rm{h}}_{v_{0}}(\boldsymbol{\alpha})+m\sum_{i=1}^{m}\log\mu_{v_{0}}(s_{i})&\ \text{if}\ v\nmid\infty,\\ \log|\beta|_{v_{0}}-\sum_{i=1}^{m}{\rm{h}}_{v_{0}}(\alpha_{i})-m{\rm{h}}_{v_{0}}(\boldsymbol{\alpha})-m\log|2|_{v_{0}}&\ \text{if}\ v\mid\infty,\end{cases}
Uv0(β)=(i=1mhv(αi)+(m1)(hv(𝜶)+hv(β)))+mi=1mlogμv0(si).\displaystyle U_{v_{0}}(\beta)=\left(\sum_{i=1}^{m}{\rm{h}}_{v}(\alpha_{i})+(m-1)({\rm{h}}_{v}(\boldsymbol{\alpha})+{\rm{h}}_{v}(\beta))\right)+m\sum_{i=1}^{m}\log\mu_{v_{0}}(s_{i}).

We now restate Theorem 1.1 together with a linear independence measure.

Theorem 5.1.

We use the same notations in Theorem 1.1. Let v0𝔐Kv_{0}\in\mathfrak{M}_{K} such that Vv0(β)>0V_{v_{0}}(\beta)>0. Then the series fj(z)f_{j}(z) for 0jm20\leq j\leq m-2 converge around β\beta in Kv0K_{v_{0}} and for any positive number ε\varepsilon with ε<Vv0(β)\varepsilon<V_{v_{0}}(\beta), there exists an effectively computable positive number H0H_{0} depending on ε\varepsilon and the given data such that the following property holds. For any 𝛌=(λ,λj)0jm2Km{𝟘}{{\boldsymbol{\lambda}}}=({{\lambda}},{{\lambda_{j}}})_{0\leq j\leq m-2}\in K^{m}\setminus\{\mathbb{0}\} satisfying H0H(𝛌)H_{0}\leq{\mathrm{H}}({{\boldsymbol{\lambda}}}), then

|λ+j=0m2λjfj(β)|v0>C(β,ε)Hv0(𝝀)H(𝝀)μ(β,ε),\displaystyle\left|{{\lambda}}+\sum_{j={0}}^{m-2}{{\lambda_{j}}}f_{j}(\beta)\right|_{v_{0}}>C(\beta,\varepsilon){\mathrm{H}}_{v_{0}}({{\boldsymbol{\lambda}}}){\mathrm{H}}({{\boldsymbol{\lambda}}})^{-\mu(\beta,\varepsilon)}\kern 5.0pt,

where

μ(β,ε)=𝔸v0(β)+Uv0(β)Vv0(β)ϵ,\displaystyle\mu(\beta,\varepsilon)=\dfrac{\mathbb{A}_{v_{0}}(\beta)+{{U}}_{v_{0}}(\beta)}{V_{v_{0}}(\beta)-\epsilon},
C(β,ε)=exp((log(2)Vv0(β)ε+1)(𝔸v0(β)+Uv0(β)).\displaystyle C(\beta,\varepsilon)=\exp\left(-{{\left(\frac{\log(2)}{V_{v_{0}}(\beta)-\varepsilon}+1\right)}}(\mathbb{A}_{v_{0}}(\beta)+{{U}}_{v_{0}}(\beta)\right).
Proof..

Firstly, we claim that MnM_{n} is invertible by applying Theorem 3.1. To this end, we show that equations (18) and (19) in Theorem 3.1 hold for our a(z),b(z)a(z),b(z).

First, we claim that the assumptions (1) and (2) imply (18). Suppose not. Then there exist a positive integer nn and a root αi\alpha_{i} of a(t)a(t) such that

na(αi)+b(αi)=0.na^{\prime}(\alpha_{i})+b(\alpha_{i})=0.

The assumption (1) implies a(αi)0a^{\prime}(\alpha_{i})\neq 0, and the above equality yields

b(αi)a(αi)=n1.\dfrac{b(\alpha_{i})}{a^{\prime}(\alpha_{i})}=-n\in\mathbb{Z}_{\leq-1}.

This contradicts the assumption (2). Moreover, the assumption (3) implies that equation (19) holds. Therefore, Theorem 3.1 ensures that detMnK{0}\det M_{n}\in K\setminus\{0\}.

For v𝔐Kv\in\mathfrak{M}_{K}, we define functions Fv:0F_{v}:\mathbb{N}\longrightarrow\mathbb{R}_{\geq 0} by

Fv(n)\displaystyle F_{v}(n) =n(i=1mhv(αi)+(m1)(hv(𝜶)+hv(β)))\displaystyle=n\left(\sum_{i=1}^{m}{\rm{h}}_{v}(\alpha_{i})+(m-1)({\rm{h}}_{v}(\boldsymbol{\alpha})+{\rm{h}}_{v}(\beta))\right)
+{mi=1mlog|μn(si)|v1+log|d(m1)(n+1)(bm1)|v+o(n)ifv,mlog|4|v+o(n)ifv,\displaystyle+\begin{cases}m\sum_{i=1}^{m}\log|\mu_{n}(s_{i})|_{v}^{-1}+\log|d_{(m-1)(n+1)}(b_{m-1})|_{v}+o(n)&\ \text{if}\ v\nmid\infty,\\ m\log|4|_{v}+o(n)&\ \text{if}\ v\mid\infty,\end{cases}

where o(n)=0o(n)=0 for almost all non-Archimedean places vv. Notice that

lim supn1nFv0(n)=Uv0(β).\limsup_{n\to\infty}\dfrac{1}{n}F_{v_{0}}(n)=U_{v_{0}}(\beta).

Proposition 4.3 allows us to get

max0jm20m1logmax{|Pn,j(β)|v,|Qn,j,(β)|v}Fv(n).\displaystyle\max_{\begin{subarray}{c}0\leq j\leq m-2\\ 0\leq\ell\leq m-1\end{subarray}}\log\,\max\{|P_{n,j}(\beta)|_{v},|Q_{n,j,\ell}(\beta)|_{v}\}\leq F_{v}(n).

Then Proposition 4.13 yields

max0jm20m1log|n,j,(β)|v0𝔸v0(β)n+o(n).\displaystyle\max_{\begin{subarray}{c}0\leq j\leq m-2\\ 0\leq\ell\leq m-1\end{subarray}}\log\,|\mathfrak{R}_{n,j,\ell}(\beta)|_{v_{0}}\leq-\mathbb{A}_{v_{0}}(\beta)n+o(n).

Lemma 4.4 (iii)(iii) yields

lim supn1nlogd(m1)(n+1)(bm1)=(m1)den(bm1)φ(den(bm1))1jden(bm1)(j,den(bm1))=11j,\displaystyle\limsup_{n\to\infty}\dfrac{1}{n}\log d_{(m-1)(n+1)}(b_{m-1})=(m-1)\dfrac{{\rm{den}}(b_{m-1})}{\varphi({\rm{den}}(b_{m-1}))}\sum_{\begin{subarray}{c}1\leq j\leq{{\rm{den}}(b_{m-1})}\\ (j,{\rm{den}}(b_{m-1}))=1\end{subarray}}\dfrac{1}{j},

and we obtain

𝔸v0(β)lim supn1nvv0Fv(n)Vv0(β),\mathbb{A}_{v_{0}}(\beta)-\limsup_{n}\dfrac{1}{n}\sum_{v\neq v_{0}}F_{v}(n)\leq V_{v_{0}}(\beta),

where Vv0(β)V_{v_{0}}(\beta) is the real number defined in (4). Using a qualitative linear independence criterion in [9, Proposition 5.65.6] for

ϑj=fj(β)for 0jm2,\vartheta_{j}=f_{j}(\beta)\ \ \text{for}\ \ 0\leq j\leq m-2,

and the family of invertible matrices (Mn)n(M_{n})_{n}, and applying above estimates, we obtain Theorem 1.1. ∎

6 Appendix: Jordan-Pochhammer equation

The equation of the following form is called the Jordan-Pochhammer equation (confer [19, 18.418.4]):

(46) i=0m(μi)Q(i)(z)dmidzmij=0m1(μ1j)R(j)(z)dm1jdzm1j\displaystyle\sum_{i=0}^{m}\binom{-\mu}{i}Q^{(i)}(z)\dfrac{d^{m-i}}{dz^{m-i}}-\sum_{j=0}^{m-1}\binom{-\mu-1}{j}R^{(j)}(z)\dfrac{d^{m-1-j}}{dz^{m-1-j}}

where μ\mu is a complex number, (μ0)=1\binom{\mu}{0}=1 and

(μi)=μ(μ1)(μi+1)i!fori1,\displaystyle\binom{\mu}{i}=\dfrac{\mu(\mu-1)\ldots(\mu-i+1)}{i!}\ \ \text{for}\ \ i\geq 1,
Q(z)=(zα1)(zαm)[z]with disinct roots,\displaystyle Q(z)=(z-\alpha_{1})\ldots(z-\alpha_{m})\in\mathbb{C}[z]\ \ \text{with disinct roots},
R(z)[z]of degree at mostm1.\displaystyle R(z)\in\mathbb{C}[z]\ \text{of degree at most}\ m-1.

We observe that the equation (46) is of Fuchsian type with singularities α1,,αm,\alpha_{1},\ldots,\alpha_{m},\infty, and has the following Riemann scheme:

{α1αm00μ111μ2m2m2μm+1m+μ+s11m+μ+sm1μγ},\begin{Bmatrix}\alpha_{1}&\cdots&\alpha_{m}&\infty\\ 0&\cdots&0&-\mu-1\\ 1&\cdots&1&-\mu-2\\ \vdots&\ddots&\vdots&\vdots\\ m-2&\cdots&m-2&-\mu-m+1\\ m+\mu+s_{1}-1&\cdots&m+\mu+s_{m}-1&-\mu-\gamma\end{Bmatrix},

where si:=R(αi)/Q(αi)s_{i}:=R(\alpha_{i})/Q^{\prime}(\alpha_{i}) for 1im1\leq i\leq m and γ=s1++sm\gamma=s_{1}+\ldots+s_{m}.

Example 6.1.

Let m2m\geq 2 be an integer and a(z),b(z)[z]a(z),b(z)\in\mathbb{C}[z], where a(z)a(z) is a monic polynomial of degree mm with distinct roots, and b(z)b(z) is a polynomial of degree at most m1m-1. Denote the differential operator

L=dm1dzm1(a(z)ddzb(z)).L=\dfrac{d^{m-1}}{dz^{m-1}}\left(a(z)\dfrac{d}{dz}-b(z)\right).

Then, the following identity holds:

L\displaystyle L =i=0m(mi)a(i)(z)dmidzmij=0m1(m1j)(a(z)+b(z))(j)dm1jdzm1j.\displaystyle=\sum_{i=0}^{m}\binom{m}{i}a^{(i)}(z)\dfrac{d^{m-i}}{dz^{m-i}}-\sum_{j=0}^{m-1}\binom{m-1}{j}(a^{\prime}(z)+b(z))^{(j)}\dfrac{d^{m-1-j}}{dz^{m-1-j}}.

Thus, LL takes a form of a Jordan-Pochhammer equation with Q(z)=a(z),R(z)=a(z)+b(z)Q(z)=a(z),R(z)=a^{\prime}(z)+b(z) and μ=m\mu=-m.

Acknowledgements.

The author thanks to Akihito Ebisu for his invaluable comments for Jordan-Pochhammer equation. This work is partially supported by the Research Institute for Mathematical Sciences, an international joint usage and research center located at Kyoto University. The author is supported by JSPS KAKENHI Grant Number JP24K16905.

References

  • [1] Y. André, GG-functions and geometry, Aspects of Mathematics, E13. Friedr. Vieweg & Sohn, Braunschweig, 1989.
  • [2] Y. André, Séries Gevrey de type arithmétique, I. Théorèmes de pureté et de dualité, Ann. of Math. 151 (2000), 705-740.
  • [3] Y. André, Arithmetic Gevrey series and transcendence. A survey, J. Théor. Nombres Bordeaux 15 (1) (2003), 1–10.
  • [4] A. I. Aptekarev, A. Branquinho and W. Van Assche, Multiple orthogonal polynomials for classical weights, Trans. Amer. Math. Soc. 355. 10 (2003), 3887-3914.
  • [5] P. Bateman, J. Kalb and A. Stenger, Problem 1079710797: A limit involving least common multiples, Amer. Math. Monthly 109 (2002), 393–394.
  • [6] F. Beukers, Irrationality of some pp-adic LL-values, Acta Math. Sin. 24, no. 4, (2008), 663–686.
  • [7] D. V. Chudnovsky, G. V. Chudnovsky, Padé and rational approximations to systems of functions and their arithmetic applications, Lecture Notes in Mathematics, volume 1052, 1984, 37–84.
  • [8] D. V. Chudnovsky, G. V. Chudnovsky, Use of computer algebra for Diophantine and differential equations, In: Computer Algebra (New York, 1984), Lecture Notes in Pure and Applied Math. 113, Marcel Dekker, 1989, 1–81.
  • [9] S. David, N. Hirata-Kohno and M. Kawashima, Can polylogarithms at algebraic points be linearly independent?, Mosc. J. Comb. Number Theory 9 (2020), 389–406.
  • [10] S. David, N. Hirata-Kohno and M. Kawashima, Linear Forms in Polylogarithms, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 23 (2022), 1447–1490.
  • [11] S. David, N. Hirata-Kohno and M. Kawashima, Linear independence criteria for generalized polylogarithms with distinct shifts, Acta Arith. 206 (2) (2022), 127–169.
  • [12] S. David, N. Hirata-Kohno and M. Kawashima, Linear independence of values of hypergeometric functions and arithmetic Gevrey series, available at https://arxiv.org/pdf/2511.06534
  • [13] S. Fischler and T. Rivoal, On the denominators of the Taylor coefficients of GG-functions, Kyushu J. Math. 71.2 (2017), 287–298.
  • [14] S. Fischler and T. Rivoal, A note on GG-operators of order 22, Colloq. Math. 170.2 (2022), 321–340.
  • [15] C. Hermite, Sur la fonction exponentielle, C.r. Acad. Sci. Paris, 77, 1873, 18–24, 74–79, 226–233, 285–293. Oeuvres m, 150-181.
  • [16] C. Hermite, Lettre de M. C. Hermite de Paris à M. L. Fuchs de Göttingue “Sur quelques équations différentielles linéaires” ,Journal de Crelle, 79, 1875, 324–338.
  • [17] M. Huttner, On linear independence measure of some abelian integrals, Kyusyu J. Math, 57, 2003, 129–157.
  • [18] M. Huttner, On a paper of Hermite and Diophantine Approximation of Abelian Integrals ((Analytic Number Theory :: Expectations for the 21st Century)), RIMS Kokyuroku, 1219, 2001, 185–194.
  • [19] E. L. Ince, Ordinary differential equations, Dover, 1956.
  • [20] M. Kawashima, Rodrigues formula and linear independence for values of hypergeometric functions with parameters vary, J. of the Aust. Math. Society, 117, 3, 308–344.
  • [21] M. Kawashima and A. Poëls, Padé approximation for a class of hypergeometric functions and parametric geometry of numbers, J. of Number Theory, 243, (2023), 646–687.
  • [22] M. Kawashima and A. Poëls, On the linear independence of pp-adic polygamma values, Mathematika 71 (4) (2025), DOI: 10.1112/mtk.70040 .
  • [23] E. M. Nikišin and V. N. Sorokin, Rational Approximations and Orthogonality, American Math. Soc., Translations of Mathematical Monographs, 92 (1991).
  • [24] H. Padé, Sur la représentation approchée d’une fonction par des fractions rationnelles, Ann. Sci. École Norm. Sup. 9 (1892), 3–93.
  • [25] H. Padé, Mémoire sur les développements en fractions continues de la fonction exponentielle, pouvant servir d’introduction à la théorie des fractions continues algébriques, Ann. Sci. École Norm. Sup. 16 (1899), 395–426.
  • [26] O. Perron, U¨\ddot{U}ber Summengleichungen and Poincarésche Differenzengleichungen, Math. Annalen., 84, 1–15 (1921).
  • [27] M. Pituk, More on Poincaré’s and Perron’s theorems for difference equations*, J. Differ. Equ. Appl., 8(3), 201–216 (2002).
  • [28] G. Rhin and P. Toffin, Approximants de Padé simultanés de logarithmes, J. Number Theory, 24, (1986), 284–297.
  • [29] C. L. Siegel, U¨\ddot{\text{U}}ber einige Anwendungen diophantischer Approximationen, Abh. Preuß. Akad. Wiss., Phys.-Math. Kl. (1), 1929, 70S, English transl. in On Some Applications of Diophantine Approximations, (with a commentary by C. Fuchs and U. Zannier), Quad. Monogr. 2, Edizioni della Normale, Pisa, 2014, 1–80.
 

Makoto Kawashima kawasima@mi.meijigakuin.ac.jp Institute for Mathematical Informatics Meiji Gakuin University Totsuka, Yokohama, Kanagawa 224-8539, Japan